content
stringlengths 1
15.9M
|
---|
\section{Introduction}
\end{Large}
\end{center}
\subsection{Motivations}
The present work is an attempt to give an appropriate definition of tropical convergence for abstract families of curves in terms of their moduli, with a view towards applications to the constructive and enumerative aspects of tropical geometry. In particular, one of the goals of this paper is to provide an alternative proof of Mikhalkin's theorem for approximation of phase tropical curves as stated in \cite{Mikh06}, with references for various proofs.\\
A family of immersed curve $ \mathcal{C}_t \subset ( \mathbb{C}^{\ast} )^{2}$ is classically said to converge to a tropical curve $C \subset \mathbb{R}^2$ if the family of rescaled amoebas $$ \dfrac{1}{\log (t)} \mathcal{A} ( \mathcal{C}_t) \subset \mathbb{R}^2 $$ converges in Hausdorff distance to $C$, when $t$ becomes arbitrarily large.
If $\left\lbrace S_t \right\rbrace_{t>>1} \subset \mathcal{M}_{g,n}$ is the family of abstract curves parametrizing $\left\lbrace \mathcal{C}_t \right\rbrace_{t>>1}$, what can be said about the moduli of $S_t$ when $t$ is arbitrarily large?
Conversely, how can we construct, among all immersed curves of degree $deg \, C$, a family of curves approximating $C$ by prescribing their moduli?\\
We will go beyond the algebraic settings and answer both questions in a wider context to eventually come back to the algebraic case in the last section.
\subsection{Main results}
\textbf{Harmonic amoebas and harmonic tropical curves.} Amoebas have been introduced in the now classical \cite{GKZ}. They are defined as images of algebraic varieties in complex tori by the amoeba map
\[
\begin{array}{rcl}
\mathcal{A} \: : \: \big( \mathbb{C}^\ast \big)^n & \rightarrow & \mathbb{R}^n \\
(z_1,...,z_n) & \mapsto & \big( \log \vert z_1 \vert, ..., \log \vert z_n \vert \big)
\end{array}.
\]
For an immersed curve $S \rightarrow \big( \mathbb{C}^\ast \big)^n$, the map $ \mathcal{A}$ is given on $S$ by integrating the 1-forms $Re \big( d \log z_j \big)$. These are special instances of the so-called imaginary normalized differentials (i.n.d. for short), namely holomorphic differentials on $S$ with at worst simple poles at the punctures and purely imaginary periods.\\
A theorem due to Riemann ensures that such differentials are in correspondence with the line vector of their residues at the $n$ punctures of $S$. Hence, a collection of $m$ residue vectors encoded in a real $m \times n$ matrix $R$ defines a map
$$\mathcal{A}_R : S \rightarrow \mathbb{R}^m$$
(up to translation) given by integration of the real parts of the induced i.n.d.'s on $S$. As the coordinate functions of $ \mathcal{A}_R$ are harmonic on $S$, we call the set $\mathcal{A}_R (S) \subset \mathbb{R}^m$ the \textbf{harmonic amoeba}\footnote{The present terminology is due to the author.} of $S$ with respect to $R$. We refer to the founding work \cite{Kri}, and the further details given below.\\
In the present text, we mimic Krichever's generalization of amoebas in the case of tropical curves. Following \cite{Mikh05}, tropical curves are graphs with 1-valent vertices removed, and equipped with a complete inner metric. Tropical morphisms on them are integer affine linear map on any edge, such that their images are piecewise linear graph in $\mathbb{R}^m$ with integer slope and satisfying the so-called \textbf{balancing condition}. Tropical morphisms can be alternatively described by integration of exact 1-forms on tropical curves. Following \cite{MZ}, a 1-form $\omega$ on a tropical curve $C$ is roughly speaking the data of a current on $C$, seen as an electrical circuit, where the leaves of $C$ are either electrical sources or sinks (positive or negative residues of $\omega$). \\
We show that exact 1-forms on a tropical curve $C$ are in correspondence with the line vector of their residues at the $n$ leaves of $C$. As in the case of amoebas, an $m \times n$ residue matrix $R$ defines a map
$$\pi_R : C \rightarrow \mathbb{R}^m$$
(up to translation) given by integration of the induced forms on $C$. We call the set $\pi_R (C) \subset \mathbb{R}^m$ a \textbf{harmonic tropical curve} (parametrized by $C$). The current vectors on the edges of $C$ induced by $R$ should be thought of as a vectors of tropical periods.
Then, harmonic amoebas ($resp.$ harmonic tropical curves) are honest amoebas ($resp.$ tropical curves) when the periods of the 1-forms involved satisfy a certain integrality condition.\\
$\,$\\
\noindent
\textbf{Tropical convergence and approximation.}
The definition of tropical convergence presented here is expressed in terms of hyperbolic moduli, using Fenchel-Nielsen coordinates, see $e.g.$ \cite{B}. We formulate this coordinate-system as a surjective map
\[\mathcal{FN}_{G} \: : \: \big( \mathbb{C}^\ast\big)^{3g-3+n} \rightarrow \mathcal{M}_{g,n}\]
Here, $G$ is a cubic graph (only 1- and 3-valent vertices) of genus $g$, with $n$ leaves such that $2g-2+n>0$, and $G$ is implicitly equipped with a fixed ribbon structure. The parameters of $\mathcal{FN}_{G}$ are functions $(l,\Theta)$ that assign to any edge $e$ of $G$ a length $l(e) >0$ and a twist $\Theta(e) \in S^1$. From the latter data, we construct a Riemann surface $S_{(G,l,\Theta)}$ by gluing (hyperbolic) pair of pants such that the resulting decomposition is dual to $G$, the length of their boundary geodesics is given by $l$ and they are glued together along the edges of $G$ according to the twists $\Theta$. Note that the pair $(G,l)$ is an equivalent presentation of a simple tropical curve $C$.
\begin{definition}[\textbf{Tropical convergence}]\label{tropconv} $\,$\\
Let $C=(G,l)$ be a simple tropical curve of genus $g$ with $n$ leaves. A family of Riemann surfaces $ \left\lbrace S_t \right\rbrace_{t >> 1} \subset \mathcal{M}_{g,n} $ is said to converge to $C$ if for any $t$, there exist simple tropical curves $C_t=(G,l_t)$ and twist functions $\Theta_t$ such that $ S_t = S_{(C_t, \Theta_t)}$ and
$$ \displaystyle l_t(e) \underset{t\rightarrow \infty}{\sim} \frac{4\pi}{l(e) \log(t)}$$
for any non-leaf edge $e$ of $C$.
\end{definition}
\noindent
The geodesics of the decomposition of $S_t$ are vanishing cycles of the whole family, cycles of which we prescribe the speed of contraction. Then, a subsequence of $ \left\lbrace S_t \right\rbrace_{t >> 1} $ might converge to a dilatation of the original limiting tropical curve. The latter notion of convergence can also be understood in terms of Gromov-Hausdorff convergence toward the metric graph $C$, see below. Note that whenever $S_{(C_t,\Theta_t)}$ converges to $C$, $C_t$ does not converge to $C$. There is indeed a duality between the length of hyperbolic geodesics and the width of their collar.\\
Regarding our initial motivations, we provide the following answer.
\begin{mytheorem}\label{approxtrop}
Let $C$ be a simple tropical curve of genus $g$, with $n$ leaves and $R$ be an $m \times n$ matrix of residues.\\
For any sequence $ \left\lbrace S_t \right\rbrace_{t >>1} \subset \mathcal{M}_{g,n} $ converging to $C$, the sequence of maps
$ \big(1/\log (t)\big) \cdot \mathcal{A}_R : S_t \rightarrow \mathbb{R}^m $ converges to the map $\pi_R : C \rightarrow \mathbb{R}^m$.\\
Conversely, if $\pi_R$ is injective and $ \left\lbrace S_t \right\rbrace_{t >>1} \subset \mathcal{M}_{g,n} $ is a sequence of Riemann surfaces such that $\big(1/\log (t)\big) \cdot \mathcal{A}_R \big( S_t \big)$
converges in Hausdorff distance to $\pi_R \big(C\big) $, then $\left\lbrace S_t \right\rbrace_{t >>1} $ converges to $C$.
\end{mytheorem}
\noindent
Here, convergence is to be understood in the Hausdorff sense where it aims to describe convergence of mappings rather than simple convergence of their images, see definition \ref{GHconv}.\\
$\,$\\
\noindent
\textbf{Phase-tropical curves and Mikhalkin's theorem.}
Phase-tropical curves play a central role in the computation of planar Gromov-Witten invariants provided in \cite{Mikh05}. They are complexified version of tropical curves arising as Hausdorff limits of families of curves in complex tori, and have been studied intensively in the unpublished work \cite{Mikhunpub}.\\
They are topological surfaces equipped with a piecewise conformal structure and given as singular $S^1$-fibration over tropical curves. Simple phase-tropical curves can be constructed by gluing charts of ``the" phase-tropical line, itself fibered over ``the" tropical line. The transition functions over the edges of the underlying simple tropical curve $C$ are rotations between cylinders. Therefore, they can be encoded by twist parameters $\Theta$ on the non-leaf edges of $C$ (once a ribbon structure is fixed). Hence, they can be described by tropical Fenchel-Nielsen coordinates $(l,\Theta)$ and denoted by $V_{(C,\Theta)}$, where $C=(G,l)$.
\begin{definition}[\textbf{Phase-tropical convergence}]\label{comptropconv}$\,$\\
Let $C=(G,l)$ be a simple tropical curve of genus $g$ with $n$ leaves. A family of Riemann surfaces $ \left\lbrace S_t \right\rbrace_{t >> 1} \subset \mathcal{M}_{g,n} $ is said to converge to the simple phase-tropical curve $V_{(C,\Theta)}$ if for any $t$, there exist simple tropical curves $C_t=(G,l_t)$ and twist functions $\Theta_t$ such that $ S_t = S_{(C_t, \Theta_t)}$,
$$ \displaystyle l_t(e) \underset{t\rightarrow \infty}{\sim} \frac{4\pi}{l(e) \log(t)} $$
for any non-leaf edge $e$ of $G$ and $ \Theta_t $ converges to $\Theta$.
\end{definition}
\noindent
This refined notion of tropical convergence allows to control i.n.d.'s on sequences of Riemann surface. If $R$ is an $m \times n$ matrix of residue on $S \in \mathcal{M}_{g,n}$, one can consider the period matrix $ \mathcal{P}_{R,S}$ of the $m$ induced i.n.d.'s on a fixed basis of $H_1\big( S, \mathbb{Z} \big)$. The following theorem will be stated more precisely later.
\begin{mytheorem}\label{convperiod}
Assume $ \left\lbrace S_t \right\rbrace_{t>>1} \subset \mathcal{M}_{g,n} $ converges to a simple phase-tropical curve $V$. Then, for any $m \times n$ matrix of residue $R$, the associated sequence of periods matrices $\mathcal{P}_{R,S_t}$ converges to a matrix $\mathcal{P}_{R,V}$ depending only on $R$ and $V$.
\end{mytheorem}
\noindent
In the case the period matrices have coefficients in $2 \pi i \mathbb{Z} $ for any $t$, the maps $ \mathcal{A}_R$ factorize through immersions $\iota_R : S_t \rightarrow \big( \mathbb{C}^\ast \big)^m$. Deforming the complex structure of the complex torus via the map
\[
\begin{array}{rcl}
H_t \: : \: \big( \mathbb{C}^\ast \big)^m & \rightarrow & \big( \mathbb{C}^\ast \big)^m \\
(z_1,...,z_m) & \mapsto & \left( \vert z_1 \vert^{\frac{1}{\log(t)}} \frac{z_1}{\vert z_1 \vert} ,..., \vert z_m \vert^{\frac{1}{\log(t)}} \frac{z_m}{\vert z_m \vert} \right)
\end{array},
\]
the sequence of maps $ H_t \circ \iota_R : S_t \rightarrow \big( \mathbb{C}^\ast \big)^m $ converges to a so-called phase-tropical morphism $V \rightarrow \big( \mathbb{C}^\ast \big)^m$, $i.e.$ given by toric maps on the charts of $V$.\\
Mikhalkin's approximation theorem is about finding reasonable and general conditions under which tropical morphisms are obtained as limits of sequences $ H_t \circ \iota_R $. This condition is known as regularity and will be described below.
In the present framework, Mikhalkin's theorem amount to the existence of a family $\left\lbrace S_t \right\rbrace_{t>>1}$ converging to $V$ and having $2 \pi i$ integer period matrices with respect to the data $R$. The statement and its proof are given in section \ref{Mik}.
\subsection{Techniques and perspectives}
The key point in this paper is the study of the limit of the analytic linear system of i.n.d.'s on sequences of Riemann surfaces, in a spirit similar to \cite{GK}. Convergence of such linear systems along tropically converging sequence of curves can be understood in the classical Hodge bundle over $\overline{\mathcal{M}}_{g,n}$ (see \cite{ELSV}) twisted by the $n$ punctures. The control of the speed of the vanishing cycles (see definition \ref{tropconv}) allows to control the limiting period of i.n.d.'s along the same cycles (see theorem \ref{convdif}), whereas the refined notion of phase-tropical convergence allows to control all periods (see theorem \ref{convperiod}). The approximation theorems obtained here fall as corollaries.\\
The present techniques provide interesting connection with various compactifications of moduli spaces (curves, abelian varieties) as well as it suggest other possible compactifications.\\
One can extend the definition \ref{tropconv} of tropical convergence to any tropical curves, vertices of which can have any valency $v \geq 1$. Considering equivalence classes of tropical curves with respect to dilatation of the metric provides a proper definition of tropical convergence, meaning stable by taking subsequences. It induces a tropical compactification $\overline{\mathcal{M}}^{T}_{g,n}$ of $\mathcal{M}_{g,n}$. This construction is very similar to the one of Odaka, obtained by considering Gromov-Hausdorff collapse of Riemann surfaces, $cf$ \cite{O}. The resulting compactifications turns out to be topologically equivalent. However, the notions of convergence involved are different, see below.\\
Similarly, one can consider a wider class of phase-tropical curves supported on any tropical curves, see \cite{Mikhunpub}. Again, extending the definition \ref{comptropconv} to equivalence classes of such phase-tropical curves induces a compactification $\overline{\mathcal{M}}^{\varphi T}_{g,n}$ of $\mathcal{M}_{g,n}$, where forgetting the phases (or twists) gives rise to a map $\overline{\mathcal{M}}^{\varphi T}_{g,n} \twoheadrightarrow \overline{\mathcal{M}}^{ T}_{g,n}$. There seems to be an interesting connection between this compactification and the work of Thurston on surfaces in which he provides a compactification of the Teichmüller space $\mathcal{T}_g$ by considering measured foliation on surfaces, see $e.g.$ \cite{FLP}. As the action of the modular group extends to this compactifiaction, the quotient space provides a compactification of $\mathcal{M}_{g,0}$ that should be compared with $\overline{\mathcal{M}}^{\varphi T}_{g,0}$.\\
Eventually, the techniques developed here suggest a compactification of the moduli space $A_g$ of principally polarized abelian varieties (ppav). Indeed, the asymptotic of Jacobians can be totally determined along phase-tropically convergent sequences of Riemann surfaces. This asymptotic can be described in terms of period matrices according to an appropriate basis of 1-cycles. The imaginary part of such matrices collapse to 0 at a logarithmic speed that can be explicitly determined, whereas their real part defines tropical jacobians as studied in \cite{MZ} and \cite{BMV}. It suggests then a phase-tropical compactification of the moduli space $A_g$ as well as a compactification of the Torelli map that should be compared with the work \cite{Al}, \cite{BMV}, \cite{CaVi} and \cite{O}.\\
It also motivate a continuation of the work done in \cite{MZ} by studying divisors and 1-forms on phase-tropical curves.
Some of the latter perspectives will be discussed throughout the rest of the paper and will be the subject of the further work \cite{Lappear}.
\vspace{1cm}
\tableofcontents
\newpage
\section{Prerequisites}
\subsection{Imaginary normalized differentials and harmonic amoebas}
In this section, one briefly overviews the facts of main interest for us that appear in \cite{Kri}. Unless specified otherwise, the proofs of the statements to follow can be found there.
\begin{definition}
Let $n$ and $g$ be natural numbers and $S \in \mathcal{M}_{g,n}$ a Riemann surface. An \textbf{imaginary normalized differential} (\textbf{i.n.d.} for short) $\omega$ on $S$ is an holomorphic differential on $S$ having simple poles at the $n$ punctures of $S$ and such that
\[ Re \left( \int_{\gamma} \omega \right) = 0 \]
for any $ \gamma \in H_1 (S, \mathbb{Z})$.
\end{definition}
\begin{theorem}\label{thmRiem}
Let $n$, $g$ and $S \in \mathcal{M}_{g,n}$ be as in the previous definition and denote by $p_1, ..., p_n$ the $n$ punctures of $S$. For any collection of real numbers $r_1, ..., r_n$ such that $\sum_j r_j =0$, there exists a unique imaginary normalized differential $\omega$ on $S$ such that $Res_{p_j} \omega = r_j$ for any $1 \leq j \leq n$.
\end{theorem}
\noindent
\begin{proof}
See e.g. theorem 5.3 in \cite{Lang82}.
\end{proof}
\begin{definition}\label{defAr}
A \textbf{collection of residues of dimension $m$} on a Riemann surface $S \in \mathcal{M}_{g,n}$ is real $n \times m$ matrix $R:= \big( r^{(k)}_j \big)_{k,j}$ such that $\sum_j r_j^{(k)} =0$ for any $1 \leq k \leq m$.
Denote by $\omega_{R,S} := (\omega_{R,S}^{(1)}, ..., \omega_{R,S}^{(m)})$ the collection of imaginary normalized differentials on $S$ induced by the $m$ lines of $R$, see theorem \ref{thmRiem}. Finally, define the \textbf{harmonic amoeba map}
\[
\begin{array}{rcl}
\mathcal{A}_R \; : \; S & \rightarrow & \mathbb{R}^m\\
z & \mapsto & \left( Re \big( \int^z_{z_0} \omega_{R,S}^{(1)} \big),..., Re \big( \int^z_{z_0} \omega_{R,S}^{(m)} \big) \right)
\end{array}.
\]
up to the choice of an initial point $z_0 \in S$.
\end{definition}
\noindent
One also introduces here the following terminology
\begin{definition}
Let $S \in \mathcal{M}_{g,n}$ and $R$ be a collection of residues of dimension $m$. The set $\mathcal{A}_R (S) \subset \mathbb{R}^m$ is the \textbf{harmonic amoeba} of $S$ with respect to $R$.
\end{definition}
\begin{remark}
The space of deformation of harmonic amoebas of $S$ in $\mathbb{R}^m$ is a real vector space of dimension $m(n-1)$ given by the coefficients of the collection of residues $R$.
\end{remark}
The terminology is motivated by the fact that any coordinate function of the map $\mathcal{A}_R$ is harmonic on the punctured Riemann surface, and that the definition of harmonic amoebas generalizes the one of classical amoebas in the case of curves.
Indeed, for an immersion of $S$ in $(\mathbb{C}^\ast)^m$, the amoeba map $ \mathcal{A}$ is given coordinate wise by integrating the real part of the differentials $ d \log z_j $, where the $z_j$'s are the coordinate functions of the immersion. The periods of such differentials are always integer multiples of $2 \pi i$.\\
In the more general settings of i.n.d.'s, the periods can be any imaginary number.
One of the resemblance with the classical amoebas is the following.
\begin{proposition}\label{convamoeba}
Let $S \in \mathcal{M}_{g,n}$ and $R$ be a collection of residues of dimension 2. Then, all the connected components of the complement of $\mathcal{A}_R (S) \subset \mathbb{R}^2$ are convex, and the unbounded components are separated by tentacle-like asymptotes of $\mathcal{A}_R (S) $.
\end{proposition}
It is also shown in \cite{Kri} that harmonic amoebas possess a logarithmic Gauss map, a Ronkin function and extend many classical properties of amoebas. In particular, one can carry the definition of spine of planar amoebas as introduced in \cite{PR} to the case of harmonic amoebas. Such consideration motivates the introduction of a more general class of affine tropical curve with non rational slopes, namely harmonic tropical curves, see section \ref{secharmtrop}.
\begin{proposition}\label{areaamoeba}
Let $S \in \mathcal{M}_{g,n}$ and $R$ be a collection of residues of dimension 2. Then, the Euclidean area of $\mathcal{A}_R (S)$ in $\mathbb{R}^2$ is bounded from above by a constant $M_R$ independent of $S$.
\end{proposition}
For $S \in \mathcal{M}_{g,n}$, a basis $\gamma_1, ..., \gamma_{2g+n-1}$ of $H_1(S,\mathbb{Z})$ and a collection of residues $R$ of dimension $m$, the corresponding (normalized) period matrix $ \mathcal{P}_{R,S} $ is the $(2g+n-1) \times m$ matrix given by
\[ \left( \mathcal{P}_{R,S} \right)_{l,k} := \frac{1}{2 i \pi} \int_{\gamma_l} \omega_{R,S}^{(k)}. \]
The period matrix relative to a different basis of $H_1(S, \mathbb{Z})$, after the base change $A \in Sl_{2g+n-1}(\mathbb{Z})$ is given by $A \cdot \mathcal{P}_{R,S}$.
\begin{definition}\label{defpermat}
For $S \in \mathcal{M}_{g,n}$ and $R$ be a collection of residues of dimension $m$, the \textbf{period matrix} of $S$ with respect to $R$ is the equivalence class of $ \mathcal{P}_{R,S} $ by the left action of $Sl_{2g+n-1}(\mathbb{Z})$, and is still denoted by $ \mathcal{P}_{R,S} $. \\
$ \mathcal{P}_{R,S} $ is an \textbf{integer period matrix} if one of its representative (and then all of them) has all its entries in $\mathbb{Z}$. In such case, one can define the holomorphic map
\[
\begin{array}{rcl}
\mathcal{\iota}_{R} \; : \; S & \rightarrow & (\mathbb{C}^\ast)^m\\
z & \mapsto & \left( e^{\int^z_{z_0} \omega_{R,S}^{(1)}},..., e^{ \int^z_{z_0} \omega_{R,S}^{(m)}} \right)
\end{array}.
\]
\end{definition}
\begin{theorem}\label{thmcodim}
Let $S \in \mathcal{M}_{g,n}$ and $R$ be a collection of residues of dimension$ \; 1$. The level set
\[\left\lbrace S' \in \mathcal{M}_{g,n} \: \big| \: \mathcal{P}_{R,S} = \mathcal{P}_{R,S'} \right\rbrace\]
is a smooth analytic subvariety of $\mathcal{M}_{g,n}$ of real codimension $2g$.
\end{theorem}
\noindent
\begin{proof}
see remark 3.2 and theorem 2.6 in \cite{GK}.
\end{proof}
\subsection{A twisted Hodge bundle}
In the sequel, we will study the convergence of imaginary normalized differential over families of Riemann surfaces. In order to do so, we need an appropriate ambient space.\\
Recall that the classical Hodge bundle $\Lambda_{g,n} \rightarrow \overline{\mathcal{M}}_{g,n}$ is the rank $g$ vector
bundle whose fiber at a point $(S;p_1,...,p_n)$ is the space of holomorphic sections of the dualizing sheaf over $S$, see
\cite{ELSV}. Geometrically, the fiber of $\Lambda_{g,n}$ over a smooth curve $(S;p_1,...,p_n) \in \mathcal{M}_{g,n}$ is
the vector space of holomorphic differentials. Now for a singular curve $(S;p_1,...,p_n) \in
\overline{\mathcal{M}}_{g,n}$ with simple node $q_1$, ..., $q_k$, the fiber of $\Lambda_{g,n}$ over $S$ is the vector
space of meromorphic differentials such that their pullback on the normalization $\tilde{S}$ have at
most simple poles at the preimages of the nodes of $S$, and such that their residues at the 2 preimages of any of the
$q_j$'s are opposite to each other. Note that this vector space is also $g$-dimensional. The fact that $\Lambda_{g,n}$ is a vector bundle on $\overline{\mathcal{M}}_{g,n}$ can be found in \cite{ACG}, and references therein.
Now, consider the following twisted version of the latter bundle.
\begin{definition}
One defines $\Lambda_{g,n} ^m \rightarrow \overline{\mathcal{M}}_{g,n}$ to be the vector bundle of rank $g+n-1$ obtained by tensoring $\Lambda_{g,n}$ by $\mathcal{O}_S (p_1+...+p_n)$.
A point of $\Lambda_{g,n} ^m$ lying over a curve $S$ will be called a generalized meromorphic differential on $S$.
\end{definition}
A point of $\Lambda_{g,n} ^m$ is simply allowed to have extra simple poles at the punctures of $S$. Then, an i.n.d. $\omega$ on a Riemann surface $S \in \mathcal{M}_{g,n}$ can be considered as a point in the total space of $\Lambda_{g,n} ^m$.\\
Let us now introduce a local set of coordinates on the total space of $\Lambda_{g,n} ^m$. Convergence on $\mathcal{M}_{g,n}$ will be described in term of Fenchel-Nielsen
coordinates, so that we should only care about a coherent coordinate system on the fibers over the
sequence $\left\lbrace S_t \right\rbrace_{t >>1}$. If one restricts to a simply connected open subset $\mathscr{U} \subset \mathcal{M}_{g,n}$ such that $\left\lbrace S_t \right\rbrace_{t > t_0} \subset \mathscr{U}$ for some $t_0$, then $H_1(S,\mathbb{Z})$ can be trivialized on $\mathscr{U}$. In the sequel, we will consider sequences converging to maximally degenerated stable curves. Such sequence specifies $3g-3$ vanishing cycles in $H_1(S,\mathbb{Z})$. Now, choose $g$ of these cycles and $(n-1)$ small loops around all the punctures except one such that they are linearly independant in $H_1(S,\mathbb{Z})$. Denote them by $\gamma_1, ..., \gamma_{g+n-1}$, then computing the periods along those cycles gives a local system of coordinates
\[
\begin{array}{rcl}
\Lambda_{g,n} ^m \big( \mathscr{U} \big) & \rightarrow & \mathscr{U} \times \mathbb{C}^{g+n-1} \\
(S, \omega) & \mapsto & \Big( S, \, \big( \int_{\gamma_1} \omega, ..., \int_{\gamma_{g+n-1}} \omega \big) \Big)
\end{array}.
\]
on the fibers of $\Lambda_{g,n} ^m$ above $\mathscr{U}$.
\subsection{A bit of hyperbolic geometry}
In this subsection, one recalls some basic facts about geometry of hyperbolic surfaces, uniformization of punctured Riemann surfaces and Fenchel-Nielsen coordinates on moduli spaces of curves. One refers to \cite{B} or \cite{ACG} for the proofs.
\begin{definition}
A \textbf{pair of pants} $Y$ is a Riemannian surface with constant curvature -1 and boundary geodesics that is conformally equivalent to $ \mathbb{CP}^1 \setminus \big( E_1 \cup E_2 \cup E_3 \big) $ where each of the $E_i$'s is either a point or an open disc, and such that the $E_i$'s are pairwise disjoint.
\end{definition}
Recall that the upper half-space $ \mathbb{H} := \left\lbrace z \in \mathbb{C} \; \big| \; Im(z)>0 \right\rbrace $ equipped with the Riemannian metric
\[g:= \dfrac{(dx^2 +dy^2)}{y^2}\]
where $ z=x+iy$, is a complete metric space with constant curvature -1.
\begin{theorem}
For any $\alpha, \beta, \gamma >0$, there exists an hyperbolic right-angled hexagon $H \subset \mathbb{H}$ with consecutive side lengths $a, \alpha, b, \beta, c, \gamma$. The lengths $a,b,c$ are determined by $\alpha, \beta, \gamma$, and $H$ is unique up to isometry.
\end{theorem}
\begin{definition}
A generalized hyperbolic right-angled hexagon $H \subset \mathbb{H}$ is the limit for the Hausdorff distance on compact sets of a sequence of hyperbolic right-angled hexagons $ \left\lbrace H_t \right\rbrace_{t>>1} $ of respective lengths $\alpha_t, \beta_t, \gamma_t$ converging in $ \mathbb{R}_{\geq 0} $.
\end{definition}
From now on, a boundary components of a pair of pants $Y$ will be meant to be a boundary geodesic as well as a puncture of $Y$.
\begin{theorem}
For any pair of pants $Y$, there exists a unique orientation reversing isometry $\sigma$ fixing the boundary components. The quotient $Y/\sigma$ is isometric to a generalized hyperbolic right-angled hexagon $H \subset \mathbb{H}$.
\end{theorem}
\begin{definition}
For any pair of pants $Y$, we define $\mathbb{R} Y$ to be the fixed locus of the isometry $\sigma$ of the previous theorem.
\end{definition}
Now let us briefly recall what are Fenchel-Nielsen coordinates on the moduli space $\mathcal{M}_{g,n}$. These coordinates allow to describe any Riemann surface as a gluing of pair of pants, up to the choice of some combinatorial data we introduce right now.
\begin{definition}
A \textbf{cubic graph} $G$ is a graph with only 3-valent and 1-valent vertices. The edges adjacent to a 1-valent vertex are called leaves. \\
A \textbf{ribbon structure} $\mathscr{R}$ on $G$ is a the data for every vertex of $G$ of a cyclical ordering of its adjacent leaves-edges.
\end{definition}
\noindent
\textbf{Notations.}
We denote by $V(G)$ the set of vertices, by $L(G) $ the set of leaves and by $E(G)$ the set of non-leaf edges of $G$. We also denote $LE(G):= L(G) \cup E(G)$.\\
We fix the following data: a cubic graph $G$ with $n$ leaves and genus $g$, together with a fixed ribbon structure $\mathscr{R}$. A quick Euler characteristic computation shows that $G$ has $(3g-3+n)$ edges and $ (2g-2+n)$ 3-valent vertices. Assume the latter quantity is strictly positive.\\
Here, $S^1$ is the set of complex numbers of modulus 1. Given $l \, : \, E(G) \rightarrow \mathbb{R}_{> 0} $ (length function) and $\Theta \, : \, E(G) \rightarrow S^1$ (twist function), one constructs the Riemann surface $S_{(C,\Theta)} \in \mathcal{M}_{g,n}$ where $C:=(G,l)$:\\
first, $S_{(C,\Theta)}$ is given with the pair of pants decomposition dual to the graph $G$. The pair of pants $Y_v$ dual to the vertex $v$ is determined by the length of its boundary geodesics, prescribed by the length function $l$ on the dual edges. It remains to specify how pair of pants are glued together. We need to define a framing on each boundary geodesic of $Y_v$. We orient each geodesic such that the normal vector field points inwards. Then, we use the ribbon structure $\mathscr{R}$ to define an origin on each boundary geodesic of $Y_v$: $\mathscr{R}$ defines a orientation on $\mathbb{R} Y_v$ and then defines a first intersection point of $\mathbb{R} Y_v$ with every boundary geodesic, that we choose to be its origin. With this framing, every such geodesic is canonically isomorphic to the algebraic group $S^1$. Now, for 2 nearby vertices $v_1, v_2 \in V(G)$ connected by an edge $e\in E(G)$, glue the corresponding geodesics in $ Y_{v_1} $ and $ Y_{v_1} $ by the isometry
\[
\begin{array}{rcl}
S^1 & \rightarrow & S^1\\
z & \mapsto & -\overline{\Theta(e)z}.
\end{array}\]
We have just constructed the Riemann surface $S_{(C,\Theta)} \in \mathcal{M}_{g,n}$. Define the map
\[
\begin{array}{rcl}
\mathcal{FN}_{G} \: : \: \big( \mathbb{C}^\ast\big)^{3g-3+n} & \rightarrow & \mathcal{M}_{g,n} \\
(l,\Theta) & \mapsto & S_{(C,\Theta)}
\end{array}
\]
\begin{theorem}\label{thmhyperb}
The map $\mathcal{FN}_{G} \: : \: \big( \mathbb{C}^\ast\big)^{3g-3+n} \rightarrow \mathcal{M}_{g,n}$ is surjective.
\end{theorem}
\noindent
The 2 maps $l$ and $\Theta$ are known as Fenchel-Nielsen coordinates for the curve $S$, relatively to $G$ (and $\mathscr{R}$). For the sake of concision, we choose not to refer to $\mathscr{R}$ in the notation $S_{(C,\Theta)}$. It is an auxiliary datum that we fix one for all.
\begin{remark}
The latter map turns out to be an intermediate covering to the universal covering of $\mathcal{M}_{g,n}$ by the Teichmüller space $ \mathcal{T}_{g,n} \simeq \mathbb{R}^{6g-6+2n}$. Rigorously speaking, this map is not a coordinate system on $\mathcal{M}_{g,n}$ as it is far from being injective.
\end{remark}
\noindent
We end up this section by recalling some classical facts around the collar theorem.
\begin{definition}
Let $Y$ be a pair of pants. For any boundary geodesic $\gamma$ of $Y$, the \textbf{half-collar} associated to $\gamma$ is the tubular neighbourhood
\[K^{1/2}_\gamma := \left\lbrace z \in Y \; \big| \; d(z, \gamma) \leq w\big( l(\gamma) \big) \right\rbrace\]
where $ w\big( l(\gamma) \big) := arcsinh \left( 1 / \sinh\left( \frac{1}{2} l(\gamma)\right)\right)$. For any puncture $p$ of $Y$, the \textbf{cusp} $K_p$ associated to $p$ is the neighbourhood of $p$ in $Y$ isometric to the space $\left]- \infty, \log(2) \right] \times \mathbb{R}/\mathbb{Z}$ of coordinates $(\rho,t)$ with the metric $d\rho^2 + e^{2 \rho}dt^2$.
\end{definition}
\begin{proposition}\label{propcolcusp}
Let $\left\lbrace Y_t \right\rbrace_{t \in \mathbb{R}}$ be a sequence of pair of pants and $\gamma_t$ a boundary geodesic of $Y_t$ such that $\lim_{t \rightarrow \infty} l(\gamma_t)=0$. Denote by $p \in Y_\infty$ the puncture obtained by shrinking $\gamma_t$.
Then, the sequence of half-collars $\left\lbrace K^{1/2}_{\gamma_t} \right\rbrace_{t \in \mathbb{R}}$ converges in Gromov-Hausdorff distance to the cusp $K_p$.
\end{proposition}
\begin{theorem}
Let $Y$ be a pair of pants. The half-collars and cusps of $Y$ are pairwise disjoint.
\end{theorem}
\begin{definition}\label{popfinite}
Let $Y$ be a pair of pants. Define the \textbf{bounded part} $Y^{bd}$ to be the closure in $Y$ of the complement of all the cusps and half-collars of \nolinebreak$Y$.
\end{definition}
\begin{definition}\label{defcollar}
For any Riemann surface $S \in \mathcal{M}_{g,n}$, and any geodesic $\gamma \subset S$, the collar associated to $\gamma$ is the tubular neighbourhood
\[K_\gamma := \left\lbrace z \in S \; \big| \; d(z, \gamma) \leq w\big( l(\gamma) \big) \right\rbrace.
\]
\end{definition}
\subsection{Asymptotic equivalence of conformal invariants on holomorphic annuli}\label{secconfinv}
Unlike holomorphic discs, holomorphic annuli have a non trivial modulus. For example, the annuli $A_{r,R} = \left\lbrace z \in \mathbb{C} \: \big| \: r < \vert z \vert < R \right\rbrace$ are determined up to biholomorphism by the single conformal invariant $\log \left( \frac{R}{r}\right)$, see e.g. \cite{Rud}. This conformal invariant is related to so-called extremal lengths on complex domains, see e.g. \cite{Ahl}. It is a classical fact that any holomorphic annulus is biholomorphic to $A_{r,R}$ for some $0<r<R$. Hence, they have a single modulus. Another model can be chosen as
\[A_m = \left\lbrace z \in \mathbb{C} \: \big| \: 0 < Re( z ) < m \right\rbrace \big/ i \mathbb{Z},\]
for any $m>0$.
Choosing the rectangle of side $[0,m]$ and $[0,i]$ as a fundamental domain for $A_m$, the map $ z \mapsto e^{2i\pi z} $ sends $A_m$ onto $A_{1,e^m}$ and has modulus $m$. A consequence is that
\begin{center}
\parbox[c][][l]{10cm}{\textit{every holomorphic annulus is biholomorphic to a single $A_m$ for some $m>0$.}}
\end{center}
In the sequel, we will need to compare these conformal models with another model coming from hyperbolic geometry. Recall that the collar $K_l$ is the the tubular neighbourhood of width $w(l)= arcsinh\left( \frac{1}{\sinh (l/2)} \right)$ of a geodesic of length $l$ in an hyperbolic Riemann surface.
\begin{lemma}\label{lemmaconfcollar}
Let $m(l)$ be the unique positive number such that $K_l \simeq A_{m(l)}$. Then
\[ m(l) \underset{l \rightarrow 0}{\sim} \dfrac{2}{l} .\]
\end{lemma}
\noindent
\begin{proof}
According to \cite{B}, $K_l$ is isometric to $\left] - w(l), w(l) \right[ \times \mathbb{R}/\mathbb{Z}$ equipped with the metric $d\rho^2 + l^2 \cosh(\rho)^2 dy^2$. In order to present it as an annulus $A_{m(l)}$, let us look for a change of variable $x \mapsto \rho(x)$ for which the latter Riemannian metric becomes conformally equivalent to the euclidean metric on the $(x,y)$-space. It amounts to solve the differential equation
\[\rho'(x) = l \cosh \big(\rho(x) \big).\]
The solution is given by
\[\rho(x) = arccosh \left( \dfrac{1}{\cos(lx)}\right).\]
It follows that $K_l$ is isometric to $\left] - \rho^{-1}\big(w(l)\big), \rho^{-1}\big(w(l)\big) \right[ \times \mathbb{R}/\mathbb{Z}$ equipped with the metric $\left(\frac{l}{\cos(lx)}\right)^2 (dx^2 + dy^2)$. One deduces that $m(l)= 2\rho^{-1}\big(w(l)\big)$. It remains to compute the asymptotic of $m(l)$ when $l$ tends to $0$. One has
\[m(l)= \frac{2}{l}\arccos\left(\dfrac{1}{\cosh\big( w(l)\big)}\right). \]
Then, the result follows from the facts that $$\displaystyle \lim_{l\rightarrow 0} w(l)= + \infty \text{ and } \displaystyle \lim_{l\rightarrow 0} \arccos\left(\frac{1}{\cosh\big( w(l)\big)}\right) =1.$$
\end{proof}
Now, suppose we are given a sequence of holomorphic differentials $\left\lbrace \omega_m \right\rbrace_{m \in \mathbb{R}_{>0}}$ defined respectively on $A_m$, and normalized in the following way
\begin{eqnarray}\label{normdiff}
\dfrac{1}{2 \pi i} \int_0^i \omega_m = 1.
\end{eqnarray}
One would like to compute the asymptotic of the integral of $\omega_m$ along the width of $A_m$ while $m$ tends to $\infty$. For technical reasons, let us introduce the following definition
\begin{definition}
An admissible sequence of path is a sequence of piecewise $C^1$ maps $\rho_m \: : \: [0,m] \rightarrow A_m$ such that
\begin{description}
\item[i)] $Re \big( \rho_m (0) \big) =0$ and $Re \big( \rho_m (m) \big) =m$,
\item[ii)] $ \exists \: M>0 $ independent from $m$ such that $ \left| \int_0^m \frac{d}{dt}Im\big( \rho_m(t) \big) dt \right| <M$, $i.e.$ $\rho_m$ wraps at most finitely many times around $A_m$, independently on $m$.
\end{description}
\end{definition}
In order to determine the asymptotic of $\int_{\rho_m} \omega_m$ on $A_m$, one needs reasonable convergence assumptions on the sequence $\left\lbrace \omega_m \right\rbrace_{m \in \mathbb{R}_{>0}}$. Let us introduce a third model for holomorphic annuli. Let
\[C_\lambda = \left\lbrace (z,w) \in \mathbb{C}^2 \, \big| \; \vert z \vert < 1, \; \vert w \vert < 1, \: zw= \lambda \right\rbrace\]
for $0 \leq \lambda <1$. For $\lambda >0$, $C_\lambda$ is a holomorphic annulus of modulus $-\log(\lambda)$. For $\lambda=0$, $C_\lambda$ is the union of the $z$-unit disc and the $w$-unit disc in $\mathbb{C}^2$.
\begin{definition}\label{deflocconv}
Let $\left\lbrace \omega_m \right\rbrace_{m \in \mathbb{R}_{>0}}$ be a sequence of holomorphic differentials on the $A_m$'s. The sequence $\left\lbrace \omega_m \right\rbrace_{m \in \mathbb{R}_{>0}}$ converges when $m$ tends to $\infty$ if, in the $C_\lambda$-model, the limit is a meromorphic differential on each of the two irreducible components of $C_0$, with opposite residues.
\end{definition}
\begin{remark}
This convergence can be expressed in terms of point-wise convergence in total space of the relative dualizing sheaf of the morphism $(z,w) \mapsto zw$ from the bi-disc of bi-radius $(1,1)$ to the unit disc.
\end{remark}
\begin{lemma}\label{lemmaconfdiff}
Let $\left\lbrace \omega_m \right\rbrace_{m \in \mathbb{R}_{>0}}$ be a converging sequence of holomorphic differentials on the $A_m$'s and normalized as in \eqref{normdiff}. For any admissible sequence of paths $\left\lbrace \rho_m \right\rbrace_{m \in \mathbb{R}_{>0}}$, one has
\[\int_{\rho_m} \omega_m \underset{m \rightarrow \infty}{\sim} 2 \pi m.\]
\end{lemma}
\noindent
\begin{proof}
Consider first the sequence $\omega_m^0 = 2 \pi dz$ on $A_m$. This sequence is obviously normalized as in \eqref{normdiff}.
Moreover, it converges to the meromorphic differential given by $\pm dz/z$ and $ \mp dw/w$ on the 2 components of $C_0$ (the
indeterminacy of the signs holds on the choice of a canonical isomorphism between the $A_m$- and $C_\lambda$-models).\\
Consider now the admissible sequence of paths $\rho_m^0 (t)=t$. Then
\[\int_{\rho_m^0} \omega_m^0 = 2 \pi m.\]
Now for any converging sequence $\left\lbrace \omega_m \right\rbrace_{m \in \mathbb{R}_{>0}}$ normalized as in \eqref{normdiff}, the sequence $\left\lbrace \omega_m - \omega_m^0 \right\rbrace_{m \in \mathbb{R}_{>0}}$ converges to holomorphic differentials on the components of $C_0$. It follows that $\int_{\rho_m^0} ( \omega_m - \omega_m^0 ) $ converges and that
\[\int_{\rho_m^0} \omega_m \underset{m \rightarrow \infty}{\sim} 2 \pi m.\]
Now for any admissible sequence of paths $\left\lbrace \rho_m \right\rbrace_{m \in \mathbb{R}_{>0}}$, there exist 2 sequences of locally injective paths $\left\lbrace \rho_{1,m} \right\rbrace_{m \in \mathbb{R}_{>0}}$ and $\left\lbrace \rho_{2,m} \right\rbrace_{m \in \mathbb{R}_{>0}}$, respectively contained in the boundaries $ \left\lbrace Re(z)=0 \right\rbrace$ and $ \left\lbrace Re(z)=m \right\rbrace$ of $A_m$, and such that
\[ \rho_m \sim \rho_{2,m} \circ \rho_m^0 \circ \rho_{1,m}. \]
Hence
\[\int_{\rho_m} \omega_m = \int_{\rho_{1,m}} \omega_m + \int_{\rho_m^0} \omega_m + \int_{\rho_{2,m}} \omega_m .\]
By the convergence of $\left\lbrace \omega_m \right\rbrace_{m \in \mathbb{R}_{>0}}$ and the admissibility of $\left\lbrace \rho_m \right\rbrace_{m \in \mathbb{R}_{>0}}$, the integrals along $\rho_{1,m}$ and $\rho_{2,m}$ converge. As a consequence
\[\int_{\rho_m} \omega_m \underset{m \rightarrow \infty}{\sim} \int_{\rho_m^0} \omega_m \underset{m \rightarrow \infty}{\sim} 2 \pi m.\]
\end{proof}
\noindent
Putting everything together, one gets
\begin{proposition}\label{convdifloc}
Let $\left\lbrace l_t \right\rbrace_{t \in I}$ be a sequence of positive numbers such that
\[ l_t \underset{t \rightarrow \infty}{\sim} \dfrac{4 \pi}{l \cdot \log(t)} \]
for some positive number $l$. For any converging sequence $\left\lbrace \omega_t \right\rbrace_{t \in I}$ of holomorphic differentials on the $K_{l(t)}$'s, and any admissible sequence of paths $\left\lbrace \rho_t \right\rbrace_{t \in I}$, one has
\[\dfrac{1}{\log(t)} \int_{\rho_t} \omega_t \underset{t \rightarrow \infty}{\sim} l \Lambda\]
where
\[ \Lambda := \lim_{t \rightarrow \infty} \dfrac{1}{2 i \pi} \int_0^i \omega_t. \]
\end{proposition}
\noindent
\begin{proof} Denote $\Lambda_t := 1/(2 i \pi) \int_0^i \omega_t$. Suppose first that $\Lambda \neq 0$ and remove the finite number of $t$'s for which $\Lambda_t := 0$.
Applying lemma \ref{lemmaconfdiff} to the normalized family $\left\lbrace (1/\Lambda_t)\omega_t \right\rbrace_{t \in I}$, and the equivalence given in lemma \ref{lemmaconfcollar}, one gets
\[ \frac{1}{\Lambda_t} \int_{\rho_t} \omega_t \underset{t \rightarrow \infty}{\sim} 2 \pi \cdot m\big( l(t)\big) \underset{t \rightarrow \infty}{\sim} \frac{4 \pi}{l(t)} \underset{t \rightarrow \infty}{\sim} l\cdot \log(t).\]
The result obviously holds when $\Lambda = 0$.
\end{proof}
\subsection{Tropical curves}
In this section, one reviews tropical curves and tropical 1-forms on them. As abstract objects, the definition of tropical curves follows the classical framework, see $e.g.$ \cite{Mikh05}, \cite{Mikhunpub} and \cite{Viro}.
In the present paper, we restrict ourselves to simple tropical curves: firstly because it fits perfectly to the hyperbolic framework given previously, and secondly because considering general tropical curves would obscure vainly both statements and proofs. Finally, all the results in the present paper can be generalized to every tropical curve by density arguments, as simple tropical curves are dense in their respective moduli space, see \cite{Mikh05}.
\tocless\subsubsection{Tropical curves and morphisms}\label{sectropcurvmorph}
\begin{definition}\label{deftropcurv}
A \textbf{simple tropical curve} $C$ is a topological space homeomorphic to a cubic graph with all 1-valent vertices removed, and equipped with a complete inner metric. The leaves of a simple tropical curve are always marked.
\end{definition}
$\,$\\
\textbf{Notations} We carry the notation from $G$ to $C$ by making use of $V(C)$, $L(C)$, $E(C)$ and $LE(C)$.
For a vertex $v \in V(C)$, denote by $T_v \subset C$ the tripod obtained as the union of $v$ and the three open leaves-edges adjacent to $v$ and $T^{1/2}_v \subset T_v$ the tripod obtained by taking only half of the edges.
\begin{remark}\
A simple tropical curve can also be presented as a couple $C :=(G,l)$, where $G$ is the cubic graph supporting $C$ and $l : E(C) \rightarrow \mathbb{R}_{>0} $ is the length function of $C$.
\end{remark}
\begin{definition}\label{tropmorph}
Let $C $ be a simple tropical curve. A \textbf{tropical morphism} $ \pi \, : \, C \rightarrow \mathbb{R}^m $ is a continuous map subject to the following
\begin{description}
\item[\textbf{Integrality:}] for any $ e \in LE( C )$, the map $ \pi_{\vert_e} $ is integer affine linear, with respect to the metric on $e$.
\item[\textbf{Balancing:}] For any $ v \in V ( C )$, denote by $\vec{e}_1$, $\vec{e}_2$ and $\vec{e}_3 $ the 3 outgoing unitary tangent vectors to the 3 leaves-edges adjacent to $v$. Then
\[ \pi (\vec{e}_{1}) +\pi (\vec{e}_{2}) +\pi (\vec{e}_{3}) =0. \]
\end{description}
Tropical morphisms are tacitly considered up to isometry of the source.
\end{definition}
\begin{definition}
A \textbf{tropical curve} $\Gamma \subset \mathbb{R}^m$ is the image of tropical morphism $ \pi : C \rightarrow \mathbb{R}^m$.
\end{definition}
\begin{remark}
Tropical curves in affine spaces can be naturally equipped with positive integer weights, respecting to which they satisfy a balancing condition, see \cite{Mikh05}.
\end{remark}
\tocless\subsubsection{Tropical 1-forms}
The definition of tropical 1-forms below follows the one given in \cite{MZ} in the case of compact tropical curves. \\
\begin{definition}\label{def1form}
A \textbf{1-form} $\omega$ on a simple tropical curve $C$ is a locally constant 1-form on every $e \in LE( C )$ and satisfying the following
\begin{description}
\item[\textbf{Balancing:}] for any $v \in V( C )$, and $\vec{e_1},\vec{e_2},\vec{e_3}$ its 3 adjacent outgoing unitary tangent vectors, one has
\[w(\vec{e}_1) + w(\vec{e}_2) + w(\vec{e}_3)= 0. \]
\end{description}
The set of 1-forms on $C$ is denoted by $\Omega( C )$. For any inward oriented leaf $\vec{e}$ of $C$, define the number $w(\vec{e})$ to be the \textbf{residue} of $\omega$ at $e$. An element $\omega \in \Omega( C )$ is \textbf{holomorphic} if all its residues are zero. The set of holomorphic 1-form on $C$ is denoted by $ \Omega_\mathcal{H}( C )$.
\end{definition}
A 1-form $\omega$ on $C$ corresponds locally to the data $a \, dx$ on any leaf-edge $e$, where $a \in \mathbb{R}$ and $x \, : \, e \rightarrow \mathbb{R}$ is an isometric coordinate. For any other isometric coordinate $y$, one has $\omega = \pm a \, dy$ depending whether $x \circ y^{-1}$ preserves orientation or not. Therefore, $\omega$ is locally equivalent to the data of an orientation on any leaf-edge of $C$ plus the corresponding real number $a$. Globally, $\omega$ is equivalent to the datum of an electric current on $C$ seen as an electrical circuit, where the balancing condition corresponds to the conservation of energy at the node of the circuits.
\begin{definition}\label{pathloop}
Let $C$ be a simple tropical curve. A path $c$ in $C$ is an injective map $c \, : \, \llbracket 1, m \rrbracket \rightarrow LE(C)$, for some $m\in \mathbb{N}$, such that each $c(j)$ is oriented and such that the terminal vertex of $c(j)$ is the initial vertex of $c(j+1)$ for all $1 \leq j < m $. \\
A loop $\lambda$ in $C$ is an injective map $ \lambda \, : \, \mathbb{Z}/ m \mathbb{Z} \rightarrow E(C)$, for some $m\in \mathbb{N}$, such that each $c(j)$ is oriented and such that the terminal vertex of $c(j)$ is the initial vertex of $c(j+1)$ for all $j \in \mathbb{Z}/ m \mathbb{Z}$.\\
The leaves-edges of a path, or a loop $\rho$ in $C$ are oriented by definition. Hence, one can integrate any 1-form $\omega$ on $C$ along $\rho$, and one has the formula
\[\int_\rho \omega = \sum_{e\in \rho} l(e)\omega(\vec{e}) \; . \]
For a path from a leaf to another, or a loop $\rho$, one defines the 1-form $\omega^\rho$ dual to $\rho$ by
\[
\omega^\rho(\vec{e}) =
\left\lbrace
\begin{matrix}
1 & if \; \vec{e} \in Im(\rho) \\
0 & \text{for any other leaf-edge}
\end{matrix}
\right. .
\]
\end{definition}
\begin{definition}
An \textbf{exact 1-form} $\omega$ on a simple tropical curve $C$ is an element of $\Omega( C )$ such that $\int_\rho \omega = 0$
for any loop $\rho$ in $C$. The set of exact 1-forms on $C$ is denoted by $ \Omega_0( C )$.
\end{definition}
\begin{remark}
As for the classical terminology, exact 1-forms are exactly those obtained as gradient of particular functions on tropical curves, that we introduce later. It can be compared with one part of the tropical Abel-Jacobi theorem, see \cite{MZ}.
\end{remark}
\noindent
Before ending this subsection, we introduce some useful results on exact 1-forms.
\begin{proposition}\label{tdiff}
Let $C$ be a simple tropical curve of genus $g$ with $n$ leaves and assume $n\geq2$. Then, one has the following
\begin{description}
\item[1)] For any 1-form $\omega$, the sum of all its residues is zero.
\item[2)] $\Omega (C) = \Omega_0 (C) \oplus \Omega_\mathcal{H} (C)$ with respective dimension $(n-1)$ and $g$.
\item[3)] Any element of $\Omega_0 (C) $ is determined by its residues.
\end{description}
\end{proposition}
\noindent
\begin{proof} Assume $2g-2+n>0$, the remaining case being trivial. Pick one of the leaves of $C$ and choose $n$ paths going from this leaf to the $n$ remaining ones and denote by $\omega_1$, ..., $\omega_{n-1}$ the 1-forms dual to these paths, see definition \ref{pathloop}. Then, choose $g$ loops in $C$ linearly independent in $H_1(C,\mathbb{Z})$ and denote by $\omega_{n}$, ..., $\omega_{g+n-1}$ the 1-forms dual to these loops. Then, $\omega_1$, ..., $\omega_{g+n-1}$ are linearly independent as the dual cycles in $H_1(C,\partial C, \mathbb{Z})$ are independent. They also form a basis by the same argument. The residues of any element of this basis sum to zero, hence \textbf{1)} is proved. One check easily that $\omega_{n}$, ..., $\omega_{g+n-1}$ form a basis for $\Omega_\mathcal{H} (C)$, that $\Omega_0 (C) \cap \Omega_\mathcal{H} (C) = \emptyset$ and that $\Omega (C) = \Omega_0 (C) + \Omega_\mathcal{H} (C)$. Then, \textbf{2)} is proved. The last point is a consequence of the second. Details are left to the reader.
\end{proof}
\begin{definition}\label{defPir}
A \textbf{collection of residues of dimension $m$} on a simple tropical curve $C$ with $n$ leaves is real $n \times m$ matrix $R:= \big( r^{(k)}_j \big)_{k,j}$ such that $\sum_j r_j^{(k)} =0$ for any $1 \leq k \leq m$.
Denote by $\omega_{R,C} := (\omega^{(1)}_{R,C}, ..., \omega^{(m)}_{R,C})$ the collection of exact 1-forms on $C$ induced by the $m$ lines of $R$, see the previous proposition. Finally, define the map
\[
\begin{array}{rcl}
\pi_R \; : \; C & \rightarrow & \mathbb{R}^m\\
p & \mapsto & \left( Re ( \int^p_{p_0} \omega^{(1)}_{R,C}),..., Re ( \int^p_{p_0} \omega^{(m)}_{R,C} ) \right)
\end{array}
\]
up to the choice of an initial point $p_0 \in C$.
\end{definition}
\tocless\subsubsection{Phase-tropical curves}
\noindent
As in \cite{Mikh05}, consider the diffeomorphism
\[
\begin{array}{rcl}
H_t \: : \: ( \mathbb{C}^{\ast} )^{2} & \rightarrow & ( \mathbb{C}^{\ast} )^{2} \\
(z,w) & \mapsto & \left( \vert z \vert^{\frac{1}{\log(t)}} \frac{z}{\vert z \vert} , \vert w \vert^{\frac{1}{\log(t)}} \frac{w}{\vert w \vert} \right)
\end{array}.
\]
This corresponds to a change of the holomorphic structure of $( \mathbb{C}^{\ast} )^{2}$, see section 6 in \cite{Mikh05}. Note that
\[\mathcal{A} \circ H_t = \frac{1}{\log(t)} \mathcal{A}.\]
Denote $ \mathcal{L} := \left\lbrace (z,w) \in ( \mathbb{C}^{\ast} )^{2} \: \big| \: z+w+1=0 \right\rbrace$. The sequence of topological surfaces $\left\lbrace H_t(\mathcal{L}) \right\rbrace_{t >1}$ converges in Hausdorff distance to a topological surface $L$, when t goes to $\infty$, see for instance \cite{Mikh05}. Topologically, $L$ is a sphere minus 3 points. It is obtained as the gluing of 3 holomorphic annuli to the coamoeba of $\mathcal{L}$, as pictured in figure \ref{fig:L0}. It implies that $L$ comes naturally equipped with a piecewise conformal structure. Its amoeba $ \mathcal{A}(L)$ is the classical tropical line with vertex at the origin. Denote it by $\Lambda$. \\
\begin{figure}[h]
\centering
\input{L0.tex}
\caption{The fibration $ \mathcal{A} : L \rightarrow \Lambda$.}
\label{fig:L0}
\end{figure}
\begin{definition}
A simple phase-tropical curve is a topological surface $V$ together with a fibration $ \mathcal{A}_V : V \rightarrow C$ onto a simple tropical curve $C$, satisfying the following
\begin{itemize}
\item[$\ast$] for any $v \in V(C)$, and $T_v \subset C$ the associated tripod, there exist a tropical embedding $\pi_v : T_v \rightarrow \Lambda$ and a homeomorphism $\varphi_v : \mathcal{A}_V^{-1} \big( T_v\big) \hookrightarrow \nolinebreak[4] ( \mathbb{C}^{\ast} )^{2}$ such that the following diagram is commutative
\begin{center}
\begin{tikzpicture}
\matrix (m) [matrix of math nodes, row sep=1 em,
column sep= 1.5 em]
{ \mathcal{A}_V^{-1} \big( T_v\big) & & L. \\
& & \\
T_v & & \Lambda \\ };
\path[->,font=\scriptsize]
(m-1-1) edge node[above] {$ \varphi_v $} (m-1-3)
(m-3-1) edge node[above] {$ \pi_v $} (m-3-3)
(m-1-1) edge node[left] {$ \mathcal{A}_V $} (m-3-1)
(m-1-3) edge node[left] {$ \mathcal{A} $} (m-3-3);
\end{tikzpicture}
\end{center}
\item[$\ast$] For any pair of adjacent vertices $v_1, v_2 \in V(C)$, the transition function $ \varphi_{v_1} \circ \big( \varphi_{v_2}\big)^{-1} $ is biholomorphic.
\end{itemize}
\end{definition}
\begin{remark}
The preimage of $ \mathcal{A}_V$ over an edge $e$ is a holomorphic annulus of modulus $l(e)$. As transition functions are just rotations on such annuli, they preserve the fibers of the maps $ \varphi_{v_1} \circ \mathcal{A} $ and $ \varphi_{v_2} \circ \mathcal{A} $ where $v_1$ and $v_2$ are the two vertices adjacent to $e$. Hence, the fibration $ \mathcal{A}_V : V \rightarrow C$ is well-defined. Note also that the collection of $\pi_v$'s provides an atlas on the tropical curve $C$. Indeed, if each tripod $T_v$ is oriented outwards, each transition function is the only orientation reversing isometry on every edge.
\end{remark}
\textbf{Notations.} For the sake of concision, we denote a phase-tropical curve $ \mathcal{A}_V : V \rightarrow C$ simply by $V$.\\
Simple phase-tropical curves can be alternatively described by tropical Fenchel-Nielsen coordinates. Recall that the coamoeba of a curve $\mathcal{C} \subset ( \mathbb{C}^{\ast} )^{2}$ is the image of $\mathcal{C}$ by the argument map $ Arg : ( \mathbb{C}^{\ast} )^{2} \rightarrow T := \big( \mathbb{R} / 2 \pi \mathbb{Z} \big)^2 $ to the argument torus. The coamoeba $ Arg ( \mathcal{L} ) $ is the union of 2 open triangles delimited by 3 geodesics plus their 3 common vertices $ (\pi, 0)$, $(0, \pi)$ and $(\pi,\pi)$, see figure \ref{fig:coamoeba}.
\begin{figure}[h]
\centering
\scalebox{0.75}{
\input{coamoeba.tex}}
\caption{$ Arg ( \mathcal{L} ) $ (grey), and a framing on its 3 boundary geodesics (blue).}
\label{fig:coamoeba}
\end{figure}\\
The choice of a ribbon structure $\mathscr{R}$ on $C$ induces a framing on the boundary geodesics of $ \mathcal{A}_V^{-1}(v)$, for any $v \in V(C)$. This is done in the exact same fashion as the Fenchel-Nielsen construction of Riemann surfaces presented above. Once again, such framing induces a unique isomorphism with $S^1$ for each boundary geodesic.\\
Now, for any $e \in E(C)$ and $v_1$, $v_2$ its adjacent vertices, the holomorphic annulus $ \mathcal{A}_V ^{-1}(e)$ is mapped to a boundary geodesic of $ \mathcal{A}^{-1}\big(\pi_{v_1}(v_1)\big) $ $\big( resp$. of $ \mathcal{A}^{-1}\big(\pi_{v_2}(v_2)\big) \: \big)$ in $T$ by the map $Arg \circ \varphi_{v_1}$ ($resp$. $Arg \circ \varphi_{v_2}$). The transition function $\varphi_{v_1}\circ \big( \varphi_{v_2} \big)^{-1}$ descends to a self-inverse orientation reversing isometry
\[
\begin{array}{rcl}
S^1 & \rightarrow & S^1\\
z & \mapsto & -\overline{e^{i\theta}z}
\end{array}\]
from one geodesic to another. Hence, the choice of the ribbon structure $\mathscr{R}$ on $C$ induces a function $ \Theta : E(C) \rightarrow S^1$ by collecting the $e^{i\theta}$'s.
\begin{remark}
The building block $L$ for phase-tropical curves has a unique real structure. The three special points of $Arg ( L )$ correspond to the three connected components of $\mathbb{R} L$. Similarly to Riemann surfaces, the twist on an edge measures how much the local real structures on the two nearby building blocks fail to glue into a real structure on their union.
\end{remark}
\begin{proposition}
A phase-tropical curve $ \mathcal{A}_V : V \rightarrow C$ is uniquely determined by $C$ and its associated twist distribution $\Theta$, as constructed above.
\end{proposition}
\noindent
\begin{proof}
The phase-tropical curve $ \mathcal{A}_V : V \rightarrow C$ is determined by $C$ and the transition functions along its edges. The latter are easily recovered from $\Theta$.
\end{proof}
$\,$\\
\textbf{Notations.} The phase-tropical curve supported on $C$ with twist distribution $\theta$ will be denoted by $V_{(C, \theta)}$.
\section{Approximation of harmonic tropical curves}
\subsection{Convergence of imaginary normalized differentials}
The purpose of this section is to study the limit of imaginary normalized differentials on punctured Riemann surfaces along tropically converging families. Limits of linear systems on reducible curves are complicated in general, as addressed e.g. in section $5$ of \cite{GK}.
Thanks to the preliminary work on conformal invariants done in section \ref{secconfinv}, we will see that the definition of tropical convergence is the appropriate one that ensures the convergence of i.n.d.'s.\\
A collection of residues $R :=(r_1,...,r_n)$ and a sequence $ \left\lbrace S_t \right\rbrace_{t \in \mathbb{N}} \subset \mathcal{M}_{g,n} $ give rise to the sequence $ \left\lbrace ( S_t , \omega_{R,S_t}) \right\rbrace_{t \in \mathbb{N}} \subset \mathcal{M}_{g,n} $ in the total space of the twisted Hodge bundle $\Lambda_{g,n} ^m$, see definition \ref{defAr}. A simple tropical curve $C$ gives a point $S_{C} \in \overline{\mathcal{M}}_{g,n}$ simply by considering the stable curve dual to the underlying cubic graph of $C$. Now, any $\omega \in
\Omega (C)$ induces a generalized meromorphic differential on $S_{C}$: by duality between $C$ and $S_C$, the form $\omega$ restricted to any tripod $T^{1/2}_v \subset C$ induces a meromorphic differential on the corresponding component of $S_C$, a sphere with 3 punctures, by simply carrying the residues of $\omega_{\vert_{T^{1/2}_v}}$ to the corresponding punctures.
\begin{definition}\label{maphodge}
For a simple tropical curve $C$ and a 1-form $\omega \in \Omega(C)$, one denote by $ \big[ C,
\omega \big]$ the corresponding point in $ \Lambda_{g,n} ^m$.
\end{definition}
\noindent
The main result of this subsection is the following.
\begin{mytheorem}\label{convdif}
Let $ \left\lbrace S_t \right\rbrace_{t >>1} \subset \mathcal{M}_{g,n} $ be a sequence converging to a simple tropical curve $C$ and $R:=(r_1,...,r_n)$ be a collection of residues. Let $\left\lbrace \omega_{R,S_t} \right\rbrace_{t>>1} $ be the associated sequence of imaginary normalized differentials, see definition \ref{defAr}, and $\omega_{R,C} $ the associated 1-form on $C$, see definition \ref{defPir}. \\
Then, the sequence $\left\lbrace (S_t, \omega_{R,S_t}) \right\rbrace_{t >>1} \subset \Lambda_{g,n} ^m$ converges to $\big[ C, \omega_{R,C} \big]$.
\end{mytheorem}
Before proving theorem \ref{convdif}, let us introduce a technical definition : to a loop $ \rho_{ \mathbb{T}}$ in a cubic graph $G$, we will associate a piecewise geodesic loop in any curve $S = \mathcal{FN}_G(l ,\Theta)$. Recall that a loop is a map $ \rho_{ \mathbb{T}} : \mathbb{Z} / m \mathbb{Z} \rightarrow E(G)$. It can also be presented as a map $V : \mathbb{Z} / m \mathbb{Z} \rightarrow V(G)$ where $V(j)$ is the vertex between $\rho_{ \mathbb{T}} (j)$ and $\rho_{ \mathbb{T}} (j+1)$. Denote by $Y_{V(j)}$ the pair of pants in the decomposition of $S$ corresponding to the vertex $V(j)$, by $\gamma_j$ the geodesic in the decomposition of $S$ corresponding to the edge $\rho_{ \mathbb{T}} (j)$ and orient it such that the associated normal vector field points toward $Y_{V (j-1)}$.\\
Now construct the piecewise geodesic loop $\breve{\rho}$ in $S$ as follows : $\breve{\rho} \cap Y_{V(j)}$ is the oriented connected component of $ \mathbb{R} Y_{V(j)}$ going from $\gamma(j)$ to $\gamma(j+1)$, and $\breve{\rho} \cap \gamma_j$ is the positive arc of $\gamma_j$ connecting $\breve{\rho} \cap Y_{V(j-1)}$ to $\breve{\rho} \cap Y_{V(j)}$.
\begin{definition}\label{assloop}
For any loop $\rho_{ \mathbb{T}}$ in a cubic graph $G$ and any curve $S = \mathcal{FN}_G(l ,\Theta)$, define by $\breve{\rho} \subset S$ the loop in $S$ associated to $\rho_{ \mathbb{T}}$, as constructed above.
\end{definition}
\noindent
\begin{proof}{\textbf{of theorem \ref{convdif}.}}
Recall that from definition \ref{tropconv}, $S_t = \mathcal{FN}_G(l_t, \Theta_t)$. Suppose first that for any
geodesic $\gamma_t$ of the pair of pants decomposition of $S_t$, the sequence $\left\lbrace \int_{\gamma_t} \omega_{R,S_t}
\right\rbrace_{t>>1}$ converges. It is equivalent to the convergence of the sequence $\left\lbrace (S_t, \omega_{R,S_t}) \right\rbrace_{t>>1}$ in the total space of the bundle $ \displaystyle \Lambda_{g,n} ^m$. Let us show that the limit is $ \big[ C, \omega_{R,C} \big]$.\\
First, it is clear from the definition that the sequence of curves $S_t$ converges to the stable curve dual to the underlying graph $G$ of $C$. Now, pick $g$ loops $\rho_\mathbb{T}^{1}, ... , \rho_\mathbb{T}^g $ in $ C $ forming a basis of $H_1( C ,
\mathbb{Z} ) $. For any $1 \leq j \leq g$ and any $t$, construct a loop $\rho_t ^j \subset S_t$ as follows: as an
intermediate step, consider the piecewise geodesic loop $\breve{\rho}_t ^{j} \subset S_t$ associated to
$\rho_\mathbb{T}^{j}$ (see definition \ref{assloop}). Now define $\rho_t ^j$ to be the unique geodesic in the free
homotopy class of $\breve{\rho}_t ^j$, see theorem 1.6.6 in \cite{B}.
Now for $ 1 \leq j \leq g $, index coherently the edges $ \vec{e}^{\,jk} $ as they are encountered in the loop $
\gamma_\mathbb{T}^j$ for $ 1 \leq k \leq \lambda_j$, where the orientation on the $ \vec{e}^{\,jk} $'s is induced by $
\gamma_\mathbb{T}^j$ . Denote also by $ \gamma_t^{jk} $ the geodesic of the pair of pants decomposition of $S_t$
associated to $ \vec{e}^{\,jk} $ oriented coherently to the orientation of $ \vec{e}^{\,jk} $, and $K_t^{jk}$ the
collar around $ \gamma_t^{jk} $. Define $ \psi_t^{jk} \, : \, K_t^{jk} \rightarrow A_{m_t^{jk}} $ the
biholomorphism such that :
\begin{enumerate}
\item[$\ast$] $ \psi_t^{jk} \big( \rho_t^{\, j} \cap K_t^{jk} \big) $ is a transversal path in $ A_{m_t^{jk}} $,
\item[$\ast$] $ \psi_t^{jk} \big( \rho_t^{\, j} \cap K_t^{jk} \big) \cap \left\lbrace z \in A_{m_t^{jk}} \, \big| \: Re(z) =0 \right\rbrace = \left\lbrace 0 \right\rbrace$.
\end{enumerate}
Note that the push-forward of $\omega_{R,S_t}$ on $ A_{m_t^{jk}} $ by $ \psi_t^{jk} $ gives rise to a convergent sequence, according to definition \ref{deflocconv}. Indeed, the limit is given on each connected component of the normalization of $S_{C}$ by a unique meromorphic differential. This differential admits a unique representation as $f(z)dz$ once a coordinate $z$ is chosen.\\
It implies first that $\omega_{R,S_t}$ converges toward an holomorphic differential on the complement of the collars, which stay compact in the limit, thanks to proposition \ref{propcolcusp}. So, the integral on each complement converges to a finite quantity. Hence one has
\[ \int_{\rho_t^{\, j}} \omega_{R,S_t} = \sum_{k=1}^{\lambda_j} \int_{\rho_t^{\, j} \cap K_t^{jk}} \omega_{R,S_t} \quad + \; O(1).\]
Now, applying proposition \ref{convdifloc} for each of the collars involved in the latter formula, one gets
\begin{equation}\label{lim}
\lim_{t \rightarrow \infty} \frac{1}{\log(t)} \int_{\rho_t^{\; j}} \omega_{R,S_t}= \sum_{k=1}^{\lambda_j} l(e_{jk}) \Lambda_{jk}
\end{equation}
where $ \Lambda_{jk} = \displaystyle \lim_{t \rightarrow \infty} \frac{1}{2\pi i} \int_{\gamma_t^{jk}} \omega_{R,S_t} $ which exist by assumption. As $\omega_{R,S_t}$ is an imaginary normalized differential for any $t$, it implies that $ \Lambda_{jk} \in \mathbb{R} $ and that
\[ Re \left( \int_{\rho_t^{\; j}} \omega_{R,S_t} \right) = 0 , \; \forall t, \; \forall j.\]
Considering the real part on both sides in \eqref{lim} gives in turn that
\begin{equation}\label{lim2}
\sum_{k=1}^{m_j} l(e_{jk}) \Lambda_{jk} = 0, \; \forall j.
\end{equation}
By definition of the $ \Lambda_{jk} $, one can construct a 1-form on $C$ taking the value $\Lambda_{jk}$ along $\vec{e}_{jk}$ and having residue $r_l$ at the $l$-th leaf of $C$. The equation \eqref{lim2} states exactly that this 1-form belongs to $\Omega_0 (C) $. According to proposition \ref{tdiff}, this 1-form is exactly $\omega_{R,C}$ and the theorem is proved in this case. \\
Suppose now that there is a bound for all the periods $ \int_{\gamma_t} \omega_{R,S_t} $ that is uniform in $t$. For any subsequence $\left\lbrace t_k \right\rbrace_{t\in I} $ such that all these periods converge, one can reduce to the previous case and deduce that $ \omega_{R,S_{t_k}} $ converges to $ \omega_{R,C} $. As any converging subsequence converges to the same limit, the original sequence does converge. \\
Suppose finally that $ \lim_{t\rightarrow \infty} M_t = + \infty $ where $ M_t $ is the maximum of $ \left| \int_{\gamma_t} \omega_{R,S_t} \right| $ over the $\gamma_t$'s. Consider the sequence of imaginary normalized differential $ \tilde{\omega}_t := \frac{1}{M_t} \omega_{R,S_t} $. Considering a subsequence if necessary, assume that the periods $ \int_{\gamma_t} \tilde{\omega}_t $ converge. Applying the same argument as in the first case, one construct a limit element $ \tilde{\omega}^\mathbb{T} \in \Omega_0 (C ) $. The way we rescaled $ \omega_{R,S_t} $ to get $ \tilde{\omega}_t$ insures that
\begin{enumerate}
\item[$\ast$] $ \exists \; \vec{e} \in E(C) $ such that $ \vert \tilde{\omega}^{\mathbb{T}}(\vec{e}) \vert = 1 $,
\item[$\ast$] $ \tilde{\omega}^\mathbb{T} $ has no residues.
\end{enumerate}
In other word, $ \tilde{\omega}^\mathbb{T}$ is a non zero exact 1-form which is holomorphic. This is in contradiction with \ref{tdiff} and the theorem is proved.
\end{proof}
\subsection{Convergence of periods}
The refined notion of phase-tropical convergence allows to obtain a finer result than theorem \ref{convdif}.
\addtocounter{mytheorem}{-2}
\begin{mytheorem}
Let $ \left\lbrace S_t \right\rbrace_{t>>1} \subset \mathcal{M}_{g,n} $ be a sequence converging to a phase-tropical curve $V_{(C, \Theta)}$ and $R:= (r_1,...,r_n)$ be a collection of residues. Let $\left\lbrace \omega_{R,S_t} \right\rbrace_{t>>1} $ be the associated sequence of imaginary normalized differentials and $\omega_{R,C} $ the associated 1-form on $C$. \\
For any loop $\rho_{ \mathbb{T}} \subset C$, and $\breve{\rho}_t$ the associated loop in $S_t$, see definition \ref{assloop}, one has
\[ \lim_{t \rightarrow \infty} \int_{\breve{\rho}_t} \omega_{R,S_t} = \sum_{\vec{e} \in \rho_{ \mathbb{T}}} \log \big( \Theta (e) \big) \cdot \omega_{R,C} \big( \vec{e} \big)\]
where the branch of $\log$ is chosen such that $\log : S^1 \rightarrow \left[ 0, 2i \pi \right[ \subset \mathbb{C}$.
\end{mytheorem}
\addtocounter{mytheorem}{2}
\begin{remark} Together with theorem \ref{convdif}, the latter suggests that the notions of tropical convergence defined in \ref{tropconv} and \ref{comptropconv} are suitable for the study of limits of linear system on families of curves, as treated $e.g.$ in \cite{Ca} and \cite{GK}. As the definitions of tropical convergence also apply in the compact case ($g \geq 2$), the theorems \ref{convperiod} and \ref{convdif} can be adapted to this case by replacing i.n.d.'s by real parts of homolorphic differentials in order to determine the limit of Jacobians in tropical compactifications of moduli spaces of ppav's. For a phase-tropically converging sequence, one can determine the limit of its Jacobian more accurately. Fixing an appropriate symplectic basis of 1-cycles, we claim the period matrix $B_t$ along $B$-cycles has the following asymptotic
\[B_t \sim B_\mathcal{R} +\frac{i}{\log (t)} B_\mathcal{I} + o \Big( \frac{1}{\log (t)} \Big)\]
with $B_\mathcal{R}$ and $B_\mathcal{I}$ in $M_g(\mathbb{R})$, $B_\mathcal{I}>0$, are determined by the phase tropical curve $V_{(C,\Theta)}$ at the limit. As one could expect, the limit of the Jacobians given by the real part $B_\mathcal{R}$ is nothing but the tropical Jacobian $J(C)$, see \cite{MZ}. The knowledge of the asymptotic $B_\mathcal{I}$ suggests a phase-tropical compactification of $A_g$ which records this asymptotic, and which should be compared with the compactifications constructed in \cite{BMV} and \cite{O}. it will be part of the work in preparation \cite{Lappear}.
\end{remark}
In order to prove the above statement, let us briefly study imaginary normalized differentials $\omega$ on $S \in
\mathcal{M}_{0,3}$. We are interested in paths $\rho$ in $S$ for which $ Im(\omega)_{\vert_\rho} \equiv
0$. If ones denotes by $g$ the uniformizing metric on $S$, define $Im(\omega)^\vee$ ($resp.$ $Re(\omega)^\vee$) to be the vector field dual to $Im(\omega)$ ($resp.$ $Re(\omega)$) with respect to $g$. The vector field $Im(\omega)^\vee$ is
obtained by rotating $Re(\omega)^\vee$ by $\pi/2$. As a consequence, a path $\rho$ in $S$ is such that $
Im(\omega)_{\vert_\rho} \equiv 0$ if and only if $\rho$ is parallel to $Re(\omega)^\vee$ at
any point.\\
Let us identify $ S \simeq \mathbb{CP}^1 \setminus \left\lbrace -1, 1, \infty \right\rbrace$ and assume for a moment that none of the residues of $\omega$ is zero. Up to multiplication of $\omega$ by a constant and an automorphism of $S$ (exchanging the punctures), one can assume that
\begin{equation}\label{omega}
\omega = \left( \frac{\lambda_{1}}{z-1} + \frac{\lambda_{-1}}{z+1}\right) dz
\end{equation}
with $\lambda_{-1}, \lambda_1 >0$. Consider the real oriented blow-up $\tilde{S}$ of $S$ at $-1$, $1$ and $\infty$, and denote by $\gamma_{-1}$, $\gamma_1$ and $\gamma_\infty$ the its 3 boundary components. The vector field $Re(\omega)^\vee$ does not extend to the boundary of $\tilde{S}$ as its modulus gets arbitrarily large. However, its asymptotic direction is well-defined.
\begin{lemma} Let $\omega$ be an imaginary normalized differential on $S$ as in \eqref{omega}. Then, one has
\begin{enumerate}
\item[$\ast$] the 3 connected components of $\mathbb{R} S$ are parallel to $Re(\omega)^\vee$,
\item[$\ast$] $Re(\omega)^\vee$ is asymptotically orthogonal to $\gamma_{-1}$, $\gamma_1$ and $\gamma_\infty$. It is oriented inward $\tilde{S}$ at $\gamma_{-1}$ and $\gamma_1$ and outward $\tilde{S}$ at $\gamma_\infty$.
\item[$\ast$] For any point $p$ in $\gamma_{-1}$ or $\gamma_1$ out of $\mathbb{R} \tilde{S}$, the flow line of $Re(\omega)^\vee$ starting at $p$ ends in $\gamma_\infty$.
\item[$\ast$] $[-1,1]$ is the unique flow line in $\tilde{S}$ connecting $\gamma_{-1}$ to $\gamma_1$.
\end{enumerate}
\end{lemma}
\begin{figure}[h]
\input{PhasePortrait.tex}
\caption{Phase portrait of $Re(\omega)^\vee$ on $\tilde{S}$.}
\label{fig:phpt}
\end{figure}
\noindent
\begin{remark}
The latter picture is very similar to figure 4, exposé 6 in \cite{FLP}, concerning classification of measured foliation of pairs of pants and Thurston compactification of $\mathcal{M}_g$.
\end{remark}
\noindent
\begin{proof}
The first point is an easy consequence of the fact that $\omega$ is defined over $\mathbb{R}$. For the second point, consider a complex coordinate $z$ centred at one of the 3 punctures of $S$ and consider the circular integral
\[
\begin{array}{rcl}
\displaystyle \oint_{|z|=r} \omega & = & \displaystyle \oint_{|z|=r} \left( \frac{\lambda}{z} + h(z) \right) dz \\
& & \\
& = & \displaystyle \int_0 ^{2 \pi} \left( i \lambda + h(re^{i\theta}) ire^{i\theta} \right) d\theta \\
& & \\
& = & \displaystyle \int_0 ^{2 \pi} \left( i \lambda + r O(1) \right) d\theta
\end{array}
\]
where $h$ is an holomorphic function near the origin. The integrand converges uniformly to a purely imaginary function as $r$ goes to $0$. Equivalently, $Re(\omega)^\vee$ becomes everywhere orthogonal to the tangent vector field of the circle of radius $r$ as $r$ goes to 0. As $Im(\omega)^\vee$ is obtained by rotating $Re(\omega)^\vee$ by $\pi/2$, it gets asymptotically parallel to the tangent vector field of the circle of radius $r$ as $r$ goes to 0. These vectror fields points in the same direction if $\lambda>0$, and in opposite direction otherwise.\\
For the third point, notice that $Re(\omega)^\vee$ has a unique zero at the point $ \zeta := \frac{\lambda_1 - \lambda_{-1}}{\lambda_1 + \lambda_{-1}} \in ]-1,1[$. As this vector field is the gradient field of the harmonic function $\int Re(\omega)$, by classical Morse lemma, it is locally isotopic to the gradient field of $x^2-y^2$. There are exactly 2 flow lines passing through $\zeta$. One is $]-1,1[$ and the flow is oriented towards $\eta$ on each connected component of $ ]-1,1[ \; \setminus \; \zeta$. The other flow line intersects$ ]-1,1[ $ transversally at $\zeta$, and the flow is locally oriented outward $\eta$. Hence, any flow line starting from $p$ never reaches the only singular point $\zeta$ of the vector field $Re(\omega)^\vee$ and can be extended until it reaches the boundary of $\tilde{S}$. According to the second point, it has to end up in $\gamma_\infty$. The last point falls as a corollary.
\end{proof}
\begin{lemma}
Let $\left\lbrace Y_t\right\rbrace_{t >>1}$ be a sequence of pair of pants converging to $S \in \mathcal{M}_{0,3}$ and let $\left\lbrace \omega_t\right\rbrace_{t >>1}$ be a sequence of imaginary normalized differential on $Y_t$ converging to an imaginary normalized differential on $S$. For any connected component $\rho_t \subset Y_t$ of $\mathbb{R} Y_t$, one has
\[ \lim_{t \rightarrow \infty} \int_{\rho_t} Im(\omega_t) = 0. \]
\end{lemma}
\noindent
\begin{proof}
Assume first that the limit is as in \eqref{omega}. For any $t$, there exists positive numbers $r_{t,-1}$, $r_{t,1}$ and $r_{t, \infty}$ such that $$Y_t \simeq \left\lbrace z \in \mathbb{C} \big| \vert z \vert \leq r_{t,\infty}, \; \vert z-1 \vert \geq r_{t,1}, \; \vert z+1 \vert \geq r_{t,-1} \right\rbrace,$$ with boundary components $\gamma_{t,-1}$, $\gamma_{t,1}$, and $\gamma_{t,\infty}$. According to the previous lemma, there exists $N>0$ such that for almost all $t>N$, one has
\begin{enumerate}
\item[$\ast$] $\int_{\gamma_{t,-1}} Im(\omega_t) >0$ and $\int_{\gamma_{t,1}} Im(\omega_t) >0$,
\item[$\ast$] $Re(\omega_t)^\vee$ is everywhere transversal to $\gamma_{t,-1}$, $\gamma_{t,1}$ and $\gamma_{t,\infty}$,
\item[$\ast$] $\omega_t$ has a single zero on $Y_t$.
\end{enumerate}
The same treatment as in the proof of the previous lemma implies the existence of $M>N$ such that for almost all $t>M$, one has
\begin{enumerate}
\item[$\ast$] existence of a unique curve in $Y_t$, parallel to $Re(\omega_t)$, joining $\gamma_{t,-1}$ and $\gamma_{t,1}$ and converging in Hausdorff distance to $]-1+r_{t,-1}, 1-r_{t,1}[ \subset Y_t$,
\item[$\ast$] existence of two curves in $Y_t$, parallel to $Re(\omega_t)$, going from $\gamma_{t,-1}$ ($resp.$ $\gamma_{t,1}$) to $\gamma_{t,\infty}$ and converging in Hausdorff distance to $]-r_{t,\infty}, -1-r_{t,-1}[ $ ($resp.$ $]1 +r_{t,1}, r_{t,\infty}[$).
\end{enumerate}
If for example $\rho_t := ]-1+r_{t,-1}, 1-r_{t,1}[$ and $\breve{\rho}_t$ is the sequence of integral curves converging to $\rho_t$, then $\rho_t$ is homotopic to $ \rho_{t,1} \circ \breve{\rho}_t \circ \rho_{t,-1} $ where $ \rho_{t,-1} $ is an arc in $\gamma_{t,-1}$ and $ \rho_{t,1} $ is an arc in $\gamma_{t,1}$. As $\breve{\rho}_t$ converges to $\rho_t$, the integrals of $Im(\omega_t)$ over $ \rho_{t,-1} $ and $ \rho_{t,1} $ tend to 0 as $t$ goes to $\infty$. Hence
\[\lim_{t \rightarrow \infty} \int_{\rho_t} Im(\omega_t) = \lim_{t \rightarrow \infty} \int_{\breve{\rho}_t} Im(\omega_t) = 0. \]
The same argument apply for any sequence $\rho_t$ of components of $\mathbb{R} Y_t$ and the lemma is proved in this case. \\
If one of the residues of $\omega_t$ is zero for infinitely many times $t$, it converges to zero. The phase-portrait of the limiting 1-form is a radial foliation on a cylinder. Adapting the above lemma, one shows that there is always an integral curve converging to any connected component of $\mathbb{R} S$. The details are left to the reader. When two residues are zero for infinitely many times, then $\omega_t$ is zero and there is nothing to prove.
\end{proof}
$\,$\\
\noindent
\begin{proof}{\textbf{of theorem \ref{convperiod}.}}
For any loop $\rho_ \mathbb{T} \subset C$, the associated loop $\breve{\rho}_t \subset S_t$ is piecewise geodesic, see \ref{assloop}. According to the definition, $\breve{\rho}_t$ is made out of connected components of $\mathbb{R} Y_t$ for some pair of pants $Y_t$ of the decomposition of $S_t$ and arcs in some of the geodesics $\gamma_t$ of the same decomposition. By the previous lemma, parts of $\breve{\rho}_t$ contained in the different $\mathbb{R} Y_t$'s do not contribute in the limit of the integral of $Im(\omega_t)$ along $\breve{\rho}_t$. Let $\vec{e}$ be an edge of $\rho_ \mathbb{T}$, and $ \gamma_t$ the corresponding geodesic in $S_t$. By definition, $\breve{\rho}_t \cap \gamma_t$ is an arc of length $\frac{1}{2 \pi} \cdot \log \big( \Theta_t (e) \big) \cdot l_t(e)$ of $\gamma_t$. If one shows that
\[ \lim_{t \rightarrow \infty} \int_{\breve{\rho}_t \cap \gamma_t} Im(\omega_t) = \log \big( \Theta(e) \big) \cdot \big( \omega_{R,C} \big)_{\vec{e}} \;, \]
then the proposition is proved.\\
For $t$ big enough, and $\varepsilon >0$ small enough, consider the tubular neighbourhood $ \mathcal{U}_\varepsilon $ of width $\varepsilon$ around $\gamma_t$. Let $c_{t,1}$ and $c_{t,2}$ be the flow lines of $Re(\omega_t)$ starting at the end points of $ \breve{\rho}_t \cap \gamma_t$ and ending on $\partial \mathcal{U}_\varepsilon $. Denote by $\breve{\rho}_{t, \varepsilon} $ the arc on $\partial \mathcal{U}_\varepsilon $ between the end points of $c_{t,1}$ and $c_{t,2}$ such that $ \breve{\rho}_t \cap \gamma_t$ is homotopic to $ c_{t,2}^{-1} \circ \breve{\rho}_{t, \varepsilon} \circ c_{t,1}$. For any $t$, one has
\[ \int_{\breve{\rho}_t \cap \gamma_t} Im(\omega_t) = \int_{c_{t,2}^{-1} \circ \breve{\rho}_{t, \varepsilon} \circ c_{t,1}} Im(\omega_t) = \int_{\breve{\rho}_{t, \varepsilon}} Im(\omega_t) \]
as $c_{t,1}$ and $c_{t,2}$ are flow lines of $Re(\omega_t)$. Now, one clearly has
\[
\begin{array}{rcl}
\displaystyle \lim_{\varepsilon \rightarrow 0} \lim_{t \rightarrow \infty} \int_{\breve{\rho}_{t, \varepsilon}} Im(\omega_{R,C} )&=& \displaystyle \lim_{\varepsilon \rightarrow 0} \displaystyle \int_{\alpha_{\varepsilon, \Theta(e)}} \frac{\big( \omega_{R,C} \big)_{\vec{e}}}{z}dz \\
& & \\
&=& \log \big( \Theta(e) \big) \cdot \big( \omega_{R,C} \big)_{\vec{e}}
\end{array}\]
where $ \alpha_{\varepsilon, \Theta(e)} : = \left\lbrace z \in \mathbb{C} \; \big| |z|=\varepsilon, \; \theta \leq Arg(z) \leq \theta + \log \big(\Theta(e)\big) \right\rbrace$ for some argument $\theta$.
\end{proof}
\subsection{Harmonic tropical curves}\label{secharmtrop}
For any planar curve $ \mathcal{C} \subset ( \mathbb{C}^{\ast} )^{2}$, there exists a natural tropical curve sitting inside its amoeba $ \mathcal{A}( \mathcal{C})$: the so-called spine of $ \mathcal{C}$. It is defined via the Ronkin function associated to any defining equation of $ \mathcal{C}$. Planar amoebas deformation-retracts on their spine, see \cite{PR}.\\
The same construction can be carried to the case of harmonic amoebas, thanks to the extended notion of Ronkin function provided in \cite{Kri}. Hence, it gives rise to a piecewise linear graph sitting inside any harmonic amoeba, with the same deformation retraction property. They are similar to tropical curves in any point except that they might have edges with non rational slope. They are part of a wider class of tropical curves we introduce now.
\begin{definition}\label{tropmorph}
Let $C $ be a simple tropical curve. An \textbf{harmonic tropical morphism} $ \pi \, : \, C \rightarrow \mathbb{R}^m $ is a continuous map subject to the following
\begin{description}
\item[\textbf{Linearity:}] for any $ e \in LE( C )$, the map $ \pi_{\vert_e} $ is affine linear, with respect to the metric on $e$.
\item[\textbf{Balancing:}] For any $ v \in V ( C )$, denote by $\vec{e}_1$, $\vec{e}_2$ and $\vec{e}_3 $ the 3 outgoing unitary tangent vectors to the 3 leaves-edges adjacent to $v$. Then
\[ \pi (\vec{e}_{1}) +\pi (\vec{e}_{2}) +\pi (\vec{e}_{3}) =0. \]
\end{description}
An \textbf{harmonic tropical curve} $\Gamma \in \mathbb{R}^m$ is the image of a harmonic morphism $ \pi \, : \, C \rightarrow \mathbb{R}^m $.
\end{definition}
\begin{remark} Harmonic tropical curves in affine spaces should be equipped with an extra structure in order to satisfy a balancing condition. There is no natural definition for weights here, as leaves-edges have no longer rational slopes. They should be replaced by positive real numbers, standing for the (sum of) gradient of possible parametrizations. In the case of a tropical curve, it would correspond to the classical weights.\\
As point sets, harmonic tropical curves forms a much bigger class of objects : there are any piecewise linear graphs with no restrictions on slopes, as long as they locally span, around each vertex, the smallest linear space that contains it.
\end{remark}
\noindent
Similarly to the amoeba case, every harmonic tropical morphism can be described by integration of 1-forms, as stated below.
\begin{proposition}\label{propharmmorph}
Let $C$ be a simple tropical curve and $R$ a collection of residues of dimension $m$ on $C$. Then, the map $ \pi_R : C \rightarrow \mathbb{R}^m$ is a harmonic tropical morphism. For any vertex $v \in V(C)$ and any outgoing adjacent leaf-edge $\vec{e}$, the corresponding unitary tangent vector is mapped to $\omega_{R,C} \big(\vec{e}\big)$. \\
Reciprocally, for any harmonic tropical morphism $ \pi \; : \; C \rightarrow \mathbb{R}^m$, there exists a unique collection of residues $R$ of dimension $m$ on $C$ such that $ \pi = \pi_R$.
\end{proposition}
\noindent
\begin{proof}
As every coordinates of $\omega_{R,C}$ is a constant 1-form on any leaf-edge of $C$, $\pi_R$ is affine linear on every leaf-edge. By definition, $\omega_{R,C} \big(\vec{e}\big)$ is the gradient of $\pi_R$ along the oriented edge $\vec{e}$. The balancing condition of \ref{tropmorph} follows from the definition \ref{def1form}. The first part of the statement is proven.\\
Reciprocally, it is now clear that the map $\pi$ is given by integration of an $m$-tuple of constant 1-forms on any $e \in LE(C)$. The balancing condition of \ref{tropmorph} is clearly equivalent to the balancing condition of \ref{def1form}. Exactness follows from the fact that this $m$-tuple is given by differentiating the $m$-tuple of coordinate functions of $\pi$. Uniqueness follows from \ref{tdiff}.
\end{proof}
\begin{remark}
As for harmonic amoeba maps, the space of deformation of harmonic morphisms on a fixed tropical curve $C$ in $\mathbb{R}^m$ corresponds to the space of collection of residues of dimension $m$ on $C$. It is then a real vector space of dimension $m(n-1)$, where $n $ is the number of leaves of $C$.
\end{remark}
\begin{definition}\label{defint1form}
Let $C$ be a simple tropical curve. An element $\omega \in \Omega_0 ( C ) $ is \textbf{integer} if $\omega \big(\vec{e} \big) \in \mathbb{Z}$ for any $e \in LE(C)$. A collection of residues $R$ of dimension $m$ is \textbf{integer} on $C$ if the $m$ coordinates of $\omega_{R,C}$ are integer elements of $\Omega_0 ( C ) $.
\end{definition}
\subsection{Degeneration of harmonic amoebas}
In order to prove theorem \ref{approxtrop}, one need to make precise the notion of convergence we use. we use the notations introduced in section \ref{sectropcurvmorph}.
\begin{definition}\label{GHconv}
In theorem \ref{approxtrop}, one says that the sequence of maps $ \big(1/\log (t)\big) \cdot \mathcal{A}_R : S_t \rightarrow \mathbb{R}^m $ converges to the map $\pi_R : C \rightarrow \mathbb{R}^m$ if for any $v \in V(C)$, $ \frac{1}{\log (t)} \mathcal{A}_R (Y_{v,t}) $ converges in Hausdorff distance to $\pi_R(T^{1/2}_v)$.
\end{definition}
Note that the latter definition is stronger than the Hausdorff distance of the image of $ \big(1/\log (t)\big) \cdot \mathcal{A}_R $ to the image of $\pi_R$.
\begin{remark}
If $\left\lbrace S_t \right\rbrace_{t>>1}$ converges to a simple tropical curve $C=(G,l)$ and $d^u_t$ denotes the uniform metric on $S_t$, the Gromov-Hausdorff limit of $\left\lbrace (S_t, d^u_t) \right\rbrace_{t>>1}$ is the stable curve dual to $G$, equipped with its uniformizing metric. Hence, it does not depend on the length function $l$. Simple computations show that the diameter of any collar around a vanishing cycle of $\left\lbrace S_t \right\rbrace_{t>>1}$ is asymptotically equivalent to $\log \big( \log (t) \big)$, and that up to the rescaling
\[ d_t := \frac{2 \pi}{\log (t)} \cdot\sinh \left( \dfrac{d^u_t}{2}\right),\]
the sequence of metric spaces $\left\lbrace ( S_t, d_t) \right\rbrace_{t>>1}$ converges in Gromov-Hausdorff distance to $C$. The collars ($resp.$ cusps) converge to their dual edge ($resp.$ leave) while the bounded parts of the pair of pants converge to their dual vertex. In the compact case ($n=0$), the above computation expresses the major difference with the approach of \cite{O}: the Gromov- Hausdorff limit of a sequence converging $C$ is a metric graph supported on $G$ where all edges have the same length.
\end{remark}
\noindent
\begin{proof}{\textbf{of theorem \ref{approxtrop}.}}
Let us prove the first part of the statement. Assume that $m \geq 2$, the case $m=1$ will easily follows.
Recall that one needs to specify initial points on the source spaces of the maps $\pi_R$ and $ \mathcal{A}_R$. Everything below will be defined up to translation until we fix these initial points. For any $v \in V(C)$, denote by $R_{v,t}$ the $m \times 3$ matrix of the periods of $\omega_{R,S_t}$ along the 3 boundary geodesics of $Y_{v,t}$. The matrix $R_{v,t}$ is a matrix of residues of dimension $m$ that converges, thanks to theorem \ref{convdif}, to a matrix $R_v$, columns of which correspond to the slope of $\pi_R$ on the leaves-edges of $T^{1/2}_v$. Denote by $S$ the point of $\mathcal{M}_{0,3}$. $Y_{v,t}$ is a subset of $S$ converging to $S$ when $t$ gets arbitrarily large, and the restriction of $\omega_{R,S_t}$ to $Y_{v,t}$ converges to $\omega_{R_v,S}$ on $S$. It implies that
$ \mathcal{A}_R (Y_{v,t}) $ converges to $\mathcal{A}_{R_{v}} (S)$, where the half collars and cusps of $Y_{v,t}$ are mapped in the limit to the corresponding tentacles of the harmonic amoeba $\mathcal{A}_{R_{v}} (S)$, whereas the bounded part $Y_{v,t}^{bd}$ is mapped to a compact region. Note that the latter amoeba might be degenerate to a line or a point if some lines or columns of $R_v$ are zero. Reintroducing the $\log(t)$-rescaling, one has
\[ \lim_{t \rightarrow \infty} \frac{1}{\log(t)} \mathcal{A}_R (Y_{v,t}) \subset \lim_{t \rightarrow \infty} \frac{1}{\log(t)} \mathcal{A}_{R_{v}} (S).\]
The right-hand side is the (possibly degenerate) infinite tripod supporting $\pi_R \big( T^{1/2}_v\big)$. Denote it $\Gamma_v$. Let us show that the left-hand side is $\pi_R \big( T^{1/2}_v\big)$, as desired. From what is above, one deduces that $Y_{v,t}^{bd}$ is contracted to the vertex, the cusps are asymptotically mapped to the corresponding leaves of $\pi_R \big( T^{1/2}_v\big)$, and the half collars are asymptotically mapped into the corresponding rays of $\Gamma_v$. The proposition \ref{convdifloc} clearly applies to half collars where the asymptotic has to be divided by 2. One deduces that half collars are mapped to the corresponding edges of $\pi_R \big( T^{1/2}_v\big)$. In order to conclude the proof of the first part of the theorem, one needs to fix initial points for $\pi_R$ and $ \mathcal{A}_R$. We fix a vertex $v \in C$ as an initial point for $\pi_R$ and choose any sequence of points $z_{0,t} \in Y_{v,t}^{bd}$. Then, the sequence $\left\lbrace z_{0,t} \right\rbrace_{t>>1}$ converges to $v$ and all of them are mapped to $0 \in \mathbb{R}^m$ by $ \mathcal{A}_R$ and $\pi_R$, respectively.\\
For the second part of the statement, let us first show that $\left\lbrace S_t \right\rbrace_{t>>1}$ converges to the stable curve dual to $C$ in $\overline{\mathcal{M}}_{g,n}$. For any arbitrarily small $ \varepsilon >0 $ there exists $t_\varepsilon$ such that $\big( 1/ \log (t) \big) \cdot \mathcal{A}_R (S_t)$ is contained in the $\varepsilon$-neighbourhood of $\pi_R(C)$ for any $t>t_\varepsilon$. For any edge $e$ in $\pi_R(C)$, there is a cylinder in $S_t$ ($t>t_\varepsilon$) mapping into the $\varepsilon$-neighbourhood of $e$. Indeed, applying the maximum principle to $\mathcal{A}_R$, there has to be a Riemann surface with boundaries, image of which joins the two ends of the $\varepsilon$-neighbourhood of $e$. It has negative Euler characteristic. Applying the same reasoning to any leaf or edge of $\pi_R(C)$, it has to be a cylinder by additivity of the Euler characteristic and injectivity of $\pi_R$. Now, the map $ \mathcal{A}_R$ is unbounded on each such sequence of cylinders, as only the image of $\big( 1/ \log (t) \big) \cdot \mathcal{A}_R $ converges. It means that each of them corresponds to a vanishing cycle of $\left\lbrace S_t \right\rbrace_{t>>1}$. Hence $\left\lbrace S_t \right\rbrace_{t>>1}$ converges to the stable curve dual to $C$. In order to achieve the proof, one needs to recover how fast the vanishing cycles shrink. This can be recovered from the length on the edges of $\pi_R(C)$ using lemmas \ref{lemmaconfcollar} and \ref{lemmaconfdiff}. It only remains to show that $\left\lbrace (S_t, \omega_{R,S_t}) \right\rbrace_{t >>1} \subset \Lambda_{g,n} ^m$ converges to $\big[ C, \omega_{R,C} \big]$, as in theorem \ref{convdif}. Indeed, the last argument of the proof of the latter theorem implies that the periods of $\omega_{R,S_t}$ are uniformly bounded in $t$. For any subsequence of $S_t$ such that $\omega_{R,S_t}$ converges, the limit is $\omega_{R,C}$: As observed in the first part of the proof, the slopes of the edges $ \pi_R( \vec{e}) $ are given by $ \omega_{R,S_t}(\vec{e})$. Hence $\omega_{R,S_t}$ converges to $\omega_{R,C}$ and the theorem is proven.
\end{proof}
\begin{remark}
Approximation of tropical curves by families of amoebas is treated $e.g.$ in \cite{Mikh03}, \cite{Mikh06}, \cite{NS}, or \cite{Nish}.
Going over the algebraic framework allows to extends considerably the various approximation theorems that can be found in the above references, as the theorem \ref{approxtrop} states that any harmonic tropical curve can be obtained as the Hausdorff limit of harmonic amoebas.
\end{remark}
\section{Approximation of phase-tropical curves}
In this last section, we go back to the algebraic case. Algebraicity impose severe restrictions on the realizability (or ``approximability") of tropical curves. Whereas any harmonic tropical curve in $\mathbb{R}^m$ arise as the Hausdorff limit of families of harmonic amoebas, realizability of tropical curves is much more delicate than the simple integrality of slopes. There are subtle obstructions (see $e.g.$ \cite{Mikh06} or \cite{Nish}). Here, we work under the most natural necessary condition, namely regularity.
\subsection{Phase-tropical morphisms}
Recall that a toric morphism $ A : \big( \mathbb{C}^\ast \big)^k \rightarrow \big( \mathbb{C}^\ast \big)^m $ is a map of the form
\[ (z_1, ..., z_k) \mapsto \big( b_1 z_1^{a_{11}}...z_k^{a_{1k}}, ..., b_m z_1^{a_{m1}}...z_k^{a_{mk}} \big) \]
where $(b_1,...,b_m) \in \big( \mathbb{C}^\ast \big)^m $ and $ \big( a_{ij} \big) \in M_{m \times k} \big( \mathbb{Z} \big)$.
\begin{definition}\label{defphtropmorph}
A \textbf{phase-tropical morphism} $\phi : V \rightarrow \big( \mathbb{C}^\ast \big)^m $ on a simple phase-tropical curve $ \mathcal{A}_V : V \rightarrow C$ is a map such that for any $v \in V(C)$, and the corresponding chart $\varphi_v : \mathcal{A}_V^{-1} ( T_v ) \rightarrow ( \mathbb{C}^{\ast} )^{2}$, there exists a toric morphism $A_v : ( \mathbb{C}^{\ast} )^{2} \rightarrow \big( \mathbb{C}^\ast \big)^m$ such that
\[ \phi_{\big| \mathcal{A}_V^{-1} ( T_v )} = A_v \circ \varphi_v \]
The morphism $\phi$ induces a tropical morphism $ \pi_\phi : C \rightarrow \mathbb{R}^m$ defined by
\[ \big( \pi_\phi\big)_{\big| \mathcal{A}_V^{-1} ( T_v )} = A_v \circ \pi_v. \]
\end{definition}
\begin{remark}
Phase-tropical curves originally appeared as immersed objects in $( \mathbb{C}^{\ast} )^{2}$, obtained by degeneration of families of algebraic curves, see section $6.2$ in \cite{Mikh05}. Their abstract counterpart and associated morphisms appeared later in \cite{Mikhunpub}.
\end{remark}
\begin{proposition}
Any phase-tropical morphism $\phi : V \rightarrow \big( \mathbb{C}^\ast \big)^m $ is uniquely determined by its underlying tropical morphism $ \pi_\phi : C \rightarrow \mathbb{R}^m$, up to a toric translation $(z_1, ..., z_m) \mapsto \big( b_1 z_1, ..., b_m z_m \big)$ on the target space.
\end{proposition}
\noindent
\begin{proof}
As the toric morphism $ (z_1, ..., z_k) \mapsto \big( b_1 z_1^{a_{11}}...z_k^{a_{1k}}, ..., b_m z_1^{a_{m1}}...z_k^{a_{mk}} \big) $ descends to the affine-linear map $$ (x_1, ..., x_k) \mapsto \big( \log\vert b_1\vert + a_{11} x_1+ ...+ a_{1k} x_k, ..., \log \vert b_m \vert + a_{m1} x_1 + ... + a_{mk} x_k \big)$$ by composition with $ \mathcal{A}$, the tropical morphism $\pi_\phi$ determines every $A_v$ up to the arguments of its translation part $b_v := (b_{v,1},...,b_{v,m}) \in \big( \mathbb{C}^\ast \big)^m $. Up to translation, one can assume that the latter vector is $(1,...,1)$ for a chosen vertex $v_0 \in V(C)$. One easily checks that the compatibility conditions imposed by the transition functions on $V$ determine all the other $b_v$'s.
\end{proof}
Now let us recall what are regularity and superabundancy of tropical-morphisms, see \cite{Mikh05} for more details.
Recall that any simple tropical curve can be presented as a couple $(G,l)$ of a cubic graph and a length function.
\begin{definition}
Let $C$ and $C'$ be two simple tropical curves supported on the same cubic graph $G$. One says that two tropical morphisms $\pi : C \rightarrow \mathbb{R}^m$ and $\pi' : C' \rightarrow \mathbb{R}^m$ have the same \textbf{combinatorial type} if for any oriented leaf-edge $\vec{e} \in LE(G)$,
$\pi(\vec{e})$ and $\pi'(\vec{e})$ have the same slope in $S^{m-1} \cup \left\lbrace 0 \right\rbrace$.\\
\end{definition}
The space of deformation of tropical morphisms within a fixed combinatorial type is subject to a finite number of constraints. Therefore, its dimension can be bounded by below by the ``expected dimension". Here, we are interested in the space of tropical curves supporting a tropical morphism of a fixed combinatorial type rather than the space of deformation itself.
\begin{proposition}
Let $G$ be a cubic graph of genus $g$ with $n$ leaves and $\pi : (G,l) \rightarrow \mathbb{R}^m$ be a tropical morphism. The space of tropical curves supported on $G$ and admitting a tropical morphism combinatorially equivalent to $\pi$ is the relative interior of an open convex polyhedral domain of codimension at most $mg$ in $ \big( \mathbb{R}_{>0} \big)^{3g-3+n} $.
\end{proposition}
\noindent
See section 2.4 in \cite{Mikh05} for a proof.
\begin{definition}
Let $G$ be a cubic graph of genus $g$. A tropical morphism $\pi : (G,l) \rightarrow \mathbb{R}^m$ is \textbf{regular} if the space of simple tropical curves supported on $G$ and admitting a tropical morphism combinatorially equivalent to $\pi$ has codimension $mg$ in $ \big( \mathbb{R}_{>0} \big)^{3g-3+n} $. Otherwise, $\pi$ is \textbf{superabundant}. A phase-tropical morphism $\phi$ is regular if its underlying tropical morphism $\pi_\phi$ is.
\end{definition}
\subsection{Mikhalkin's theorem}\label{Mik}
In this last section, we state and prove Mikhalkin's theorem in a framework close to the one of \cite{Mikhunpub}. The main difference is the way regularity steps in the proof. In \cite{Mikhunpub}, regularity allows to find principal divisor on appropriate sequences of Riemann surfaces, giving the desired sequence of maps up to the Abel-Jacobi theorem.\\
The approximation of phase-tropical morphisms requires more than the approximation of the underlying tropical morphism, guaranteed by the theorem \ref{approxtrop}. In the present framework, one needs to construct sequences of harmonic amoeba maps that have well-defined harmonic conjugate and lift to actual holomorphic maps. Technically, one has to look for harmonic maps coming from i.n.d.'s with $2 i \pi$ integer periods. That's exactly where regularity steps in, see proposition \ref{propreg}.\\
Recall the change of complex structure
\[
\begin{array}{rcl}
H_t \: : \: \big( \mathbb{C}^\ast \big)^m & \rightarrow & \big( \mathbb{C}^\ast \big)^m \\
(z_1,...,z_m) & \mapsto & \left( \vert z_1 \vert^{\frac{1}{\log(t)}} \frac{z_1}{\vert z_1 \vert} ,..., \vert z_m \vert^{\frac{1}{\log(t)}} \frac{z_m}{\vert z_m \vert} \right)
\end{array}.
\]
Then, one has
\begin{mytheorem}[Mikhalkin]\label{thmMik}
Let $ \phi : V_{(C,\Theta)} \rightarrow \big( \mathbb{C}^\ast \big)^m$ be a regular phase-tropical morphism on a simple phase-tropical curve $V_{(C,\Theta)}$, where $C$ has genus $g$ and $n$ leaves. Then, there exists a sequence $ \left\lbrace S_t \right\rbrace_{t>>1} \subset \mathcal{M}_{g,n} $ together with algebraic maps $ \iota_t \, : \, S_t \rightarrow (\mathbb{C}^\ast)^m $ such that $ H_t \big( \iota_t ( S_t ) \big) $ converges in Hausdorff distance to $\phi \big( V_{(C,\Theta)} \big)$. Moreover, one can require that $S_t=S_{(C_t, \theta_t)}$ with $ \Theta_t = \Theta$, for all $t$.
\end{mytheorem}
\begin{remark} In certain cases, it is of primary importance that the latter theorem holds for sequences of Riemann surfaces with constant twist distributions. The construction of real algebraic curves with prescribed topology related to Hilbert's sixteenth problem is a fundamental instance.
\end{remark}
\begin{remark} As every phase-tropical immersion to $( \mathbb{C}^{\ast} )^{2}$ is regular, theorem \ref{thmMik} takes care of this case.
\end{remark}
\noindent
When integrality is guaranteed, one has the following.
\begin{proposition}\label{approxcomptrop}
Let $ \phi : V \rightarrow \big( \mathbb{C}^\ast \big)^m$ be a phase-tropical morphism on a simple phase-tropical curve $ V:=V_{(C,\Theta)}$ and $R$ be the collection of residues giving the underlying tropical morphism $\pi_\phi : C \rightarrow \mathbb{R}^m$. For any sequence $ \left\lbrace S_t \right\rbrace_{t \in \mathbb{N}, t>>1} $ of Riemann surface converging to $V$ such that $\mathcal{P}_{R,S_t}$ is a constant family of integer period matrix, then $ H_t \big( \iota_{R} ( S_t ) \big) $ converges in Hausdorff distance to $\phi \big( V \big)$.
\end{proposition}
\noindent
\begin{proof}
The proof is similar to the one of theorem \ref{approxtrop}. We first show the convergence locally for any tripod $T^{1/2} _v$ and check that the pieces glue together in the expected way. As before, we will work up to translation until we fix the initial points in $S_t$. For any $v \in V(C)$, let us show that $ H_t \big( \iota_{R} ( Y_{v,t} ) \big) $ converges to $\phi \big( \mathcal{A}_V^{-1} (T^{1/2} _v) \big) $. Using the notation of the proof of theorem \ref{approxtrop}, one has that $\iota_{R} ( Y_{v,t} )$ converges to $\iota_{R_v} ( S ) $. Recall that $\iota_{R_v}$ is a toric embedding if the 3 columns of $R_v$ are non zero. Otherwise the image is a cylinder or a point if some lines or columns of $R_v$ are zero. Reintroducing the rescaling $H_t$, one deduces that, in the non degenerate case, $ H_t \big( \iota_{R} ( Y_{v,t} ) \big) $ converges to the part of the embedded phase-tropical line $\lim_{t\rightarrow \infty} H_t \big( \iota_{R_v} ( S ) \big)$ supported on $\pi_\phi (T^{1/2} _v) $, namely $\phi \big( \mathcal{A}_V^{-1} (T^{1/2} _v) \big) $. The 2 degenerate cases are similar.
Up to now, we have shown that each piece $H_t(\iota_{R} (Y_{v,t}))$ converge to $\phi \big( \mathcal{A}_V^{-1} (T^{1/2} _v) \big) $, up to toric translation. It is clear that one can choose the sequence of initial points $z_{0,t} \in S_t$ such that $H_t(\iota_{R} (Y_{v_0,t}))$ converge to $\phi \big( \mathcal{A}_V^{-1} (T^{1/2} _{v_0}) \big) $, for a chosen vertex $v_0 \in V(C)$. Now, one only needs to check that the pieces $H_t(\iota_{R}(Y_{v,t}))$ glue together the same way as the $\phi \big( \mathcal{A}_V^{-1} (T^{1/2} _v) \big) $ do, when $t$ gets arbitrarily large. For 2 vertices connected by an edge $e$ which is contracted by $\pi_\phi$, there is nothing to check as the cylinder $ \mathcal{A}_V^{-1}(e)$ is contracted to a point. If $\pi_\phi$ is injective on $e$, the gluing is fixed by the compatibility condition of \ref{defphtropmorph} and hence determined by the twist parameter $\Theta(e)$. It is given by the automorphism $z \mapsto -\overline{\Theta (e) z}$ on the geodesic $Arg \big( \phi \big( \mathcal{A}_V^{-1}(e) \big) \big)$ in the argument torus. On the other side, the geodesic $\gamma_{e,t} \subset S_t $ dual to $e$ is asymptotically mapped to the same geodesic $Arg \big( \phi \big( \mathcal{A}_V^{-1}(e) \big) \big)$ in the argument torus and the gluing is asymptotically given by $\Theta_t(e)$. As the sequence $\left\lbrace \Theta_t \right\rbrace_t$ converges to $\Theta$, the result follows.
\end{proof}
The two following results tell that the convergence towards a phase-tropical curve ensures integrality at the limit.
\begin{lemma}\label{intperiod}
Let $ \phi : V \rightarrow \big( \mathbb{C}^\ast \big)^m$ be a phase-tropical morphism on a simple phase-tropical curve $ V:=V_{(C,\Theta)}$ and $R$ be the collection of residues giving the underlying tropical morphism $\pi_\phi : C \rightarrow \mathbb{R}^m$. For any loop $\rho_ \mathbb{T} \subset C$, one has
\[ \sum_{\vec{e} \in \rho_{ \mathbb{T}}} \log \big( \Theta (e) \big) \cdot \omega_{R,C} (\vec{e}) \in \big( 2 i \pi \mathbb{Z} \big)^m. \]
\end{lemma}
\noindent
\begin{proof}
For any vertex $v \in \rho_ \mathbb{T}$, there is a distinguished point among the 3 special points of $ \left( \mathcal{A}_V \right)^{-1} (v) $, namely the intersection point of the 2 geodesics corresponding to the 2 edges in $\rho_ \mathbb{T}$ adjacent to $v$. Let us look at the position of the image by $ Arg \circ \phi $ of this distinguished point in $ \big( \mathbb{R}/ 2 \pi \mathbb{Z} \big)^m $ while going around $\rho_ \mathbb{T}$.
Going from a vertex $v$ to the next one via an edge $\vec{e}$, the distinguished point is translated in the argument torus by $\frac{1}{i} \Theta(e) \cdot \omega_{R,C} (\vec{e})$. After a full cycle along $\rho_ \mathbb{T}$, the distinguished point has to end up at its initial place. Summing these displacements in the universal cover $\mathbb{R}^m$ of the argument torus, it is equivalent to say that
\[ \sum_{\vec{e} \in \rho_{ \mathbb{T}}} \frac{1}{i} \log \big( \Theta (e) \big) \cdot \omega_{R,C} (\vec{e}) \in 2 \pi \mathbb{Z}^m, \]
which is equivalent to the statement.
\end{proof}
\begin{proposition}
Let $ \phi : V \rightarrow \big( \mathbb{C}^\ast \big)^m$ be a phase-tropical morphism on a simple phase-tropical curve $ V:=V_{(C,\Theta)}$ and $R$ be the collection of residues giving the underlying tropical morphism $\pi_\phi : C \rightarrow \mathbb{R}^m$. For any family $ \left\lbrace S_t \right\rbrace_{t >>1} \subset \mathcal{M}_{g,n} $
converging to $V$, the family of period matrices $ \left\lbrace \mathcal{P}_{R,S_t} \right\rbrace_{t \in \mathbb{N}}
$ of $S_t$ with respect to $R$ converges to an integer period matrix $ \mathcal{P}_{R, C}$.
\end{proposition}
$\,$\\
\begin{proof}
According to theorem \ref{convdif}, the period vectors $ \frac{1}{2 i \pi} \big( \int_{\gamma_{e,t}} \omega_{R,S_t}^{(1)},..., \int_{\gamma_{e,t}}
\omega_{R,S_t}^{(m)} \big)$ tends to the vector $ \omega_{R,C} (\vec{e}) $, where $\gamma_{e,t}$ is
the geodesic of the decomposition of $S_t$ corresponding to $e \in E(C)$. As $\pi_\phi$ is a tropical morphism,
the collection of residues $R$ is integer on $C$. Hence, the latter period vectors tend to be integers.
Now let us consider a basis $\rho_ \mathbb{T} ^{(1)},..., \rho_ \mathbb{T} ^{(g)} \in H_1(C, \mathbb{Z})$ and consider the associated family of piecewise geodesic loops $\breve{\rho}_t ^{(1)},..., \breve{\rho}_t ^{(g)} \subset S_t$ (see definition \ref{assloop}). By theorem \ref{convperiod} and lemma \ref{intperiod}, one has that
\[\lim_{t\rightarrow \infty} \int_{\breve{\rho}_t ^{(k)}} \omega^{(j)}_{R,S_t} \in 2 i \pi \mathbb{Z}\]
for $1 \leq k \leq g$ and $1 \leq j \leq m$. As the $\gamma_{t,e}$ and the $\breve{\rho}_t ^{(k)}$ generate $H_1(S_t,\mathbb{Z})$ for any $t$, the proposition is proved.
\end{proof}
For the rest of this section, $ \phi : V \rightarrow \big( \mathbb{C}^\ast \big)^m$ is a regular phase-tropical morphism on a phase-tropical curve $ V:=V_{(C,\Theta)}$ and $R$ be the collection of residues giving the underlying tropical morphism $\pi_\phi : C \rightarrow \mathbb{R}^m$. For any family $ \left\lbrace S_t \right\rbrace_{t >>1} \subset \mathcal{M}_{g,n} $. The genus and number of leaves of $C$ will be denoted $g$ and $n$ respectively.\\
In order to prove theorem \ref{thmMik}, one needs to prove integrality not only at the limit but also after a finite time $t$.\\
Now let $G$ be the cubic graph supporting $C$. If $2g-2+n>0$, recall one can define the map
\[
\begin{array}{rcl}
\mathcal{FN}_{G} \: : \: \big( \mathbb{C}^\ast \big)^{3g-3+n} & \rightarrow & \mathcal{M}_{g,n} \\
(l,\Theta) & \mapsto & S_{(C, \Theta)}
\end{array},
\]
Now consider the following partial compactification of $ ( \mathbb{C}^\ast )^{3g-3+n} $ at the origin : consider the length factor $ ( \mathbb{R}_{>0} )^{3g-3+n} $ of its polar coordinate system, and embed it in the real oriented blowup of $ \mathbb{R}^{3g-3+n} $ at the origin. The latter ambiant space is diffeomorphic to $ \left\lbrace \underline{x} \in \mathbb{R}^{3g-3+n} \: \big| \: \vert \underline{x} \vert \geq 1 \right\rbrace $ and $ ( \mathbb{R}_{>0} )^{3g-3+n} $ is presented there as $$ \left\lbrace \underline{x} \in ( \mathbb{R}_{>0} )^{3g-3+n} \: \big| \: \vert \underline{x} \vert > 1 \right\rbrace. $$
Consider its partial compactification $$ \left\lbrace \underline{x} \in ( \mathbb{R}_{>0} )^{3g-3+n} \: \big| \: \vert \underline{x} \vert \geq 1 \right\rbrace $$ and denote by $ ( \mathbb{C}^\ast )_0 ^{3g-3+n} $ the product space of $( S^1)^{3g-3+n}$ with the latter partial compactification, and define also the following subset
$$F_0 := ( S^1)^{3g-3+n} \times \left\lbrace \underline{x} \in ( \mathbb{R}_{>0} )^{3g-3+n} \: \big| \: \vert \underline{x} \vert = 1 \right\rbrace. $$
The set $F_0$ is considered as the set of all equivalence classes of all the phase-tropical curves supported on $G$ with respect to the following relation
\begin{center}
$V_{( C_1, \mathscr{O}_1)} \sim V_{( C_2, \mathscr{O}_2)}$ iff $ \mathscr{O}_1 = \mathscr{O}_2 $ \linebreak and $\exists \lambda>0 $ such that $ l_1 = l_2 $,
\end{center}
$i.e.$ the underlying tropical curves are dilatation one of each others. The map $\mathcal{FN}_{G} $ extends to $ ( \mathbb{C}^\ast )_0 ^{3g-3+n} \rightarrow \overline{\mathcal{M}}_{g,n}$, and $F_0$ is mapped to $S_{G}$, the stable curve of dual graph $G$. Now, consider the map
\[
\begin{array}{rcl}
\Pi_R \: : \: ( \mathbb{C}^\ast ) ^{3g-3+n} & \rightarrow & M_{(2g+n-1) \times m} (\mathbb{R}) / Sl_{(2g+n-1)} (\mathbb{Z}) \\
(l,\mathscr{O}) & \mapsto & \mathcal{P}_{R,S_{(C,\mathscr{O})}}
\end{array}
\]
which associates to $(l,\mathscr{O})$ the period matrix of $S_{(C,\mathscr{O})}$ with respect to $R$, see definition \ref{defpermat}.
\begin{proposition}\label{propreg}
$\Pi_R$ extends analytically to $ ( \mathbb{C}^\ast )_0 ^{3g-3+n}$. Let $[V]$ be the equivalence class of $V$ in $ F_0$, then each irreducible component of the level set $ \Pi_R^{-1} \big( \Pi_R( [V]) \big) \subset ( \mathbb{C}^\ast )_0 ^{3g-3+n}$ is a smooth analytic subset of real codimension $2mg$ near $F_0$.
\end{proposition}
\noindent
\begin{proof}
From theorem \ref{convperiod}, one remark that the limiting period matrix of a family $S_t$ converging to a phase-tropical curve $\mathscr{V} = V_{(\mathscr{C},\mathscr{O})}$ depends only on the exact 1-forms $\omega_{R,\mathscr{C}} $ on $\mathscr{C}$ and $\mathscr{O}$. As the residues of $\omega_{R,\mathscr{C}}$ are invariant under dilatation of $\mathscr{C}$, it follows that $\Pi_R$ extends to $ ( \mathbb{C}^\ast )_0 ^{3g-3+n}$. The limiting periods are expressed analytically in terms of $(l,\mathscr{O})$ as the terms $ \omega_{R,\mathscr{C}} (\vec{e})$
depend linearly on the length function of $\mathscr{C}$. It follows that the extension of $\Pi_R$ is analytic.\\
For any simply connected domain $\mathcal{U} \subset ( \mathbb{C}^\ast ) ^{3g-3+n} $, one can trivialize \linebreak $H_1(S,
\mathbb{Z})\simeq H$ for any $S \in \mathcal{FN}_{G}(\mathcal{U})$. Now fix a basis $\rho_1,...,\rho_{{2g+n-1}}
$ of $H$ such that $\rho_1,..., \rho_{n-1}$ are given by small loops around $n-1$ of $n$ punctures, and consider the
period matrix $ \mathcal{P}_{R,S} \in M_{(2g+n-1) \times m} (\mathbb{R}) $ of any curve $S\in \mathcal{U}$ with respect to
the latter basis of $H$. This gives a map to $\mathbb{R}^{2(2g+n-1)}$, but the $m(n-1)$ coordinates corresponding to
$\rho_1,..., \rho_{n-1}$ are constant as they just compute the residues at the punctures. According to theorem \ref{thmcodim}, the remaining coordinate functions define $m$ smooth analytic subvarieties of codimension $2g$. In order to show that each connected component of their intersection is smooth and of codimension $2mg$ near $F_0$, one only needs to show that its intersection with $F_0$ is smooth of codimension $2mg$. We claim that this intersection is described on the set of classes $[\mathscr{V}]$ with $\mathscr{V}= V_{(\mathscr{C},\mathscr{O})}$ and $\mathscr{C} = (G,l) $ by
\begin{equation}\label{condperiod1}
\sum_{\vec{e} \in \rho} l (e) \cdot \omega_{R,C} (\vec{e}) = 0 \in \mathbb{R}^m
\end{equation}
\begin{equation}\label{condperiod2}
\sum_{\vec{e} \in \rho} \log \big( \mathscr{O} (e) \big)\cdot \omega_{R,C} (\vec{e}) = 0 \in (S^1)^m
\end{equation}
for any loop $\rho \subset \mathscr{C}$. As $\mathscr{C}$ and $C$ are both supported on $G$, their edges are in natural bijection. By proposition \ref{tdiff}, the bunch of equation \eqref{condperiod1} is equivalent to saying that the $m$ exact 1-forms $\omega_{R,C}$ and $\omega_{R,\mathscr{C}}$ are equal in the following sense : for any $\vec{e} \in E(G)$, $\omega_{R,C} (\vec{e})= \omega_{R,\mathscr{C}} (\vec{e})$. These are necessary conditions, as these values are indeed limits of corresponding periods by theorem \ref{convdif}. For the bunch of equations \eqref{condperiod2}, the left-hand side can be seen as a limit period vector in $\mathbb{R}^m$ for a well chosen basis $\rho_1,...,\rho_{{2g+n-1}}
$ of $H$, see theorem \ref{convperiod}. The equations \eqref{condperiod2} are equivalent to the fact that the associated periods are in the lattice $2i\pi \mathbb{Z}^m$, which is necessary by lemma \ref{intperiod}.
In order to see that those conditions are also sufficient, one can extract out of them a subset of $2mg$ linearly independent conditions, by picking $g$ independant loops in $\mathscr{C}$. Clearly the conditions \eqref{condperiod1} are independent of the conditions \eqref{condperiod2}, and both have the same rank. The fact that the conditions of \eqref{condperiod1} are $mg$ dimensional follows from the regularity assumption.
\end{proof}
$\,$\\
\noindent
\begin{proof}{\textbf{of theorem \ref{thmMik}.}}
There are only two types of graphs for which the condition $2g-2+n>0$ is not satisfied : a segment and a loop. The second one cannot be described by a cubic graph. Nevertheless, the theorem obviously holds in both cases.\\
Let's assume that $2g-2+n>0$. Thanks to the previous proposition, one can pick a sequence of points $ \left\lbrace \underline{z}_t \right\rbrace_{ t>>1} \subset \Pi_R^{-1} \big( \Pi_R ( [ V ] ) \big) $ converging to the point $\left[ V \right] \in F_0$ and such that $z_t:=(l_t, \Theta_t)$ with $l_t \sim 4\pi /\big( \log(t) \cdot l\big)$ where $l$ is the length function on $V$. According to definition \ref{comptropconv}, the sequence $S_t : = \mathcal{FN}_{G} (\underline{z}_t) $ converges to $V$. The sequence $ \left\lbrace S_t \right\rbrace_{ t>>1} $ has been constructed such that the period matrix $ \mathcal{P}_{R,S_t}$ is constant and integer, thanks to lemma \ref{intperiod}. Applying proposition \ref{approxcomptrop}, one deduces that $H_t \circ \iota_R (S_t) $ converges to $\phi$.\\
To see that one can pick the sequence $ \left\lbrace \Theta_t \right\rbrace_{t>>1}$ to be constant, remark that the same arguments as in the previous proposition imply that $ \Pi_R^{-1} \big( \Pi_R ( [ V ] ) \big) \cap \left\lbrace \mathscr{O} = \Theta \right\rbrace$ is analytic, smooth and of codimension $mg$ in the length factor of $( \mathbb{C}^\ast )_0 ^{3g-3+n}$.
\end{proof}
\bibliographystyle{alpha}
|
\section{Introduction}
\noindent
Ultimately, we would like to understand massive stars and their progeny
both locally as well as in the distant Universe. What is clear is
that rotation, mass loss, and the link between them, play a pivotal role
in the fate of massive stars.
However, in order to test mass-loss predictions for rotating stars, we
need to probe the density contrast between the stellar pole and
equator. In the local Universe, this may potentially be achievable through
the technique of long-baseline interferometry, as discussed during this
meeting. However, in order to determine wind asymmetry in the more
distant Universe we necessarily rely on the technique of {\it linear}
spectropolarimetry. The only limiting factor is then the collecting
power of the mirror of the largest
telescopes.
\section{2D Wind Predictions}
\noindent
Until 3D radiation transfer models with 3D
hydrodynamics become available, theorists
have necessarily been forced to make assumptions with respect
to either the radiative transfer (e.g. by assuming a power law
approximation for the line force due to \citealt{CAK})
or the hydrodynamics, e.g. by assuming an empirically motivated
wind terminal velocity in Monte Carlo predictions \citep{AL85,V00}. Albeit
recent 1D and 2D models of \cite{MV08,MV14} no longer
require the assumption of an empirical terminal wind velocity.
There are 2D wind models on the market that
predict the wind mass loss predominately emanating
from the equator \citep{FA86,BC93,LP91,Pel00}, whilst other models predict
higher mass-loss rates from the pole, in particular as a result of
the \cite{vonZeipel24} theorem, resulting in a larger polar Eddington factor
than the equatorial Eddington factor \citep{O96,PP00,MM00,MV14}.
The key point is that mass loss from the equator results in more angular
momentum loss than would 1D spherical or 2D polar mass loss, so we need
2D data to test this.
\section{Line polarization versus depolarization}
\begin{figure}
\begin{center}
\includegraphics[width=\textwidth]{noline.pdf}
\caption{Cartoon indicating ``no line effect'. On the left,
polarization spectrum ``triplot'' and a Stokes $QU$ diagram on the right.
A typical Stokes I emission is shown in the lower panel of the triplot,
the \%Pol in the middle panel, while the Position Angle (PA) is
sketched in the upper panel of the triplot. See \citet{V02} for further details.}
\label{fig1}
\end{center}
\end{figure}
\noindent
Whilst {\it circular} Stokes $V$ spectropolarimetry is oftentimes employed
to measure stellar magnetic fields, {\it linear} Stokes $QU$ polarimetry can
be utilized to measure large-scale 2D asymmetry in a stellar wind or any other
type of circumstellar medium, such as a disk. In this sense, the Stokes $QU$
plane plays an analogous role to the interferometric $UV$ plane, with the
additional advantage that it can measure the smallest spatial scales, such
as the inner disk holes of order just a few stellar radii
in pre-main sequence (PMS) stars \citep{V05}, which would otherwise
remain ``hidden'', or the driving region of stellar winds in massive stars, that we explore in the following.
In principle, linear continuum polarimetry would already be able to inform us about the presence
of an asymmetric (e.g. a disk or flattened wind) structure on the sky, but in practice, this
issue is complicated by the roles of intervening circumstellar and/or interstellar
dust, as well as instrumental polarization. The is one of the reasons linear {\it spectro}polarimetry, measuring
the change in the degree of linear polarization across emission lines is such a powerful tool, as ``clean'' or ``intrinsic'' information
can be directly obtained from the $QU$ plane.
The second reason is the additional bonus that it may provide kinematic information
of the flows around PMS as well as massive stars.
\begin{figure}
\begin{center}
\includegraphics[width=\textwidth]{depolariz.pdf}
\caption{Cartoon indication ``depolarization'' or ``dilution''. Note that the
depolarisation across the line is as broad as the Stokes $I$ emission. Depolarisation
translates into Stokes $QU$ space as a linear excursion. See \citet{V02} for further details.}
\label{fig2}
\end{center}
\end{figure}
Figures 1-3 show linear line polarization
cartoons (both in terms of polarization ``triplot'' spectra and Stokes $QU$ planes)
for the case that the spatially unresolved object under consideration is (i)
spherically symmetric on the sky showing ``no line effect'', (ii) asymmetric
showing line ``depolarization'' where the emission line simply acts to ``dilute'' the polarized
continuum, or (iii) cases where the line effects are more subtle, involving
position angle (PA) flips across intrinsically polarized lines.
\begin{figure}
\begin{center}
\includegraphics[width=\textwidth]{narrow.pdf}
\caption{Cartoon indicating a compact source of line photons scattered off a {\it rotating} disk.
Note that the polarisation signatures are relatively narrow compared
to the Stokes $I$ emission. The PA flip is associated with a loop
in Stokes $QU$ space. See \citet{V02,V05} for further details.}
\label{fig3}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=0.48\textwidth]{hd45314.pdf}
\includegraphics[width=0.48\textwidth]{hd120678.pdf}
\caption{H$\alpha$ line polarization ``triplots'' of the Oe stars HD\,45314 and HD\,120678.
HD\,45314 shows a line effect indicating that it
is intrinsically polarized, but HD\,120678 is not intrinsically polarized. See \citet{V09} for further details.}
\label{fig4}
\end{center}
\end{figure}
Whilst the third situation of intrinsic line polarization in a rotating disk has been
encountered in PMS (see \citealt{V05}), it is the second case of ``depolarization'' that
is most familiar to the massive-star community through its application to classical
Be stars, starting as early as the 1970s (see the
various works by Poeckert, Marlborough, Brown, Clarke, and McLean).
Interestingly, the same method has in more recent years also been applied to
Oe stars, the alleged more massive counterparts of Be stars, see Fig.\,4.
Note that although the Oe star HD\,120678 (on the right hand side of Fig.\,4) has a
significant observed level of linear polarization, the
lack of a line effect implies that the object is not intrinsically polarized \citep{V09}.
This could either mean the object is spherically symmetric or that it has a disk that is
too ``pole on'' to provide intrinsic polarization. It is for these reasons vital
to consider a {\it sample} of objects. For Oe stars \cite {V09} found that the incidence of line
effects (1/6) was much lower than for Be stars. This implies
that the chance the Oe and the Be stars are drawn from the same parent distribution
is small, providing relevant constraints on the formation of Be stars.
\section{Survey results of O and Wolf-Rayet winds}
\noindent
We now turn to more massive stars with stronger winds than Oe/Be stars.
Linear spectropolarimetry results have been performed on relatively large
samples (of order 40-100) for both O \citep{H02,V09} and Wolf-Rayet (WR) stars
\citep{H98,V07}, and the key result from these surveys is that the vast
majority of 80\% of them is to first order spherically symmetric.
This is of key importance for the accuracy of mass-loss predictions from 1D models
for rotating stars.
However, the above studies also found a number of interesting {\it exceptions}. With respect to O stars,
\citet{V09} found that certain O-type subgroups involving Of?p and Onfp
class are more likely polarized than the garden-variety of spherical O-stars.
For instance, \citet{V09} highlighted that HD\,108 is linearly polarized, which
may be related to its probably magnetic properties. Indeed, it was later found that
HD\,108 and several other Of?p stars form a magnetic sub-class.
The line effects in the Onfp stars (involving famous objects like $\lambda$\,Cep and
$\zeta$\,Pup) may may involve intrinsic line polarization effects due to the
rapid rotation of this O-type subgroup in addition to (or instead of) depolarization.
Turning to WR stars, \citet{V11} and \citet{G12} uncovered
that the small 20\% minority of WR stars that display a depolarization line effect indicating
stellar rotation are highly significantly correlated with the subset of WR stars that have ejecta nebulae.
These objects have most likely only recently transitioned from a red sugergiant (RSG) or luminous blue variable (LBV)
phase. As these presumably youthful WR stars have yet to spin-down, they are the best
candidate gamma-ray burst (GRB) progenitors identified to date. However, in our own Milky Way these
WR stars are still expected to spin down before explosion (due to WR winds).
However, in lower metallicity ($Z$) environments WR stars are thought to be weaker and
WR stars in low $Z$ environments, such as those studied in the
Magellanic Clouds may offer the best way to directly pinpoint
GRB progenitors \citep{V07}.
\section{Future}
\noindent
In addition to the quest for WR GRB progenitors, there is a whole range of interesting wind physics
to be constrained from linear spectropolarimetry. The main limitation at this point
is still sensitivity. We are currently living in an exciting time as we are
at a point where the possibility of extremely large telescopes (ELTs) may become reality. If these telescopes
materialize with the required polarization optics, we might -- for the first time in history --
be able to obtain spectropolarimetric data at a level of precision that has been feasible
with 1D Stokes $I$ data for more than a century.
It is really important to note that current 3D Monte Carlo radiative transfer
is well able to do the required modelling, but the main limitation is the necessary 3D data!
Another interesting future application will involve polarimetric monitoring.
Whilst we now know that on large scales the 1D approximation is appropriate
for stellar winds, we have also become aware of the intrinsic 3D clumpy nature of stellar winds
on smaller scales (but with macroscopic implications!). In particular the existence of wind
clumps on small spatial scales near the stellar photosphere \citep{Cant09} has
been confirmed by linear polarization variability studies \citep{D07}, but to probe
further -- mapping wind clumps in detail -- we need good monitoring data.
\bibliographystyle{iau307}
|
\section{Introduction}
\label{sect:intro}
The {CoRoT}\footnote{\underline{Co}nvection, \underline{Ro}tation \& planetary \underline{T}ransits} mission has been the first space mission dedicated to the search for transiting planets \citep{2007AIPC..895..201B}. Overviews of the mission were given by \citet{Baglin+2009}, \citet{Deleuil+2011}, and \citet{Moutou+2013}. Statistics on candidates and planet detections have been reported for several fields monitored by {CoRoT} \citep{Carpano+2009,Moutou+2009a,Carone+2012,Erikson+2012,Cabrera+2009,Cavarroc+2012}.
The goals of ground-based follow-up are manifold \citep{Carpano+2009,Carone+2012}, among others the precise derivation of the parameters of a detected planet. While the {CoRoT} light curve constrains the radius of a planet, ground-based radial-velocity (RV) measurements are needed to derive its mass \citep{Moutou+2009a}. Only the relative radius and mass of the planet are constrained. Therefore, the absolute mass and the radius of the host star have to be known very accurately.
False positives play an important role and low-resolution spectroscopy is a step towards their identification. In the case of a false positive, a planetary transit is mimicked by other configurations \citep{Brown2003}. The case of a background eclipsing binary close to the actual target can be resolved by photometry \citep{Almenara+2009,Guenther+2013}. The present work aims at another kind of false positive. A planetary transit can be mimicked by a low-mass star orbiting a giant star \citep[see][]{Cavarroc+2012}. The identification of giant stars relies on the availability of accurate spectral types.
A first clue about the spectral type of {CoRoT} targets is given by the {CoRoT} input catalogue (\textit{ExoDat}\footnote{http://cesam.oamp.fr/exodat/}; \citealp{Deleuil+2009}) which is based on a massive $UVBr'i'$ photometric survey carried out during the mission preparatory phase and has been updated continuously since then. However, in regions of inhomogeneous extinction, it is difficult to obtain spectral class, luminosity class, and extinction simultaneously. Although the photometric classification is correct on average, it can be deviant for individual stars \citep{sebastian12}. Therefore, spectroscopic information is needed to better assess stellar parameters of individual targets \citep[cf.][]{Carone+2012,Gazzano+2010a}.
Precise information is contained in atmospheric parameters like effective temperature and surface gravity. Those are obtained most accurately from high-resolution spectroscopy \citep[e.g.][for {CoRoT}]{Gazzano+2013a}. As the necessary signal-to-noise ratio (SNR) can be hardly attained at high spectral resoluton for faint stars, one needs to resort to low-resolution spectroscopy. Then, a precise measurement of atmospheric parameters is out of reach but spectral classification is still possible.
In practice, spectral classification is done by comparing stellar spectra to template spectra of well-known stars, in particular MK standards. There are different approaches of how to obtain template spectra and to compare them to the target spectra. Computer-based classification has proven efficient. \citet{sebastian12} classified more than 10,000 stars in the fields of CoRoT automatically and showed that an accuracy of one to two sub-classes can be achieved. While templates are commonly taken from libraries taken with other instruments \citep[e.g.][]{2008ApJ...687.1303G,sebastian12}, the classification is considered most accurate when the templates are taken from an internal library, i.e. have been observed with the same instrument and setup as the target spectra \citep{2009ssc..book.....G,2011RAA....11..924W}. Then, the full spectral range is available for comparison and no convolution is needed to match the spectral resolution.
A low-resolution spectrograph attached to an intermediate-size telescope, like the Nasmyth spectrograph ($R\approx1,000$) at the Tautenburg 2m telescope, offers a sufficient dynamical range to observe the faint {CoRoT} targets ($V<16\,$mag) as well as bright nearby template stars. The Nasmyth spectrograph provides a wide spectral range $360-935\,$nm covering a wealth of stellar features which can be used for classification.
In the present study, we follow the initial work of \citet{Guenther+2012} and \citet{sebastian12}. Based on spectra taken with the Nasmyth spectrograph, we explore the capabilities of low-resolution spectroscopy to derive the spectral types of {CoRoT} targets. Of course, the accuracy of spectral classification will depend on the signal-to-noise ratio (SNR) of the target spectra. In the course of the follow-up of planet candidates, a large number of stars has to be classified efficiently and to the accuracy required. Since telescope time is limited and the exposure time scales with the square of the SNR, we are interested in a good knowledge of the required SNR.
The {CoRoT} targets were selected from various internal and public lists, in particular from Deleuil et al. (in preparation) \footnote{The present work includes a statistical study on the type and quality of information available from spectral classification. Therefore, additional follow-up information on particular planet candidates is not included here and a discussion of the status of individual candidates is beyond the scope of the present work (but see Deleuil et al.).}. Most of the {CoRoT} objects are FGK stars \citep{Guenther+2012} and these are the most interesting planet host stars. How well does spectral classification distinguish giant stars and dwarf stars, especially at spectral types F, G, and K?
We followed a two-fold approach. Firstly, we followed the canonical approach of spectral classification \citep{2009ssc..book.....G} and built an internal library of spectral templates with the same spectrograph and setup used to observe the {CoRoT} targets. Secondly, a classification was done using templates from the Indo-US Coud\'e-feed library \citep[CFLIB,][]{valdes04}, selected for its good coverage of spectral types, luminosity classes, and metallicity. We investigate whether this external library gives similar results. Simultaneously, we analyse the impact of the SNR of the target spectra on the results. In addition, we compare the spectral types obtained to previous photometric classifications given by {\textit{ExoDat}}. We identify the merits of the inclusion of a spectroscopic classification over a sole photometric classification.
Section~\ref{sect:temp} explains the selection of stars to build the internal template library. The observation of templates and {CoRoT} targets is described in Sect.~\ref{sect:obs}. Sect.~\ref{sect:red} addresses the data reduction and the coverage of the internal template catalogue is assessed in Sect.~\ref{sect:coverage}. We describe the steps of the spectral classification in Sect.~\ref{sect:class} and the results in Sect.~\ref{sect:results}. The results are discussed in Sect.~\ref{sect:disc} before we conclude in Sect.~\ref{sect:concl}.
\section{Selecting stars for a new library of template spectra}
\label{sect:temp}
The template stars need to fulfil a set of requirements. Of course, the spectral types need to be known very well. Having in mind quantitative spectral classification in future work, we adopted as an additional criterion the availability of accurate atmospheric parameters.
Although quantitative classification is beyond of the scope of the present work, we point out here that it has many advantages over the use of spectral types. There is no need to use a conversion scale, e.g. from spectral type to effective temperature which is intrinsically prone to errors. Furthermore, it promises higher precision and flexibility than the classification by spectral type which is limited by the discreteness of the classification scheme. An extension of the traditional grid of spectral types to include metallicity or even abundance patterns would be too demanding. Instead, atmospheric parameters can be used defining a continuous grid which can be sampled by templates according to the requirements of the classification task. For previous work and reviews, see \citet{Cayrel+1991,Gray+1991,Stock+1999,Malyuto+2001,Bailer-Jones2002, Singh+2002}.
As another criterion, the template stars should be bright so that spectra can be taken quickly with a high SNR. The templates should cover the range of spectral types of interest.
In order to be able to classify most {CoRoT} targets, we selected templates for early-type and Sun-like stars of different luminosity class from the CFLIB. \citet{Wu+2011a} obtained stellar parameters and chemical abundance homogeneously for the CFLIB library.
FGK dwarfs are the best targets to look for planets with {CoRoT}. To fill the grid with templates, we selected well-studied FGK stars from Fuhrmann (1998-2011)\nocite{Fuhrmann1998,Fuhrmann2000,Fuhrmann2004,Fuhrmann2008,Fuhrmann2011}. Several of those are MK standards according to \citet{2009ssc..book.....G}.
The work of Fuhrmann is restricted to solar-type main-sequence and sub-giant stars. However, in the distant {CoRoT} fields, the fraction of early-type stars and giants is relatively high. {CoRoT} covers these luminous stars more completely but there is a lack of cooler main-sequence stars. While early-type stars are covered by the CFLIB, we supplement the set of stellar templates with K giants compiled by \citet{Doellinger2008} who derived the stellar parameters and iron abundance of 62 K giants.
\section{Observations}
\label{sect:obs}
\begin{table}
\caption{\label{tab:specinfo} Journal of observations sorted by the SNR achieved. The first two columns give the CoRoT and Win identifiers. The Win identifier is composed of the CoRoT run (LR=long run, SR=short run, IR=initial run; a=galactic anti-centre, c=galactic centre; plus a consecutive number), the CCD identifier (E1/E2), and a consecutive number. The run ID identifies the spectroscopic runs carried out in February and July 2012/2013. The $R$ band magnitude is followed by the number of spectra taken per object. The last column gives the SNR of the co-added spectra.}
\begin{tabular}{rllrrr}
\hline
{CoRoT} ID&Win ID&run ID &$R$&\#&SNR\\
\hline
\input{specinfo}
\hline
\end{tabular}
\end{table}
\begin{table*}
\caption{\label{tab:templates} Stellar templates observed with the Nasmyth spectrograph and their spectral type. The second column gives the run ID which identifies the spectroscopic runs carried out in February and July 2012/2013.}
\begin{tabular}{rll|rll|rll}
\hline
HD number & run ID & spectral type&HD number & run ID & spectral type&HD number & run ID & spectral type\\
&&&&&&&&\\
\hline
\input{templates_mod}
\hline
\end{tabular}
\end{table*}
All spectra were taken with the low-resolution long-slit spectrograph mounted at the Nasmyth focus of the 2m Alfred Jensch telescope in Tautenburg (Germany). A slit width of $1''$ was used to ensure a resolution of $\approx1,000$. The V200 grism was chosen as dispersing element, thus covering the visual wavelength range from 360 to 935\,nm. It is a BK7 grism with 300 lines per millimeter and a dispersion of $225\,\mathrm{\AA}\mathrm{mm}^{-1}$. The detector is a SITe\#T4 CCD with 2048x800 pixels and a pixel size of $15\,\mathrm{{\mu}m}$. We used channel A with a gain of $1.11$\,$\mu$V per electron.
{CoRoT} monitors several stellar fields consecutively for up to 40\,days (short runs) or for up to 150\,days (long runs). These fields are selected in two opposite directions in the sky where the galactic plane crosses the equatorial plane. These are the so-called galactic “centre” and “anti-centre”-eyes of {CoRoT}. Seasonal observations of the {CoRoT} fields are possible in winter (galactic anti-centre) and summer (galactic centre) despite the high latitude of the observing site of $51^\circ$N. The {CoRoT} fields can be observed at an air mass of approx. 1.5. A total of 42 {CoRoT} targets covering a brightness range of $R=11.0-15.5$ was observed in February and July 2012/2013 when the visibility of the {CoRoT} eyes was best. Table~\ref{tab:specinfo} presents the journal of observations.
In addition, we have observed a set of 149 template stars (Table~\ref{tab:templates}). They are bright so that the SNR is high. Exposure times range from a few seconds for the brightest template stars to 30\,min for the faintest {CoRoT} targets. In general, at least two spectra were taken per target in order to remove cosmics. While the usually faint {CoRoT} targets are observable only under good conditions for a few hours per night, the bright template stars are distributed over the whole sky. They perfectly fill the observing time when {CoRoT} targets are not visible or when conditions are bad.
In order to estimate the bias level and the read-out noise, dark frames were taken with closed shutter and zero exposure time. In order to assess the pixel-to-pixel variation of sensitivity, the wavelength-dependent transmission of the spectrograph, and bad or hot pixels, flat-field frames were taken pointing the telescope to a flat-field screen which is installed in the dome and illumated by an incandescent lamp.
In principle, the wavelength calibration can be done using sky emission lines in the long-slit spectra. However, the sparseness of sky emission lines in the blue part of the spectra does not permit a precise calibration in this region. It is the blue part of the spectrum which displays a wealth of features and is very valuable for spectral classification. Therefore, spectra of gas discharge lamps (He and Kr) were used for wavelength calibration. At least one set of gas discharge exposures was taken per observing run. A wavelength precision of a few {\AA} is achieved this way and is sufficient for the purpose of spectral classification.
To estimate the signal-to-noise ratio (SNR), we assumed pure photon noise accounting for read-out noise. The total system gain used depends on the amplifier setting chosen. In the present case (setting '20'), we estimated a total system gain of 3.2 electrons per data unit using the full well at the noise roll-over specified by the manufacturer (84,000 electrons)\footnote{The full well was scaled to the saturation level obtained at setting '50' and then rescaled to setting '20'.}. The readout-noise amounts to 7.8 electrons. This estimate of the SNR is fully sufficient to compare the spectra since all spectra were taken with the same instrument and identical settings. The median is used to distinguish good spectra with high signal and bad spectra with a high noise level.
\section{Data reduction}
\label{sect:red}
The long-slit spectra have been reduced using the data reduction and analysis system IRAF\footnote{IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation.} \citep{iraf86,Tody1993}. This procedure comprises bias subtraction, an automatic removal of cosmic rays, and removal of pixel-to-pixel variations by flat-field correction.
All spectra have been wavelength-calibrated using He and Kr spectra. Using IRAF tasks, the wavelengths of prominent lines have been identified and used to derive a wavelength solution which has been applied to all stellar spectra. The spectrograph is mounted at the fork of the telescope so that instrument flexure causes shifts depending on the exact orientation of the telescope. Therefore, sky emission lines have been employed to apply corrections. For each frame with a long exposure time of 20 minutes at least, the sky lines are bright enough to use them as calibration source \citep{osterbrock96}. They have been used to derive offsets to correct the wavelength scale. The corrections were small and of the order of the resolution limit.
\begin{figure}[h]
\centering
\includegraphics[width=8cm]{plot_spec_sky}
\caption{The black line shows a Nasmyth spectrum of the G0V star \object{HD\,157214}. For comparison, a library spectrum with medium resolution \citep{valdes04} is shown (light grey). To match the SED, the template spectrum was used to adjust the flux of the Nasmyth spectrum. Both spectra were normalised to unity at $5550\,${\AA}. In addition, the subtracted night sky spectrum of the site is shown (grey) and is offset w.r.t. the stellar spectra for clarity. The most important spectral features are indicated.}
\label{fig:sky}
\end{figure}
The sky background could be subtracted easily from the long-slit spectra. For this purpose, long exposures provide an excellent source to analyse the night-sky emission in Tautenburg. Figure~\ref{fig:sky} shows the Nasmyth spectrum of a Sun-like star together with the subtracted sky-background.
The most prominent features in the night-sky spectrum are the O\,I, Hg\,I, and Na\,I atomic lines, the broad continuum centred at 5890\AA, and the OH bands in the red part \citep{Osterbrock+1992,Slanger+2003}.
These lines originate from the night glow as well as from artificial sources.
The broad continuum centered at 5890\,{\AA} is characteristic of the presence of artificial light and in this case originates from high-pressure sodium lamps used in the nearby city of Jena, Germany.
The flux calibration of the spectra of {CoRoT} targets was not possible because of the spectrograph design and the high airmass of {CoRoT} targets. The orientation of the spectrograph slit cannot be adjusted and the slit-rotation of the spectrograph depends on the hour angle of the object. This leads to a dependency of the detected flux level on atmospheric refraction. In addition, the star is slightly moving on the slit and the exact location on the slit cannot be reproduced. Therefore, the stellar continuum could not be recovered and used for classification. Instead, one has to rely on spectral lines only and the continuum has to be adjusted to match the continuum of the template (Fig.~\ref{fig:spec_class}). Hence, the comparison is restricted to the strength and the profile of selected absorption lines.
\section{The coverage of the new template library}
\label{sect:coverage}
The CFLIB and the new internal template library have many stars in common. 54 CFLIB stars have been included in the sample chosen from Fuhrmann (1998-2011). The Fuhrmann sample does not cover all spectral types. Therefore, 55 stars have been chosen from the CFLIB catalogue and reobserved for the new library. In total, 109 CFLIB stars have been reobserved.
Within the present work, 149 template spectra of bright stars have been taken. MK spectral types have been adopted from \citet{valdes04} when available and from the SIMBAD database otherwise.
Table~\ref{tab:grid_spt} shows the coverage in spectral type. The catalogue is being filled continuously but a good coverage has already been achieved for dwarf stars. Particularly, FGK dwarfs are densely covered. A finely graduated classification becomes feasible in the FGK regime.
\begin{table}
\caption{\label{tab:grid_spt} The coverage of the new template library in spectral type. At each spectral type, the number of templates is given.}
\begin{center}
\begin{tabular}{l|rrrrrrr}
\hline
spectral&\multicolumn{7}{c}{luminosity class}\\
class&I&II&II-III&III&III-IV&IV&V\\
\hline
\input{grid_spt}
\hline
\end{tabular}
\end{center}
\end{table}
\section{Computer-based classification}
\label{sect:class}
The spectral types of the CoRoT targets have been obtained using a computer-based classification.
The software applied (described in \citealp{sebastian12}) compares all target spectra to a
library of template spectra. Each target spectrum is matched to template spectra
adjusting the radial velocity shift, the level of the continuum, and the slope of the continuum. We selected spectral chunks that contain sensitive lines for spectral classification thus removing
parts that are not useful for classification (see Fig.~
\ref{fig:spec_class}).
Afterwards the ${\chi}^2$ is calculated and compared for each template spectrum.
The five best-matching templates are validated by visual inspection. This
validation rules out false classifications due to low S/N or stellar
activity. The spectral type of the best-matching validated template is adopted for the
target spectrum.
\citet{Guenther+2012} found that the error bar of this classification
method depends on the spectral type and the template library used. They analysed the
accuracy of the method on a sample of more than 3000 stars and found that the accuracy of this method for low resolution spectra is
on average two subclasses. It is slightly better for
early-type stars (1.3 subclasses) and less accurate for solar-type
stars (2-3 subclasses). The classification is not affected by rotational broadening since the spectral resolution of the Nasmyth spectrograph is too low.
Firstly, we used the internal template library taken
with the Nasmyth spectrograph. In the second approach, we
used a set of 281 template spectra provided by CFLIB
(Valdes et al. 2004)\footnote{These CFLIB templates have a high SNR of $100$ at least and cover almost the full range of spectral types.}.
To compare the {CoRoT} target spectra directly with the CFLIB templates, we have to ensure that the resolution, the wavelength range, and the continuum flux level are roughly the same. Since the Nasmyth spectra have been taken with a resolution of $\sim5$\,{\AA} FWHM (at $5500$\,{\AA}), we convolved the CFLIB templates ($\sim1$\,{\AA} FWHM) with a Gaussian kernel.
For the comparison with internal templates, the full wavelength range from $3600$ to $9350\,${\AA} can be used in principle since the target spectra have been obtained with the same spectrograph. However, especially the red part of the covered wavelength region contains telluric bands. The strength of these bands is variable. Although a correction of telluric bands is possible in principle, the scarcity of useful features does not justify this effort. The strongest telluric oxygen bands appear at $6884\,${\AA} \citep{catanzaro97}. Therefore, the wavelength region applied here starts in the blue at 3950\,{\AA} where we get sufficient signal and ends at 6800\,{\AA} at the blue edge of the oxygen bands. In the extreme red, there is a very short range of $\approx400$\,{\AA} not strongly affected by telluric absorption. This region contains three prominent CaII lines ($8498$, $8542$, and $8662\,${\AA}) at almost all spectral types which can be used for spectral classification \citep{2009ssc..book.....G}. For early-type stars, some HI lines of the Paschen series were included in addition. In spite of some night-sky emission lines that could affect the spectra (see Fig.~\ref{fig:sky}), we chose the region from 8400 to 8880\,{\AA} in addition to the blue part ($4200-6800${\,\AA}) for comparison (Fig.~\ref{fig:spec_class}). For spectra with low SNR, we had to set the blue edge to $4800\,${\AA} to skip the blue part with very low signal.
When using the external CFLIB templates, the wavelength range is limited to the overlapping region of the {CoRoT} target spectra and the CFLIB templates. In particular, the very red part with the Ca\,II triplet and the Paschen series cannot be used. Moreover, parts with strong telluric absorption had to be excluded again and once more, the wavelength range had to be reduced to $4800-6850$\,{\AA} in the case of noisy spectra.
The design of the spectrograph does not allow one to reproduce the continuum flux (Sect.~\ref{sect:red}) so that the continuum flux could not be used for classification. Therefore, the {CoRoT} spectra and the Nasmyth templates were normalised and seven chunks (five in the case of low SNR, resp.) were compared separately (Fig.~\ref{fig:spec_class}). This way, the influence of the unknown absolute flux level and the extinction was optimally removed. This method ensures that the spurious continuum flux levels of both spectra do not distort the results of the classification.
The CFLIB templates are not flux-calibrated either but \citet{valdes04} recovered the SED of the CFLIB spectra from spectrophotometric templates. In order to compare the Nasmyth spectra and the CFLIB templates, they were normalised again in an appropriate way and compared in different chunks.
\begin{figure}[h]
\centering
\includegraphics[width=4cm,angle=270]{spec_class.eps}
\caption{Comparison of a target spectrum (grey line) with the
best-matching template (black line). The vertical lines mark the borders of chunks which are compared separately and used to adjust the continuum level of the spectra.}
\label{fig:spec_class}
\end{figure}
Figure~\ref{fig:spec_class} shows the comparison of a {CoRoT} target to its best-matching Nasmyth template. In this case, the target spectrum matches the spectrum of HD 121560 perfectly (spectral type F6V).
\section{Results}
\label{sect:results}
\begin{table*}
\caption{\label{tab:classinfo} Results of spectral classification (sorted by SNR). The table contrasts the classification obtained via two different sets of templates -- the internal library obtained by us with the Nasmyth spectrograph and the template library of \citet{valdes04} (CFLIB). \object{LRc10\_E2\_1984} is an M star for which no template was taken with the Nasmyth spectrograph.}
\renewcommand{\tabcolsep}{0.3em}
\footnotesize
\centering
\begin{tabular}{lrl|ll|ll}
\hline
&&&\multicolumn{2}{c|}{templates taken with }&\multicolumn{2}{c}{templates taken from} \\
&&&\multicolumn{2}{c|}{the Nasmyth spectrograph}&\multicolumn{2}{c}{\citet{valdes04}}\\
Win ID&SNR&spec. type&templ.&spec.&templ.&spec.\\
&&(Exodat)&&type&&type\\
\hline
\input{classinfo}
\hline
\end{tabular}
\end{table*}
The results of the classification are listed in Table~\ref{tab:classinfo}. First of all, there are some interesting qualitative findings. In neither case, the best-matching external template originates from the same star as the best-matching internal template, in spite of a high number of stars common to both libraries. Also it occurs that an external template spectrum matches another {CoRoT} object. The CFLIB spectrum of HD\,120136 matches LRc05\_E2\_3718 while the Nasmyth spectrum of HD\,120136 matches the spectrum of LRc07\_E2\_0182. Apparently, the CFLIB spectrum of HD\,120136 is as different from the Nasmyth spectrum as is the difference between the spectra of LRc07\_E2\_0182 and LRc05\_E2\_3718. Although both stars are late-F or early-G type stars, we note, that the SNR of the spectrum of LRc05\_E2\_3718 is only half of that of LRc07\_E2\_0182. Interestingly, the Nasmyth spectrum of HD\,84737 matched Nasmyth spectra of as many as five {CoRoT} objects. Again, the SNR is obviously important since all the five target spectra suffer from low signal!
\begin{figure}[h]
\centering
\subfigure[\label{fig:sn_spec_nas}external templates (CFLIB) - internal templates]{\includegraphics[width=7cm]{sn_spec_nas}}
\subfigure[\label{fig:sn_spec_exo}photometric classification ({\textit{ExoDat}}) - internal templates]{\includegraphics[width=7cm]{sn_spec_exo}}
\subfigure[\label{fig:sn_spec_val}photometric classification ({\textit{ExoDat}}) - external templates (CFLIB)]{\includegraphics[width=7cm]{sn_spec_val}}
\caption{\label{fig:sn_spec} Residuals of classifications when using external templates and photometric classifications. The difference of spectral types (in sub-classes) is displayed vs. the SNR. The vertical line shows the median SNR found in Sect.~\ref{sect:obs}, and is used to distinguish good and bad spectra. The $2\sigma$ outliers among the good spectra are highlighted by filled circles and discussed in the text.}
\end{figure}
\begin{figure}[h]
\centering
\includegraphics[width=7cm]{hist_spec_nas}
\caption{\label{fig:hist_spec} Histogram of residuals of classifications when using the CFLIB library (corresponding to Fig.~\ref{fig:sn_spec_nas}). The difference of spectral types is shown in units of sub-class. The filled style distinguishes spectra with good (grey) and bad signal (hatched) according to the division line shown in Fig.~\ref{fig:sn_spec}.}
\end{figure}
The qualitative discussion above shows that the SNR plays a role. A more quantitative presentation of the data confirms that the agreement depends on the noise level of the {CoRoT} target spectra (Fig.~\ref{fig:sn_spec_nas}). Fig.~\ref{fig:hist_spec} presents another view on the data by projecting along the noise axis and showing the distribution of the residuals separately for the good and bad spectra. The spread tends to increase with decreasing signal, i.e. the strongest discrepancies are usually encountered for spectra of least quality. At highest signal, the discrepancies tend to vanish.
The mean difference in spectral type is as low as 2 sub-classes when considering the good spectra only. Remarkably, hardly any systematic offsets or trends are seen in our sample.
The distribution of residuals appears far from normal in the case of the bad spectra. In the case of the good spectra, however, an identification of outliers seems possible. There are two outliers in spectral type, LRc09\_E2\_0308 and LRc05\_E2\_3718. The deviation cannot be readily explained since the SNR is reasonably good.
The spectroscopic classifications were compared with photometric classifications taken from the online version of {\textit{ExoDat}} as of May 16$^\mathrm{th}$, 2014. We note that the scatter is increased for bad spectra (Figs.~\ref{fig:sn_spec_exo}, \ref{fig:sn_spec_val}). But even for good spectra, the scatter is dramatically larger and we note that photometry tends to assign earlier spectral types. The level of agreement is the same whatever the choice of the template library - internal or external.
\begin{table}
\caption{\label{tab:lcnas} Contingency table showing the agreement of luminosity classifications obtained with the external CFLIB template library and the internal Nasmyth library. The number of classifications based on good spectra is shown and succeeded in brackets by the number of classifications based on bad spectra. Good and bad spectra are distinguished using the median of the SNR (cf. Fig.~\ref{fig:sn_spec}). The solid lines in the table divide giant-like classes and dwarf-like classes.}
\centering
\begin{tabular}{l|rrrr|rr}
\hline
&\multicolumn{6}{c}{Nasmyth templates}\\
CFLIB templates&I&II&III&III-IV&IV&V\\
\hline
I&&&1&&(1)&\\
II&&&(1)&&&(1)\\
III&&&1(3)&&1&1(2)\\
III-IV&&&&&&\\
\hline
IV&&&1&&1&4(2)\\
V&&(2)&&1&2&8(8)\\
\hline
\end{tabular}
\end{table}
\begin{table}
\caption{\label{tab:lcexo} Contingency table showing the agreement of spectroscopic and photometric luminosity classifications: internal Nasmyth catalogue vs. photometric Exodat classification. The layout follows Table~\ref{tab:lcnas}.}
\centering
\begin{tabular}{l|rrrr|rr}
\hline
&\multicolumn{6}{c}{Nasmyth templates}\\
Exodat&I&II&III&III-IV&IV&V\\
\hline
I&&&&&(1)&\\
II&&&(1)&&&\\
III&&&3(2)&&&(2)\\
III-IV&&&&&&\\
\hline
IV&&&&&2&8(6)\\
V&&(2)&(1)&1&2&5(5)\\
\hline
\end{tabular}
\end{table}
\begin{table}
\caption{\label{tab:lcval} Contingency table showing the agreement of spectroscopic and photometric luminosity classifications: CFLIB catalogue vs. photometric Exodat classification. The layout follows Table~\ref{tab:lcnas}.}
\centering
\begin{tabular}{l|rrrr|rr}
\hline
&\multicolumn{6}{c}{CFLIB templates}\\
Exodat&I&II&III&III-IV&IV&V\\
\hline
I&(1)&&&&&\\
II&&(1)&&&&\\
III&1&&1(2)&&1(2)&\\
III-IV&&&&&&\\
\hline
IV&&&(1)&&3&7(5)\\
V&&(1)&2(2)&&2&4(5)\\
\hline
\end{tabular}
\end{table}
\citet{sebastian12} showed that for solar-type stars the computer-based spectral classification cannot distinguish between main-sequence stars and sub-giants but between main-sequence stars and giants or super-giants. Therefore, we distinguish giant-like (I, II, III, III-IV) and dwarf-like luminosity classes (IV, V); also because of the low numbers in the present work. The luminosity classification based on the two different sets of templates is compared in a statistical sense (Table~\ref{tab:lcnas}). Each field contains the number of {CoRoT} targets with corresponding classifications. Again, good and bad spectra are distinguished. The total number of good(bad) spectra is 21(20). 15(10) stars are dwarf-like according to either set of templates while there is agreement on a giant-like luminosity class for 2(4) stars. Still, there is substantial disagreement for 4(6) stars. Two(four) dwarf-like stars are given giant-like classifications when using the CFLIB templates and 2(2) giant-like stars are assigned dwarf-like classifications. In summary, only 19\,\% of the classifications of good spectra are not reproduced by the external library compared to 30\,\% of the classifications of the bad spectra.
The contingency tables are reproduced for the comparisons with the photometric classification from {\textit{ExoDat}} (Tables \ref{tab:lcexo}, \ref{tab:lcval}). When comparing the internal library with {\textit{ExoDat}}, there are 1(6) off-diagonal elements. Comparing the CFLIB to {\textit{ExoDat}}, there are 3(6) off-diagonal elements.
For an update of the status of {CoRoT} candidates with a spectroscopic classification, we recommend a rather conservative approach in order to not discard planet candidates prematurely. For the present sample and for the discussion below, we adopt a giant-like luminosity class only in the case that both template libraries yield a giant-like classification. Five out of these 6 stars were classified giants by photometry, too. Furthermore, we identified seven early-type stars (F3 or earlier), six out of these also early-type according to {\textit{ExoDat}}.
However, there is disagreement in several cases. Intriguingly, seven stars are early-type according to photometry but late-type according to spectroscopy, affecting almost half of the photometric early-type targets. Furthermore, 2 out of 9 giants are actually dwarf stars. There is no single late-type star which turned out early-type. Only one early-type dwarf turned out a giant.
\section{Discussion}
\label{sect:disc}
We obtained target and template spectra with the same instrument in order to avoid any systematic effects which might be introduced by the use of another instrument. This involves a major effort since a large number of spectral types has to be covered. As many as 149 template stars have been observed to cover different luminosity classes and chemical abundance patterns. Although the coverage is not yet complete, it is very dense.
One of the main goals was to find out whether an external template catalogue, here the CFLIB, can be used to reproduce the results of a classification with an internal library. The best-matching template stars found in the present work show that the outcome mainly depends on the SNR of the target spectra. At sufficiently high SNR ($\ge100$), the mean difference in spectral type is still within the internal uncertainties of the method of a few sub-classes \citep{Guenther+2012,sebastian12}. Also a discrepancy of the luminosity classification occurred more often in the case of bad target spectra. Although this assessment is based on small numbers, too, it ascribes at least part of the discrepancy to noise.
The agreement in spectral class for good target spectra gives high confidence in the use of external templates. In contrast, the photometric classification deviates by up to an entire spectral class. Nevertheless, as the scatter w.r.t. photometry increases for spectra with bad SNR, we conclude - as is expected - that the spectral classification is not able to provide accurate spectral types if the signal is too low. The photometric {\textit{ExoDat}} classification of luminosity performs as well as a spectral classifiation employing the external library. In this respect, the present work extends the work of \citet{sebastian12} who derived spectral types from AAO spectra using templates from CFLIB and who compared the results to {\textit{ExoDat}} classifications.
The comparison of the spectroscopic and photometric classifications shows that photometry favours early-type classifications (cf. Fig.~\ref{fig:sn_spec}). Half of them turned out late-type stars which is roughly in line with the findings of \citet{sebastian12} who showed that 30\,\% of the photometric A and B stars are actually late-type stars. Planet search campaigns often remove early-type targets from the list of candidates since the radial velocity follow-up of such objects can be challenging. Moreover, the low-resolution spectroscopy was able to identify 2 dwarf stars among 9 photometric giants. Giant stars are excluded from follow-up since the transit signature would be due to a low-mass star rather than a planet. This way, low-resolution spectroscopy can recover good candidates among early-type targets (50\,\%) and giant stars (25\,\%) discarded otherwise. For the present work, we studied 17 late-type dwarf stars which are particularly interesting for follow-up. None of these turned out a giant so that photometry performs much better in this case.
\section{Summary and Conclusions}
\label{sect:concl}
Within the ground-based follow-up of {CoRoT} targets we took low-resolution spectra of 42 objects with the low-resolution Nasmyth spectrograph in Tautenburg.
The spectra of the {CoRoT} targets were classified using two different sets of templates, an internal template library taken with the same instrument as the target spectra and the external CFLIB library. Although the internal library comprises spectra of 149 stars, the coverage is not fully complete and will be complemented in upcoming observing runs. The use of the new set of templates to refine preliminary classifications is intriguing. The new template grid is densest in the regime of F and G dwarfs which is most promising for planet detections with {CoRoT}.
We found that the use of external library spectra yields similar results when the signal-to-noise ratio of the target spectrum is sufficiently high ($\ge\,100$). However, the luminosity classification with an external library does not perform better than a photometric classification. Therefore, it seems feasible to resort to a less costly photometric luminosity classification when an internal template library is not available.
As an aside, we know the atmospheric parameters of the matching templates. Therefore, a quantitative classification by atmospheric parameters becomes feasible, including effective temperature and surface gravity along with chemical abundances. This might disentangle the well-known degeneracy in spectral classification due to unknown metallicity that may certainly affect the present work. Also, the follow-up of transit-planet candidates would benefit a lot from quantitative classification. However, there are major uncertainties. At first, systematic tests will be needed to assess the accuracy of the quantitative spectral classification, e.g. by reproducing the atmospheric parameters of stars selected from the template library.
The present work highlights the importance of spectroscopic follow-up of {CoRoT} candidates when only a photometric classification has been available before. With modern multi-object spectrographs at hand (e.g. AAO, LAMOST), low-resolution spectroscopic characterisation should become an indispensable part of any effort to follow-up on planet candidates identified in wide-field surveys.
\acknowledgements
The selection of {CoRoT} targets is based on contributions by P. Bord\'e, F. Bouchy, R. Diaz, S. Grziwa, G. Montagnier, and B. Samuel within the detection and ground-based follow-up of CoRoT candidates. M.A. was supported by DLR (Deutsches Zentrum f\"ur Luft-und Raumfahrt) under the project 50 OW 0204. We would like to thank the workshops and the night assistants at the observatory in Tautenburg, Germany. This research has made use of the ExoDat Database, operated at LAM-OAMP, Marseille, France, on behalf of the CoRoT/Exoplanet program. This research has made use of the SIMBAD database, operated at the CDS, Strasbourg, France, and NASA’s Astrophysics Data System Bibliographic Services.
|
\section*{Acknowledgments}
We would like to thank K.-I. Imura and Y. Yoshimura for fruitful discussions on
the Wilson-Dirac model.
This work was supported by Grants-in-Aid for Scientific Research (KAKENHI) Numbers
25400388 (TF),
25610101(YH), and 26247064
from Japan Society for the Promotion of Science (JSPS).
|
\section{Introduction}
Low-density parity check codes (LDPC) were introduced by Gallager \cite{gallager62low} in the early 1960s. Decoding of LDPC codes, by log-likelihood ratio sum-product algorithms (LLR-SPA), are proven to achieve excellent capacity performance, by approaching the Shannon bound \cite{mackay1999good}. However, the drawbacks of LLR-SPA, namely, the high complexity and sensitivity to linear scaling, are solved by the Min-Sum algorithm \cite{fossorier1999reduced}.
Scaled Min-Sum \cite{chen2002density} is a modification of Min-Sum algorithm, where a scaling factor is used to decrease the error introduced by using the approximate minimum operation.
Scaled Min-Sum (with constant scaling factor) is suitable for regular LDPC codes. However, irregular LDPC codes require different scaling technique \cite{zhang2005improved,lechner2006improved,ahmed14CCNC}.\\
In \cite{zhang2005improved}, a two-dimensional (2D) correction of the min-sum was proposed. In this algorithm, different scaling factors are required for different check node degrees and variable node degrees. Consequently, the algorithm requires the calculation of two scaling factor vectors \underline{\textbf{$\alpha$}} and \underline{\textbf{$\beta$}} with length equal maximum check node's degree and variable node's degree respectively. These vectors are optimized by parallel differential optimization of Density Evolution (DE).\\
In \cite{lechner2006improved}, different scaling factor per iteration is proposed for irregular LDPC codes. Different scaling factor per iteration technique has good performance for irregular LDPC codes. However, adding these scaling factors requires complex calculation steps in designing stage, and requires extra storage to store the scaling factor value of each iteration.\\
In \cite{ahmed14CCNC}, we proposed simplified variable-scaled min-sum (SVS-min-sum) decoding technique. This algorithm uses simply implemented heuristic technique to update the scaling factor with iterations. It is simpler than both variable scaling factor \cite{lechner2006improved} and 2D correction Min-Sum \cite{zhang2005improved} in implementing and designing. Simulation results show that SVS-min-sum has lower Bit Error Rate (BER) than constant scaling factor for many LDPC codes \cite{ahmed14CCNC}. SVS-min-sum algorithm starts the scaling factor sequence with a constant value equals 0.5. This restriction decreases its performance and makes it unsuitable for some codes.\\
In this paper, we introduce a generalization of the SVS-min-sum algorithm by removing the restriction of starting the scaling sequence from 0.5. This generalization leads to better performance than constant scaling for all codes. In fact, constant scaling can be seen as a special sub-optimized version of the proposed algorithm as shown in section IV.
We apply Nelder-Mead optimization \cite{nelder1965simplex} on DE to jointly optimize the initial scaling factor and updating step of the scaling sequence. Simulation results illustrate the improvement of the proposed algorithm in both BER performance and decoding latency over other scaling strategies.\\
The rest of the paper is organized as follows: Section II presents the necessary background on the SPA, Min-Sum, Scaled Min-Sum, Variable Scaled Min-Sum and SVS-min-sum algorithms. Section III presents the generalized SVS-min-sum algorithm. The simulation results are displayed and discussed in Section IV. Finally, the paper is concluded in section V.\\
\section{REVIEW OF THE SPA AND MIN-SUM based ALGORITHMS}
An $(n,k)$ LDPC code is a binary code characterized by a sparse parity check matrix $\mathbf{H} \in \mathbb{F}_2^{m\times n}$ where $m=n-k$. It can be represented by a Tanner graph which contains variable nodes $j \in \{1..n\}$ and check nodes $i\in \{1..m\} $ . We denote the set of variable nodes connected to a certain check node $i$ as $V\{i\}$. Furthermore, the set $V\{i\}/j$ denotes the set of variable nodes connected to check node $i$ excluding $j$. Similarly, the set of check nodes connected to a certain variable node $j$ is denoted by $C\{j\}$. $C\{j\}/i$ denotes the set of check nodes connected to the variable node $j$ excluding $i$. \\%Coded bits of LDPC code are QAM modulated and transmitted over an Additive White Gaussian Noise (AWGN) channel. The complex received symbols $r_{l}$ are demodulated by soft output QAM demodulator. The soft output of QAM demodulator $U_{ch_j}$ represents the channel $LLR$ of bit number $j$ which is transmitted over constellation point number $l$ in bit position $z$. So $j=(l-1)*log_{2}(M)+z$. $U_{ch_j}$ is the input sequence of LDPC decoder which uses one of message forwarding decoding algorithms. \\
LDPC codes are efficiently decoded by message passing decoding algorithms. The main idea behind all message passing algorithms is processing the received symbols iteratively in concatenated steps that can be seen over the Tanner graph as horizontal step followed by vertical step. In this section, we review some message passing decoding algorithms that are either used for comparision or as the starting point of our modified algorithm.
\subsection{Sum-Product algorithm (SPA)}
One iteration of the tanh-based SPA is described in the following steps:-
\begin{enumerate}
\item \textit{Initialization step:}
The $LLR$ of bit number $j$ is initialized with its channel $LLR$ $(U_{ch_j})$. These initial values are used as $v_{j \to i}$ , messages from variable node $j$ to check nodes $i$ $\forall$ $i \in C\{j\}$.
\item \textit{Horizontal step: }
At each check node $i$, messages $v_{j \to i}$ (which come from variable nodes $V\{i\}$) are used to calculate the reply messages $U_{i \to j}$ for all $j \in V\{i\}$ by \eqref{eq:1}.
\begin{equation} \label{eq:1}
U_{i \to j}=2 \times tanh^{-1} \bigg( \prod_{\substack{j' \in V\{i\}/j}} tanh\frac{v_{j' \to i}}{2}\bigg)
\end{equation}
\item \textit{Vertical step: }
At each variable node $j$, messages $U_{i \to j}$ are used to calculate the reply messages $v_{j \to i}$ for all $i \in C\{j\}$ by \eqref{eq:2}.
\begin{dmath} \label{eq:2}
v_{j \to i}=U_{ch_j}+\sum_{\substack{i'\in C\{j\}/i}} U_{i'\to j}
\end{dmath}
\item Decision step\\
For each variable node $j$, its $LLR$ is updated by \eqref{eq:3}
\begin{dmath} \label{eq:3}
LLR_{j}=U_{ch_j}+\sum_{\substack{i'\in C\{j\}}} U_{i'\to j}
\end{dmath}
\end{enumerate}
The LLR values are applied to the hard decision to decide on the bit value to be 1 if $LLR_{j}<0$ and zero otherwise. The syndrome is calculated and checked; if it is all-zero vector, this word is successfully decoded, otherwise, if the syndrome condition is not satisfied, the decoder proceeds to the next iteration. This process continues till either the code word is successfully decoded or the maximum iterations are exhausted.
\subsection{Min-Sum algorithm}
The Min-Sum algorithm follows the same steps as SPA. It only approximates the horizontal step calculation by minimum operation as shown in \eqref{eq:4} \cite{fossorier1999reduced}
\begin{dmath} \label{eq:4}
U_{i \to j}=\prod_{\substack{j'\in V\{i\}/j}} sign(v_{j'\to i})*\min_{\substack{j' \in V\{i\}/j}} {|v_{j' \to i}|}
\end{dmath}
Min-Sum is easier to implement, as it gets rid of the tanh(.) calculation. However, the approximation of the tanh(.) to the min(.) leads to some loss of performance compared to the tanh-based SPA algorithm. This loss of performance is partially recovered by Scaled Min-Sum algorithm.
\subsection{Scaled Min-Sum algorithm}
In order to decrease the gap between the min-sum and the tanh-based SPA algorithms, a constant scaling factor ($\alpha < 1$) is applied to the check node updating equation \eqref{eq:4}. In other words, converts the Horizontal step to \eqref{eq:5}
\begin{dmath} \label{eq:5}
U_{i \to j}=\alpha \prod_{\substack{j'\in V\{i\}/j}} sign(v_{j'\to i})*\min_{\substack{j' \in V\{i\}/j}} {|v_{j' \to i}|}
\end{dmath}
This scaling factor is optimized to maximize the performance of Scaled Min-Sum algorithm.
\begin{figure}[!t]
\centering
\includegraphics[width=3.5in]{SVS_diagram.png}
\caption{Circuit representation of SVS scaling}
\label{SVS_block}
\end{figure}
\begin{table}[!t]
\renewcommand{\arraystretch}{1.3}
\caption{Calculation of scaling factor value of each iteration using SVS-min-sum algorithm}
\label{table_I}
\centering
\begin{tabular}{|c|c|}
\hline
Iteration index $i$ & $\alpha$ \\
\hline
$1 \to S$ & 0.5 \\
\hline
$(S+1) \to 2S$ & 0.75 \\
\hline
$(2S+1) \to 3S$ & 0.875\\
\hline
$(3S+1) \to 4S$ & 0.9375 \\
\hline
\end{tabular}
\end{table}
\subsection{SVS-min-sum algorithm} \label{sub:SVS}
Changing the scaling factor with iteration for irregular LDPC codes is used in \cite{lechner2006improved}. Despite of performance enhancement of this variable scaled min-sum algorithm, it requires extra storage because we need different scaling factor value per iteration, associated with different mutual information into passed messages per iteration \cite{lechner2006improved}. The general fractional values (taken by the scaling factors) make the multiplication operation complex to implement. We proposed in \cite{ahmed14CCNC} an SVS-min-sum algorithm addresses the particular point of simply per-iteration updated scaling rule.\\
As stated in \cite{lechner2006improved}, the scaling factor should increase exponentially with iterations and its final value is 1. So we approximate the scaling factor sequence to a stair sequence which is updated every $S$ iterations, increase exponentially and easy to implement. The variable scaling factor can be calculated as:
\begin{equation} \label{eq:6}
\alpha=1-2^{-\lceil i/S \rceil}
\end{equation}
Where $\lceil i/S \rceil$ is the first integer greater than or equal to $i/S$ . $i$ is the iteration index which takes values $\{1, 2, 3 ...\}$. By using \eqref{eq:6}, the scaling factor of each iteration can be calculated as shown in table \ref{table_I}. This sequence is:-
\begin{enumerate}
\item Easy to design, because it requires a single parameter $S$.
\item Does not need to store a specific scaling sequence for each code rate. It only requires to store the optimal updating step size $S$ of each code rate.
\item Easy to implement, because it only requires shifting right by $\lceil i/S \rceil $ then subtraction. Number of required shifts $\lceil i/S \rceil $ can be stored in a register and increased by 1 every $S$ iterations.
\end{enumerate}
As shown in Fig. \ref{SVS_block}, the SVS scaling strategy is implemented by two sub-circuits. The first sub-circuit calculates the number of shifts required in each iteration $\lceil i/S \rceil $, this number of shifts is calculated by dividing the iteration clock by $S$ (updating step) to generate the updating clock of the scaling factor, then this updating clock is used to count the required number of shifts $\lceil i/S \rceil $ using a counter starts by 1. The other sub-circuit multiplies the minimum operation output $U'$ by $\alpha$ specified in \eqref{eq:6} by using a Barrel shifter to shift $U'$ right by $\lceil i/S \rceil $ then subtract.
\begin{figure}[!t]
\centering
\includegraphics[width=3.5in]{GSVS_diagram.png}
\caption{Circuit representation of GSVS scaling}
\label{GSVS_block}
\end{figure}
\begin{table}[!t]
\renewcommand{\arraystretch}{1.3}
\caption{Calculation of scaling factor value of each iteration using GSVS-min-sum algorithm}
\label{table_II}
\centering
\begin{tabular}{|c|c|}
\hline
Iteration index $i$ & $\alpha$ \\
\hline
$1 \to S$ & $\alpha_{0}$ \\
\hline
$(S+1) \to 2S$ & $0.5+ 0.5*\alpha_{0}$ \\
\hline
$(2S+1) \to 3S$ & $0.75+ 0.25*\alpha_{0}$\\
\hline
$(3S+1) \to 4S$ & $0.875+ 0.125*\alpha_{0}$ \\
\hline
\end{tabular}
\end{table}
\section{Generalized SVS-min-sum (GSVS-min-sum) algorithm}
As shown in table \ref{table_I}, SVS-min-sum sequence starts with 0.5 and increases exponentially with iterations. The limitation of starting with a fixed value of 0.5 restricts the performance to be sub-optimal. As a solution, we propose a new GSVS-min-sum algorithm where the scaling factors sequence is calculated by \eqref{eq:7}:
\begin{equation} \label{eq:7}
\alpha=1-(1-\alpha_{0})*2^{-(\lceil i/S \rceil-1)}
\end{equation}
Where $\alpha_{0}$ is the initial scaling factor. By using \eqref{eq:7}, scaling factor of each iteration is calculated as shown in table \ref{table_II}, where scaling factor values start with $\alpha_0$ and increase exponentially to unity for large value of iteration index $i$.\\
Circuit representation of GSVS scaling is shown in Fig. \ref{GSVS_block}. It is similar to SVS scaling circuit, but with two main differences: the first difference is that $U'$ is multiplied by $(1-\alpha_0)$ before shifting right, this is added to generalize the initial scaling factor. $(1-\alpha_0)$ is chosen to be simply implemented. We use $(1-\alpha_{0})$ to be in the form of $2^{-i}$ or $2^{-j}+2^{-k}$, where $i$, $j$ and $k$ are integer numbers; for example, if ${i=2 \to (1-\alpha_{0})=0.25}$, ${i=3 \to (1-\alpha_{0})=0.125}$ and if ${j=2,k=3 \to (1-\alpha_{0})=0.375}$ . The second difference is that the counter of required shifts starts with 0 instead of 1 because GSVS-min-sum requires $(\lceil i/S \rceil-1)$ shifts not $\lceil i/S \rceil$ as in SVS-min-sum.\\
In SVS-min-sum, we only need to optimize the updating step size $S$, however, in GSVS-min-sum we also need to optimize the initial scaling factor $\alpha_0$. To calculate the optimal $(\alpha_{0}, S)_{opt.}$, we use Nelder-Mead optimization Method \cite{nelder1965simplex} (summarized in \ref{sub:NM}) to minimize $(E_b/N_0)_{min}$, where $(E_b/N_0)_{min}$ is the minimum $E_b/N_0$ required to achieve pre-specified BER threshold. $(E_b/N_0)_{min}$ of given $(\alpha_{0}, S)$ is calculated by DE.
\subsection{Density Evolution (DE) of Min-Sum based algorithms} \label{sub:DE}
DE checks the ability of an LDPC decoder to correctly decode messages with specific noise variance. This is done by tracing the Probability Density Function (PDF) of messages passed between check and variable nodes (using all-zero code-word). Using of all-zero code-word is valid in Binary Phase Shift Keying (BPSK) because the LLR of both 1 and 0 has a similar PDF shape, but this is not valid in Quadrature Amplitude Modulation (QAM) signaling \cite{tullberg2005serial}.\\
DE is used in \cite{mackay1999good} to obtain the optimal weight distribution of irregular LDPC codes, and is used in \cite{chen2002density} \cite{zhang2005improved} to calculate the optimal scaling factor(s) of LDPC decoder. We use the same DE as in \cite{zhang2005improved} after changing the PDF of channel LLR so that we can use it with QAM signaling.
\subsubsection{Channel LLR's PDF of BPSK over an AWGN channel}
In BPSK, all-zero code-word's bits $V_{k}=0$ are modulated to $x_{k}=1-2V_{k}=1$. Then $x_{k}$ is transmitted over an Additive White Gaussian Noise (AWGN) channel with noise variance $\sigma^{2} $, so the received sequence is $y_{k}=1+n_{k}$ where $n_{k}$ is normally distributed random variable with mean=0 and variance=$\sigma^{2} $ . Therefore, the channel LLR $(U_{ch_{k}}=2 \times y_{k}/\sigma^{2})$ is normally distributed random variable with mean=$2/\sigma^{2}$ and variance=$4/\sigma^{2}$. The PDF of $U_{ch_{k}}$ is used as the initial PDF of variable nodes' messages to check nodes.
\subsubsection{Channel LLR's PDF of higher order QAM constellation over an AWGN channel
As shown in \cite{tullberg2005serial}, for higher order constellations, we cannot assume that all-zero code-word was transmitted. Therefore, we used a similar procedure to \cite{tullberg2005serial}, where authors modified the definition of bit's LLR to be: "\textit{LLR of receiving the same bit value as was transmitted}" instead of "\textit{LLR of receiving 0}". This is equivalent to replacing $U_{ch}$ by $U_{ch}^{+}$, where $U^{+}_{ch_k}=U_{ch_k}\times(1-2V_{k})$. So $U^{+}_{ch_k}$ will be positive if and only if $U_{ch_k}$ has the same sign as given by $V_{k}$ .\\
Firstly, for any constellation point $W$, we calculate the PDF of $U_{ch}$ for bit number $l$ into $W$ given that $W$ is transmitted $f_{U_{ch}}(u_{l}/W)$ \cite{benjillali2006probability}. Then we calculate the average PDF of $U_{ch}^{+}$ by \eqref{eq:8}.
\begin{align} \label{eq:8}
f_{U_{ch}^{+}}(u)=\frac{1}{\eta} \sum_{l=1}^{log_{2}(M)/2} \bigg( &\sum_{W \in Z_{l}} f_{U_{ch}}(u_{l}=u/W)\nonumber \\ + & \sum_{W \in O_{l}} f_{U_{ch}}(u_{l}=-u/W) \bigg)
\end{align}
Where $l$ is the bit position index. Only half of bit positions were used, because of symmetry between real and imaginary axes of QAM signaling. $Z_{l}$ is the set of $W$ where bit position $l$ contains 0. $O_{l}$ is the set of $W$ where bit position $l$ contains 1. $\eta$ is PDF correction factor used to ensure that area under the PDF=1.\\
In other words, calculate the average PDF of $U_{ch}$ for zeros and $-U_{ch}$ for ones. Then use this PDF as the initial PDF of variable nodes' messages.
\subsection{Nelder-Mead (NM) optimization method } \label{sub:NM}
NM optimization \cite{nelder1965simplex} of $(\alpha_{0}, S)$ is based on constructing a simplex (polygon) of 2+1=3 random solution points $\{X_i =(\alpha_0 , S)_i | i=1,2$ or $ 3\}$ for our 2-dimension problem. After calculating $(E_b/N_0)_{min}$ of these three points, the worst point is replaced by a better point as described in the flow chart in Fig.\ref{NM_block}. This procedure is repeated until the simplex shrink enough to the optimal solution.\\
We use Nelder-Mead method for many reasons: first, it does not need the mathematical derivative of the cost function (which is not available because our cost function is $(E_b/N_0)_{min}$ which comes from DE). Second, Nelder-Mead is faster in convergence than other heuristic methods. Finally, our cost function has only one minimum, so the algorithm does not get trapped in a local minimum. To prove the last claim that $(E_b/N_0)_{min}$ has only one minimum, we calculated it for all possible combinations of $(\alpha_0, S)$ for short length LDPC code with rate 0.5 specified in DVB-T2. Fig.\ref{single_min} shows that $(E_b/N_0)_{min}$ has only one minimum. Similar results are obtained for all tested codes. Optimization of $(\alpha_0, S)$ is calculated offline for each LDPC code rate, then the optimal value is used to implement the decoding circuit.
\begin{figure}[!t]
\centering
\includegraphics[width=3.5in]{NM_digram.png}
\caption{Flow chart of Nelder-Mead method}
\label{NM_block}
\end{figure}
\begin{figure}[!t]
\centering
\includegraphics[width=2.5in]{single_min.png}
\caption{contours of $(E_b/N_0)_{min}$ for short code with rate=0.5}
\label{single_min}
\end{figure}
\section{Simulation environment and results}
For simulation, we used (16200, 7200) eIRA LDPC code specified in DVB-T2 standard \cite{bluebooka122, bluebooka133}. Data are produced as binary bits modulated using the challenging 256-QAM modulation scheme and sent over an AWGN channel. The simulations are performed using MATLAB platform. Maximum number of iterations is set to 40 iterations.
Fig.\ref{WER_05_Short} shows the WER of LLR-SPA, SVS-min-sum with $S=10$ \cite{ahmed14CCNC}, Scaled Min-Sum with $\alpha= 15/16$ (optimized by DE and the same as in \cite{ahmed14CCNC}), GSVS-min-sum with $(\alpha_0=0.75$ and $S=9)$ (optimized by DE with Nelder-Mead method) and 2D correction min-sum; where the output of the check nodes with degree 4,5,6 and 7 is multiplied by 0.94, 0.92, 0.88 and 0.86 respectively, and the output of the variable nodes with degree 1,2,3 and 8 is multiplied by 1.00, 1.00, 0.91 and 0.83 respectively \cite{zhang2005improved}. Results in Fig.\ref{WER_05_Short} show that:
\begin{figure}[!t]
\centering
\includegraphics[width=3.5in]{WER_05_Short.png}
\caption{WER of (16200, 7200) LDPC over 256QAM with different decoding algorithms}
\label{WER_05_Short}
\end{figure}
\begin{itemize}
\item GSVS-min-sum has better performance than SVS-min-sum by 0.5 dB at WER= $10^{-3}$, this indicates the importance of the proposed algorithm, which jointly optimizes the initial scaling factor $\alpha_0$ with the updating step size $S$.
\item Although 2D correction min-sum has an excellent performance after many iterations (200 iterations) \cite{zhang2005improved}, it has higher WER than GSVS-min-sum algorithm after 40 iterations for the whole simulation range and higher than scaled-min-sum for high $E_b/N_0$ range. The gap between GSVS-min-sum and 2D correction is 0.3 dB at WER= $10^{-3}$.
\item There is a small gap between GSVS-min-sum and LLR-SPA performances (0.1 dB for low $E_b/N_0$ and nearly disappears at high $E_b/N_0$), with much lower implementation complexity.
\item For the scaled min-sum algorithm, Optimizing the scaling factor of each code rate increases its performance specially for high $E_b/N_0$. This concept is illustrated in \cite{ahmed14CCNC} by showing that each code rate of DVB-T2 LDPC codes has different optimal scaling factor.
\end{itemize}
Fig.\ref{Itr_05_Short} shows clearly that GSVS-min-sum not only has lower WER than other min-sum based algorithms, but also it has the lowest average number of iterations which leads to lower latency and higher average throughput.
\\For more results, we used three different rates of the LDPC codes specified in DVB-T2 standard with BPSK, these codes are (16200,7200) short code with nominal rate = 0.5, (16200,11880) short code with nominal rate = 0.75 and (64800,48600) normal code with rate = 0.75. WER of these codes with constant scaled-min-sum, SVS-min-sum and GSVS-min-sum decoding algorithm are shown in Fig. \ref{WER_BPSK}. Simulation results show that GSVS-min-sum has the lowest WER among the three decoding algorithms, even though scaled min-sum has lower WER than SVS-min-sum or not. Note that: constant scaling (in scaled min-sum) and SVS-min-sum are special cases of the GSVS-min-sum where $S$= number of iterations for scaled min-sum and $\alpha_0=0.5$ for SVS-min-sum. So GSVS-min-sum, which has optimized values for both $\alpha_0$ and $S$, has the best performance between them as shown in Fig. \ref{WER_BPSK}. The parameters of the three decoding algorithms are shown in table \ref{table_III}. For (16200,11880) code, GSVS-min-sum has the same performance as constant scaling factor and better performance than SVS-min-sum. The poor performance of SVS-min-sum comes from the limitation of starting by 0.5 which is away from the optimal scaling sequence. For the other codes, GSVS-min-sum has better performance than both SVS-min-sum and constant scaled min-sum.
\section{CONCLUSION AND FUTURE WORK}
In this paper, we generalized the SVS-min-sum decoder by allowing it to start with any initial scaling factor $\alpha_0$. Simulation results indicated the superior performance and lower latency of GSVS-min-sum decoder to other min-sum based algorithms. In addition, GSVS-min-sum algorithm performance is very close to the LLR-SPA with much lower complexity. Moreover, the proposed algorthim is still simpler to implement than both the variable scaling factor in \cite{lechner2006improved} and the 2D correction Min-Sum in \cite{zhang2005improved}. As future work, we will apply our GSVS-min-sum decoding algorithm to layered LDPC codes implementation.
\section*{Acknowledgment}
This work is supported by E-JUST and Minister of High Education (MoHE). And is supported by NTRA as part of the project "Design and implementation of DVB-T/T2 solution".
\begin{figure}[!t]
\centering
\includegraphics[width=3.5 in]{Itr_05_Short}
\caption{Average number of iterations of (16200, 7200) LDPC over 256QAM with different decoding algorithms}
\label{Itr_05_Short}
\end{figure}
\begin{table}[!t]
\renewcommand{\arraystretch}{1.3}
\caption{Optimal parameters of different LDPC decoding algorithms for BPSK}
\label{table_III}
\centering
\begin{tabular}{|c|c|c|c|}
\hline
& constant $\alpha$ & step of SVS & $(\alpha_0,S)$ of GSVS \\
\hline
$(16200,7200)$ & 0.9375 & 5 & $(0.75,9)$\\
\hline
$(16200,11880)$ & 0.875 & 10 & $(0.75,16)$ \\
\hline
$(64800,48600)$ & 0.875 & 10 & $(0.75,18)$\\
\hline
\end{tabular}
\end{table}
\begin{figure}[!t]
\centering
\includegraphics[width=3.5in]{BPSK_fig.png}
\caption{WER of different LDPC cods over BPSK}
\label{WER_BPSK}
\end{figure}
\bibliographystyle{IEEEtran}
|
\section{Introduction}
Consider the quantum harmonic oscillator in dimensionless variables, where the Hamiltonian, in terms of position and momentum operators $\hat{x},\hat{p},$ satisfying $[\hat{x},\hat{p}]=i,$ is given by the formula
\begin{equation} \hat{H}=\frac{1}{2}(\hat{p}^2+\hat{x}^2).\notag\end{equation}
Its normalized eigenstates, when represented as square integrable functions of the position variable $x$ are
\begin{equation} \psi_n(x)=\pi^{-1/4}(2^n n!)^{-1/2} H_n(x)\,e^{-x^2/2},\notag\end{equation}
$H_n$ are Hermite's polynomials: $H_n(x)=(-1)^ne^{x^2}\frac{d^n}{dx^n}\,e^{-x^2}.$ The corresponding probability densities $P_n(x)$ are then given by $|\psi_n(x)|^2,$ i.e.
\begin{equation} P_n(x)=a_n \left(H_n(x)\,e^{-x^2/2}\right)^2,\notag\end{equation}
where
\begin{equation} a_n=\frac{1}{\pi^{1/2}2^n n!},\label{eq:an}\end{equation}
so that the normalization condition holds:
\begin{equation} \int_{-\infty}^{\infty}P_n(x){\mathrm d} x=1.\notag\end{equation}
The functions $P_n(x)$ are symmetric with respect to the origin $x=0.$ The classical turning points are at $x=\pm\sqrt{2n+1},$ thus the probability of quantum tunneling into the classically forbidden region are given by the formula
\begin{equation} P_{n,\mathrm{tun}}=2\int_{\sqrt{2n+1}}^{\infty}P_n(x)\,{\mathrm d} x=2a_n Q_n,\label{eq:pn}\end{equation}
where
\begin{equation} Q_n=\int_{\sqrt{2n+1}}^{\infty}\left(H_n(x)\,e^{-x^2/2}\right)^2\,{\mathrm d} x.\label{eq:qn}\end{equation}
The probability distributions $P_n(x)$ are usually represented (see, for example, \cite[p. 168]{thaller}) as in Fig. \ref{fig:prn}
\begin{figure}[!htb]
\center
\includegraphics[width=0.9\textwidth]{p40.eps}
\caption{Probability distribution $P_{40}(x),$ with the classical turning points scaled to $x=\pm 1,$ and the curve $P_{n,\mathrm{tun}}$ for $n$ from $0$ to $40.$ The black $U$-shaped curve represents classical probability distribution $P_{\mathrm{class}}(x)=1/(\pi\sqrt{1-x^2}\,)$ for an oscillator with known energy and unknown phase. Shadowed tails beyond the classical turning points are responsible for non-zero quantum tunneling probabilities. }\label{fig:prn}
\end{figure}
The tunneling probabilities $P_{n,\mathrm{tun}}$ are rarely discussed in textbooks on quantum mechanics. Sometimes they are discussed in quantum chemistry textbooks. For instance in \cite[p. 92]{muller} the first three are calculated (though the second one with an error: it should be 0.1116 instead of 0.116). In \cite[p. 66-67; p. 98-99 in 2nd ed (2014)]{friedman} $P_{0,\mathrm{tun}}$ is calculated with a comment: ``\textit{The probability of being found in classically forbidden regions decreases quickly with increasing $n,$ and vanishes entirely as $n$ approaches infinity}.'' Yet the question of ``how quickly" is left open there. To answer this question we will derive the asymptotic behavior, first by using a heuristic method based on a simple formula by Szeg\"{o}, then, in a rigorous way, using another formula due to Olver. Both methods lead to the same final result. The problem that we solve is closely related to a more general problem of large quantum numbers and the correspondence principle. As can be seen from Refs. \cite{kiwi,gao,eltschka,boisseau}, even for such a simple quantum system as the harmonic oscillator, there are surprises (a more general discussion can be found in the monograph \cite{bolivar}). Our result showing a very slow approach to zero of $P_{n,\mathrm{tun}}$ for large $n$ may be considered as another surprise.
In section \ref{sec:szego} we derive the leading term formula using heuristic reasoning. In section \ref{sec:olver} we use mathematically rigorous arguments based on dominated convergence theorem, while controlling the order of terms. In section \ref{sec:2order} we derive the
second order correction to the leading term. The proof of the main lemma used in this paper is given in the Appendix. Our theoretical results are also verified numerically.
\section{Derivation of the asymptotic tunneling probability formula\label{sec:szego}}
We are interested in the approximate behavior of $P_{n,\mathrm{tun}},$ as a function of $n$, for ``large'' $n.$ To this end we first use the asymptotic formula by Szeg\"{o} \cite[p. 201, (8.22.14)]{szego} ( valid in the transition regions )
\begin{equation} e^{-\frac{x^2}{2}}H_n(x)=\frac{3^{\frac{1}{3}}2^{\frac{n}{2}+\frac{1}{4}}(n!)^{\frac{1}{2}}}{\pi^{\frac{3}{4}}n^{\frac{1}{12}}} \mathrm{A}(t)\left(1+O(n^{-\frac{2}{3}})\right),\label{eq:sza}\end{equation}
\begin{equation} x=(2n+1)^{1/2}-2^{-1/2}3^{-1/3}n^{-1/6}t.\notag\end{equation}
The function $\mathrm{A}(x)$ used by Szeg\"{o} relates to the Airy functions $\mathrm{Ai}(x)$ by the following formula (cf. \cite[p.194]{olver}):
\begin{equation} \mathrm{A}(t)=3^{-1/3}\pi\, \mathrm{Ai}(-3^{-1/3}t).\notag\end{equation}
Thus Eq. \ref{eq:sza} can be written as\footnote{Another use of the Airy function in a discussion of the harmonic oscillator wave functions can be found, for instance, in \cite[Fig. 5-23]{powell}, in the section on WKB approximation.}
\begin{equation} e^{-x^2/2}H_n(x)\approx \pi^{1/4}2^{n/2+1/4}(n!)^{1/2}n^{-1/12}\mathrm{Ai}(s),\label{eq:l}\end{equation}
where \begin{equation} s=2^{1/2}n^{1/6}(x-\sqrt{2n+1}).\notag\end{equation}
The above asymptotic formula is proven to be valid for $t$ bounded. In the calculation below I will use this result for all $s > 0,$ because the main contributions to integral come from a very small neighborhood of $s = 0,$ in particular when $n$ is large in (\ref{eq:l}).\\
Since $\frac{dx}{ds}=2^{-1/2}n^{-1/6},$ we can calculate now $Q_n$ defined in Eq. \ref{eq:qn} as
\begin{equation} Q_n=\int_{\sqrt{2n+1}}^{\infty}\left(H_n(x)\,e^{-x^2/2}\right)^2\,{\mathrm d} x.\label{eq:qn1}\end{equation}
We get
\begin{equation} Q_n\approx\pi^{1/2}2^n n! n^{-1/3}\int_0^{\infty}\mathrm{Ai}^2(s){\mathrm d} s.\notag\end{equation}
The integral of the squared Airy function is known (cf. e.g. \cite[Eq. (3.75), p. 50]{airy}):
\begin{equation} \int_0^{\infty}\mathrm{Ai}^2(s){\mathrm d} s=\frac{1}{3^{2/3}\,\Gamma^2[1/3]}.\notag\end{equation}
Therefore (cf. Eqs. (\ref{eq:pn}),(\ref{eq:an}))
\begin{align} P_{n,\mathrm{tun}}&=2a_nQ_n&\\
&\approx 2\frac{1}{\pi^{1/2}2^n n!}\pi^{1/2}2^n n! n^{-1/3}\frac{1}{3^{2/3}\,\Gamma^2[1/3]}&\notag\\
&=\frac{2}{3^{2/3}\Gamma^2[1/3]}n^{-1/3}.\label{eq:pns}\end{align}
Thus, numerically,
\begin{equation} P_{n,\mathrm{tun}}\approx \frac{0.133975}{n^{1/3}}.\label{eq:fita}\end{equation}
\begin{figure}[!htb]
\center
\includegraphics[width=0.9\textwidth]{plot1.eps}
\caption{Comparison of the tunneling probabilities calculated from the exact formula with the curve given by Eq.
(\ref{eq:fita}), starting from $n=5.$} \label{fig:fit1}
\end{figure}
As it can be seen from Fig. \ref{fig:fit1}, the asymptotic formula (\ref{eq:fita}) seems to agree with the numerical calculations for $n$ up to $612.$ Yet in its derivation we have assumed the validity of Szeg\"{o}'s asymptotic formula in the unbounded interval. Can such an extension be mathematically justified? We will address this question in the next section.
\section{Using Olver's formula\label{sec:olver}}
Olver \cite[p. 403]{olver97} gives the following formula valid for $x\ge 1$:
\begin{equation} e^{-\frac{\nu^2x^2}{2}}H_n(\nu x)=c_n\left(\frac{\zeta}{x^2-1}\right)^{\frac{1}{4}}\left(\mathrm{Ai}(\nu^{\frac{4}{3}}\zeta)+\epsilon(x)\right),\label{eq:olver}\end{equation}
where
\begin{equation}\nu=\sqrt{2n+1},\quad c_n=(2\pi)^{\frac{1}{2}}e^{-\frac{\nu^2}{4}}\nu^{\frac{(3\nu^2-1)}{6}},\end{equation}
\begin{equation} \zeta=\left(\frac{3}{4}x(x^2-1)^{1/2}-\frac{3}{4}\cosh^{-1}(x)\right)^{2/3},\label{eq:zeta}\end{equation}
and
\begin{equation} |\epsilon(x)|\le 1.36\left(\exp(0.09\nu^{-2})-1\right)\mathrm{Ai}(\nu^{4/3}\zeta).\label{eq:eps}\end{equation}
\begin{remark}: The function $\cosh^{-1}(x)$ in the above formula denotes the positive branch of the inverse function of $\cosh.$ For general aspects of asymptotic methods for integrals see \cite{temme}, where, in particular, Section 23.4 provides a complete description of the Airy-type expansion of Hermite polynomials. Notice also that $\nu^{-2}=1/(2n+1) = (1/(2n)) (1-1/(2n))+O(n^{-3}).$
\end{remark}
We will need Eq. (\ref{eq:olver}) squared.
Using Eqs. (\ref{eq:olver})-(\ref{eq:eps}) we write it in the form
\begin{equation} \left(e^{-\frac{\nu^2x^2}{2}}H_n(\nu x)\right)^2= d_n \left(1+O(\frac{1}{n})\right) \left(\frac{\zeta}{x^2-1}\right)^{1/2}\mathrm{Ai}^2(\nu^{4/3}\zeta),\notag\end{equation}
where
\begin{equation} d_n=2\pi e^{-\nu^2/2}\nu^{(3\nu^2-1)/3}.\notag\end{equation}
In order to use this formula for calculating $P_{n,\mathrm{tun}},$ we use Eq. (\ref{eq:qn}) changing the integration variable to get
\begin{align} P_{n,\mathrm{tun}}&=2a_n\nu\int_1^{\infty}\left(H_n(\nu x)\,e^{-\nu^2 x^2/2}\right)^2 {\mathrm d} x&\notag\\
&= 4 a_n\pi \nu e^{-\nu^2/2}\nu^{(3\nu^2-1)/3} (1+O(\frac{1}{n}))I_n,\notag\end{align}
where
\begin{equation} I_n=\int_1^\infty \left(\frac{\zeta}{x^2-1}\right)^{1/2}\mathrm{Ai}^2(\nu^{4/3}\zeta) {\mathrm d} x.\notag\end{equation}
An easy power expansion leads to
\begin{equation} 4 a_n\pi \nu e^{-\nu^2/2}\nu^{(3\nu^2-1)/3} =n^{1/3}\left(2^{7/3}+ O(\frac{1}{n})\right),\notag\end{equation}
therefore
\begin{equation} P_{n,\mathrm{tun}}= 2^{7/3}n^{1/3}\left(1+O(\frac{1}{n})\right)I_n.\notag\end{equation}
\begin{figure}[!htb]
\center
\includegraphics[width=0.9\textwidth]{zeta.eps}
\caption{Plot of the function $\zeta(x)$ for $x>1.$}\label{fig:zet}
\end{figure}
It is convenient now to introduce a new integration variable \begin{equation} t=\nu^{4/3}\zeta.\label{eq:t}\end{equation} Fig.~\ref{fig:zet} shows the dependence of $\zeta$ on $x.$ We find
\begin{equation} \frac{{\mathrm d}\zeta}{{\mathrm d} x}=\left(\frac{x^2-1}{\zeta}\right)^{1/2},\notag\end{equation}
therefore
\begin{equation} {\mathrm d} x=\nu^{-4/3}\left(\frac{\zeta}{x^2-1}\right)^{1/2}\,{\mathrm d} t, \notag\end{equation}
and thus
\begin{equation} I_n=\nu^{-4/3}\int_0^\infty f_n(t)\mathrm{Ai}^2(t){\mathrm d} t,\notag\end{equation}
where
\begin{equation} f_n(t)=\frac{\zeta}{x^2-1}\notag\end{equation} is implicitly defined through $\zeta=\nu^{-\frac{4}{3}}t$ and Eq. (\ref{eq:zeta}) that allows us to find $x$ as a function of $\zeta.$
Since \begin{equation}
\nu^{-4/3}=(2n+1)^{-2/3}= 2^{-2/3}n^{-2/3}\left(1+O(\frac{1}{n})\right),\notag\end{equation} for $P_{n,\mathrm{tun}}$ we get
\begin{equation} P_{n,\mathrm{tun}}= 2^{5/3}n^{-1/3}\left(1+O(\frac{1}{n})\right)\int_0^\infty f_n(t)\mathrm{Ai}^2(t){\mathrm d} t.\label{eq:pnf}\end{equation}
Denote by $F_n$ the integral in Eq. (\ref{eq:pnf}):
\begin{equation} F_n=\int_0^\infty f_n(t)\mathrm{Ai}^2(t){\mathrm d} t.\label{eq:Fn}\end{equation}
We will show now that \begin{equation} F_n\sim F_\infty=2^{-2/3}\int_0^\infty \mathrm{Ai}^2(t){\mathrm d} t=\frac{1}{2^{2/3}3^{2/3}\Gamma(\frac{1}{3})^2},\label{eq:fnlim}\end{equation}
and therefore
\begin{equation} P_{n,\mathrm{tun}}\sim\frac{2}{3^{2/3}\Gamma(\frac{1}{3})^2}\,n^{-1/3},\notag\end{equation}
in agreement with Eq. (\ref{eq:pns}).
\begin{figure}[!htb]
\center
\includegraphics[width=0.9\textwidth]{rplot.eps}
\caption{Numerically calculated ratios $F_\infty/F_n,$ $6\leq n\leq 500.$}\label{fig:r}
\end{figure}
Fig. \ref{fig:r} shows numerically calculated ratios $F_\infty/F_n$ for $6\leq n\leq 500.$
In order to verify Eq. (\ref{eq:fnlim}), we notice that (see the Appendix) the function \begin{equation} f(x)=\frac{\left(\frac{3}{4}x(x^2-1)^{1/2}-\frac{3}{4}\cosh^{-1}(x)\right)^{2/3}}{x^2-1},\quad (x>1)\label{eq:s11}\end{equation}
is monotonically decreasing, from $f(1)=\lim_{x\rightarrow 1+}\,f(x)=2^{-2/3}$ to $0$ at $x\rightarrow \infty.$
It follows that the sequence $f_n(t)$ is uniformly bounded. Moreover, with fixed $t>0,$ and since $\zeta=(2n+1)^{-1/3}t,$ when $n\rightarrow\infty,$ $\zeta\rightarrow 0,$ thus $x\rightarrow 1,$ therefore $\lim_{n\rightarrow\infty}f_n(t)= f(1)=2^{-2/3}.$ Thus \begin{equation} \lim_{n\rightarrow\infty}\,F_n=\int_0^\infty \left(\lim_{n\rightarrow\infty}f_n(t)\right)\mathrm{Ai}^2(t){\mathrm d} t.\notag\end{equation}
by the dominated convergence theorem.
\section{Next order approximation\label{sec:2order}}
Using the method described in the previous section one can easily get the next term in the asymptotic formula for $F_n.$ The result is
\begin{equation} F_n=2^{-2/3}\int_0^\infty \mathrm{Ai}^2(t){\mathrm d} t-\frac{2^{-2/3}n^{-2/3}}{5}\int_0^\infty t\mathrm{Ai}^2(t){\mathrm d} t+O(\frac{1}{n}),\label{eq:fn2}\end{equation}
which, after some simplifications leads to the formula
\begin{equation} P_{n,\mathrm{tun}}= \frac{1}{n^{1/3}}\left(0.133975-\frac{0.0122518}{n^{2/3}}\right)+O(n^{-4/3}).\label{eq:2o}\end{equation}
The second term in Eq. (\ref{eq:fn2}) can be derived from the second term in the Taylor series expansion $f_n(t)=f_n(0)+(df_n/dt)|_{t=0}\,t +O(t^2),$ if we use Eq. (\ref{eq:t}) and notice that
\begin{equation} (df_n/dt)|_{t=0}=\nu^{-4/3}(df_n/d\zeta)|_{\zeta=0}\sim 2^{-2/3}n^{-2/3}(-1/5).\notag\end{equation}
Fig. \ref{fig:fit2} shows the comparison of the exact data with the two curves. The black curve, obtained from Eq. (\ref{eq:2o}) is closer to the data than the grey one, obtained from the simple approximation given by Eq. (\ref{eq:pns}).
\begin{figure}[!htb]
\center
\includegraphics[width=0.9\textwidth]{plot2.eps}
\caption{Details of the comparison for $n>512:$ exact data (dots), simple formula $P_{n,\mathrm{tun}}= 0.133975\, n^{-1/3}$ (upper curve), and the formula with the first correction term $P_{n,\mathrm{tun}}= 0.133975\,n^{-1/3}-0.0122518\, n^{-1}$ (lower curve).} \label{fig:fit2}
\end{figure}
Additional correction terms have been obtained by Paris in \cite{rp} by different means.
\section{Appendix}
Fig. \ref{fig:zeta2} shows the computer generated plot of the function $\zeta(x)/(x^2-1)$ for $x>1.$ While the behavior of this function can be easily guessed from the plot, it is not that easy to prove rigorously that the function has the properties that are necessary for the application of the dominated convergence theorem used in this paper.
\begin{figure}[!htb]
\center
\includegraphics[width=0.9\textwidth]{zeta2.eps}
\caption{Plot of the function $\zeta(x)/(x^2-1).$} \label{fig:zeta2}
\end{figure}
\begin{lemma}
Let $f(x)$ be as in Eq. (\ref{eq:s11}).
Then $f$ is monotonically decreasing, from $f(1)=\lim_{x\rightarrow 1+}\,f(x)=2^{-2/3}$ to $0$ at $x\rightarrow \infty.$
\label{lem:1}\end{lemma}
\begin{proof}
We have
\begin{equation} f(x)=\left(\frac{3}{4}\tilde{f}(x)\right)^{\frac{2}{3}},\label{eq:s2}\end{equation}
where
\begin{equation} \tilde{f}(x)=\frac{x(x^2-1)^{1/2}-\cosh^{-1}(x)}{(x^2-1)^{3/2}},\notag\end{equation}
and it is enough to show that the statement in the lemma holds for $\tilde{f}.$ Using the identity
\begin{equation} \cosh^{-1}(x)=\log (x+\sqrt{x^2-1}),\notag\end{equation}
and introducing $s,\, 0<s<\pi/2,$ through $x =\sec s,$ we can write $\tilde{f}(x)$ as a function $z(s)$ of $s$
\begin{equation} z(s)= \frac{\sec s\, \tan\, s-\log(\sec s+\tan s)}{\tan^3 s}=\frac{a(s)}{b(s)},\label{eq:s3}\end{equation}
and to prove the lemma we need to prove that $z$ is monotonically decreasing from $z(0)=\lim_{s\rightarrow 0+} z(s).$ Let \begin{equation} h(u)=\sec u \tan^2 u.\notag\end{equation}
By differentiation of the formulas below it is easy to verify that
\begin{equation}
a(s)= 2 \int_0^s h(u){\mathrm d} u,\quad b(s)= 3 \int_0^s h(u)\,\sec u\,{\mathrm d} u.
\notag\end{equation}
We notice that $a(0)=b(0),$ and that, by de l'H\^{o}spital's rule $$ z(0)=\lim_{s\rightarrow 0+}\frac{a'(s)}{b'(s)}=\frac{2}{3\sec 0}=2/3.$$
Therefore, using Eqs. (\ref{eq:s11}),(\ref{eq:s2}), $f(1)=2^{-2/3}.$
Taking the derivative of Eq. (\ref{eq:s3}) we get
\begin{equation} z'(s)=\frac{a'(s)b(s)-a(s)b'(s)}{b^2(s)}.\notag\end{equation} The denominator is positive for $s>0.$ The numerator is negative owing to the fact that, for $s>0,$ \begin{align}
a'(s)b(s)-a(s)b'(s)&=6\left(h(s)\int_0^s h(u)\sec u\,{\mathrm d} u-h(s)\sec s\int_0^s h(u){\mathrm d} u\right)&\notag\\
&=6h(s)\int_0^s h(u)(\sec u - \sec s){\mathrm d} u,
\notag\end{align}
with $h(s)>0$, and $\sec(u)<\sec s$ for $u<s.$ Therefore $z'(s)<0.$
\end{proof}
\section*{Acknowledgments}
The author would like to thank Ilia Krasikov, Richard Paris, Philippe Sosoe, Jerzy Szulga, and Nico Temme for their encouragement, helpful insights, comments and criticism. In particular Nico Temme has suggested using Olver's formula
and Richard Paris assisted me throughout the evolution of this paper. Thanks are also due to Ronald Friedman for his interest in the problem, as well as to the anonymous referee for helpful comments.
|
\section{Introduction}
\subsection{Setup}\label{Sec.Intro.Setup}
Let $R$ be a (commutative, associative) local ring over some base field $\k$ of zero characteristic.
(The simplest examples are $\k[[\ux]]$, $\C\{\ux\}$, and $C^\infty(\R^p,0)$.)
Denote by $\cm\sset R$ the maximal ideal and by
$\Mat$ the $R$-module of $m\times n$ matrices with entries in $R$. We always assume $m\le n$, otherwise one
can transpose the matrix.
\
Various groups act on $\Mat$, we consider mostly the following actions.\vspace{-0.1cm}
\bex\label{Ex.Intro.Typical.Groups}
\bee[i.]
\item
The left multiplications $G_l:=GL(m,R)$, the right and the two-sided multiplications $G_r:=GL(n,R)$, $G_{lr}:=G_l\times G_r$. Matrices
considered up to $G_r$-transformations correspond to the embedded modules, $Im(A)\sset R^{\oplus m}$. Matrices
considered up to $G_{lr}$-transformations correspond to the non-embedded modules, $coker(A)=\quotients{R^{\oplus m}}{Im(A)}$.
\item The congruence, $G_{congr}=GL(m,R)\circlearrowright Mat_{m\times m}(R)$, acts by $A\to UAU^t$. Matrices considered up
to the congruence correspond to the bilinear/symmetric/skew-symmetric forms. The conjugation, $G_{conj}=GL(m,R)\circlearrowright Mat_{m\times m}(R)$,
acts by $A\to UAU^{-1}$. This action is important in representation theory.
\item Fix some ordered compositions, $m=\suml^k_{i=1}m_i$, $n=\suml^k_{j=1}n_j$. Take the corresponding block-structure on the matrix, i.e. $A=\{A_{ij}\}_{1\le i,j\le k}$, $A_{ij}\in Mat_{m_i\times n_j}(R)$.
Denote by $Mat^{up}_{\{m_i\}\times\{n_j\}}(R)$ the set of the upper-block-triangular matrices, i.e. $A_{ij}=\zero$ for $i>j$,
this is a free direct summand of $\Mat$.
Accordingly we consider the groups
\beq
G^{up}_l:=G_l\cap Mat^{up}_{\{m_i\}\times\{m_j\}}(R),\quad
G^{up}_r:=G_r\cap Mat^{up}_{\{n_i\}\times\{n_j\}}(R)
\ \text{ and } \ G^{up}_{lr}:=G^{up}_l\times G^{up}_r.
\eeq
The upper-block-triangular matrices considered up to the $G^{up}_r$-equivalence correspond to the submodules of $R^{\oplus m}$
filtered by chains of direct summands. The upper-block-triangular matrices considered up to the $G^{up}_{lr}$-equivalence correspond to filtered
homomorphisms of filtered free modules. (See \S \ref{Sec.Proofs.Upper.Triangular} for more detail.)
\eee
\eex
\
We study deformations of matrices. In applications one often deforms a matrix not inside the whole $\Mat$ but only inside a ``deformation subspace" (a subset of prescribed deformations), $A\rightsquigarrow A+B$, $A+B\in\Si\sseteq\Mat$.
\
In this paper the subset $\Si-\{A\}\sseteq\Mat$ is a submodule. Besides the trivial choice $\Si=\Mat$, we mostly work with the following submodules.
\vspace{-0.1cm}
\bex\label{Ex.Intro.Typical.Def.Spaces}
\bee[i.]
\item The congruence of (skew-)symmetric matrices, $A\stackrel{G_{congr}}{\sim}UAU^T$, $U\in GL(m,R)$, preserves the (skew-)symmetry.
Thus it is natural to deform the matrix by the (skew-)symmetric matrices only, i.e. $\Si=Mat^{sym}_{m\times m}(R)$ or $\Si=Mat^{skew-sym}_{m\times m}(R)$.
\item Similarly, it is natural to deform an upper-block-triangular matrix only inside $Mat^{up}_{\{m_i\}\times \{n_j\}}(R)$.
\item Sometimes one deforms only by ``higher order terms", e.g. $\Si=\{A\}+Mat_{m\times m}(J)$ or $\Si=\{A\}+Mat^{sym}_{m\times m}(J)$, for some ideal $J\sset R$.
\eee
\eex
\
Recall that over a local ring any matrix is $G_{lr}$-equivalent to a block-diagonal, $A\stackrel{G_{lr}}{\sim}\one\oplus \tA$,
where all the entries of $\tA$ lie in the maximal ideal $\cm$, i.e. vanish at the origin of $Spec(R)$. Similar statements hold for (skew-)symmetric matrices \wrt $G_{congr}$, see proposition \ref{Thm.Background.Chip.off.Unit}.
This splitting is natural in various senses and is standard in commutative algebra, singularity theory and other fields.
Often it is $\tA$ that carries all the essential information.
Therefore we often assume $A|_0=\zero$, i.e. $A\in\Matm$.
\subsection{The tangent spaces}
Fix an action $G\circlearrowright\Mat$, a deformation space $\Si\sseteq\Mat$, and a matrix $A\in\Si$. We assume that the
orbit $GA$ and the deformation space possess well defined tangent spaces at $A$, which are $R$-modules.
The general definitions/conditions are in \cite{Belitski-Kerner}, for the examples relevant to our paper see \S\ref{Sec.Background.Tangent.Spaces}.
The standard approach of deformation theory is to establish the existence of the miniversal (semi-universal) deformation and, when the later exists, to understand/to compute its tangent cone.
Accordingly one passes from the study of the germs $(GA,A)\sseteq(\Si,A)$, to the study of the tangent spaces, $T_{(GA,A)}\sseteq T_{(\Si,A)}$.
Much of the information about the deformation problem is encoded in the quotient module
\beq
T^1_{(\Si,G,A)}:=\quotient{T_{(\Si,A)}}{T_{(GA,A)}}.
\eeq
This $R$-module is the tangent space to the miniversal deformation, when the later exists and is smooth. It often controls the deformation theory of modules, (skew-)symmetric forms and other objects.
In Singularity Theory such a module is known as the Tjurina algebra for the contact equivalence, and the Milnor algebra for the right equivalence.
\
This module $T^1_{(\Si,G,A)}$ is the main object of our study.
\subsection{(In)finite determinacy}
The $(\Si,G)$-order of determinacy of $A$ is the minimal number $k\le\infty$ satisfying: if $A,A_1\in\Si$ and $jet_k(A)=jet_k(A_1)$ then $A_1\in GA$. Here $jet_k$ is the projection $\Mat\stackrel{jet_k}{\to}Mat_{m\times n}(\quotients{R}{\cm^{k+1}})$. More precisely:
\bed\label{Def.Intro.Determinacy.Classical}
$ord^\Si_G(A):=min\Big\{k|\ \Si\cap\big(\{A\}+Mat_{m\times n}(\cm^{k+1})\big)\sseteq GA\Big\}\le\infty.$
\eed
(We assume that the minimum is taken here over a non-empty set.)
\
If $ord^\Si_G(A)<\infty$ then $A$ is called {\em finitely-$(\Si,G)$-determined}.
In words: $A$ is determined (up to $G$-equivalence) by a finite number of terms of its Taylor expansion at the origin.
An immediate consequence is the {\em algebraization}: $A$ is $G$-equivalent to a matrix of polynomials.
Moreover, the order of determinacy gives an upper bound on the degrees of polynomials. We show that the finite determinacy is equivalent to
the stable algebraizability (a strengthening of the ordinary algebraizability), see \S\ref{Sec.Remarks.Algebraization} and \S\ref{Sec.Background.Fin.Determin.vs.Stable.Algebraizability}.
If $ord^\Si_G(A)=\infty$ the matrix is called {\em infinitely determined}.
This condition is empty when $\cm^\infty:=\capl_{k>0}\cm^k=\{0\}$.
But for some rings $\cm^\infty\neq\{0\}\sset R$, then even the infinite determinacy is non-trivial property. (The main example is the ring of germs of smooth functions, $R=C^{\infty}(\R^p,0)$, and its sub-quotients.) In this case $A$ is determined (up to $G$-equivalence) by its image under the $\cm$-adic completion, $\hA\in Mat_{m\times n}(\hR)$, i.e. its full Taylor expansion at the origin.
\
The (in)finite determinacy is the fundamental notion of Singularity Theory. For Algebraic Geometry/Commutative Algebra the finite determinacy
means (roughly) ``the deformation theory is essentially finite dimensional''.
More generally, as the determinacy expresses the ``minimal stability", it is important in any area dealing with matrices over rings
(or matrix families or matrices depending on parameters).
See \S\ref{Sec.Results.Remarks} for further motivation and \S\ref{Sec.Results.Relation.to.Singularity.Theory} for a brief overview and the
relation of our work to the known results.
\
In \cite{Belitski-Kerner} we have reduced the study of determinacy to the understanding of the support/annihilator of the module
$T^1_{(G,\Si,A)}$. The matrix $A$ is (in)finitely determined iff $\cm^N T_{(\Si,A)}\sseteq T_{(GA,A)}$ for some $N\le\infty$, alternatively
$\cm^N T^1_{(G,\Si,A)}=\{0\}$ or $\cm^N\sseteq ann(T^1_{(G,\Si,A)})$. The order of determinacy is
fixed by the annihilator $ann(T^1_{(G,\Si,A)})$, see \S\ref{Sec.Background.BK1} for more detail.
\subsection{Contents of the paper}
The results of \cite{Belitski-Kerner} reduce the study of (in)finite determinacy to the computation of the support of $T^1_{(\Si,G,A)}$, i.e.
the annihilator ideal $ann(T^1_{(\Si,G,A)})$.
In the current paper we study the module $T^1_{(\Si,G,A)}$ and its support for the $R$-linear actions $G\circlearrowright \Mat$ of example
\ref{Ex.Intro.Typical.Groups}.
\
One could begin from a completely general (set-theoretic) action $G\circlearrowright \Mat$.
However, if an action $G\circlearrowright\Mat$ is $R$-linear (though does not necessarily preserve the row/column structure)
and sends the degenerate matrices to degenerate (quite a reasonable assumption!) then $G\sseteq G_{lr}$. This property is proved in
\S\ref{Sec.Background.Largest.Reasonable.Group}, it belongs to the area of ``preserver problems on matrices", exhaustively studied for matrices over
fields, but less known for matrices over rings. Therefore example
\ref{Ex.Intro.Typical.Groups} contains the main ``reasonable" cases.
\
We compute (or at least bound) the annihilator $ann(T^1_{(\Si,G,A)})$, the main results are stated in \S\ref{Sec.Results}.
This gives
the ready-to-use criteria to determine (or at least to bound) $ord^\Si_G(A)$ for various group-actions.
We observe the standard dichotomy: for a given data $(m,n,dim(R),G\circlearrowright\Si\sseteq\Mat)$ either there are no finitely determined matrices in $\Matm\cap\Si$, or the generic finite determinacy holds.
The proofs of these results, corollaries/examples and further developments are in \S\ref{Sec.Proofs}.
In many cases we strengthen (generalize and quantify) both the classical and the relatively recent results.
\
In \S\ref{Sec.Background} we collect the needed background material: determinantal ideals and Pfaffian ideals, annihilator-of-cokernel,
the integral closure of ideals and modules, the matrix preservers, tangent spaces to the group orbits, approximation properties and the
main result of \cite{Belitski-Kerner}.
\subsection{Acknowledgements} Many thanks to R.-O. Buchweitz, V. Grandjean, G.-M. Greuel, V. Kodiyalam and M. Leyenson for
helpful/informative discussions and to L. Moln\'{a}r, V.V. Sergeichuk
for the much needed references on the ``matrix preservers" problems.
\section{The main results}\label{Sec.Results}
Below we state the main theorems and corollaries. Their proofs and examples are in \S\ref{Sec.Proofs}.
\subsection{Notations and Conventions}
Denote by $I_j(A)$ the determinantal ideal of all the $j\times j$ minors of $A$.
Denote by $ann.coker(A)$ the annihilator-of-cokernel ideal of the homomorphism of free modules $R^{\oplus n}\stackrel{A}{\to}R^{\oplus m}$. The properties/relation/computability of these ideals are discussed in \S\ref{Sec.Background.Fitting.Ideals.ann.coker}.
We often use the ideal quotient, $I:J=\{f\in R|\ fJ\sseteq I\}$, and the integral closure of ideals $I\sseteq\overline{I}$, see \S\ref{Sec.Background.Integral.Closure.Ideals}.
Sometimes we take the $\cm$-adic completion, $R\to\hR$, denote by $\hA$ the image of $A$ under the completion map.
\
Suppose $J\supseteq\cm^\infty$. The Loewy length, $ll_R(J)\le\infty$, is the minimal number $N\le\infty$
that satisfies: $J\supseteq\cm^N$. This number also equals the degree of the socle of the quotient module $\quotients{R}{J}$. For $R=\k[[\ux]]$ we have yet another expression, via the Castelnuovo-Mumford regularity,
$ll_R(J)=reg(\quotients{R}{J})+1$, see \cite[exercise 20.18]{Eisenbud}.
\
For Artinian rings the finite determinacy is not interesting, e.g. if $\cm^{N+1}=\{0\}$ then any matrix over $R$ is trivially $N$-determined. Therefore
in this paper we usually assume that $R$ is not Artinian, i.e. $\cm^N\neq\{0\}$ for any $N<\infty$. Geometrically this means: $dim(Spec(R))>0$.
\
In some places we assume that the ring $R$ is unique factorization domain (UFD), then we state this explicitly.
\subsubsection{Generic properties}\label{Sec.Intro.Genericity.Tougeron}
In this paper the {\em genericity} is always used in the following sense of \cite{Tougeron1968}. Fix some ``parameter space" $M$,
typically an $R$-module, suppose the $jet_d$-projections $\{R\stackrel{jet_d}{\to}\quotients{R}{\cm^{d+1}}\}_{d\in\N}$ induce
the projections $M\stackrel{jet_d}{\to}jet_d(M)$. We assume that all $jet_d(M)$ are algebraic varieties over $\k$.
A property $\cP$ is said to hold generically in $M$ if it holds for all points of $M$ lying in the complement to a
subset $X\sset M$ which is ``of infinite codimension".
Namely, for any $d$ the projection of $X$ lies inside some algebraic subset, $jet_d(X)\sseteq Y_d\sset jet_d(M)$, such that $codim_{jet_d(M)}(Y_d)\to\infty$.
\
\bed\label{Def.Generic.Finite.Determinacy}
We say that the generic finite determinacy holds for a given action $G\circlearrowright\Si$ if for any $A\in\Si$, any number $N<\infty$ and
the generic matrix $B\in Mat_{m\times n}(\cm^N)$, such that $A+B\in\Si$, the matrix $A+B$ is finitely determined.
\eed
Equivalently: the $\cm$-adic closure of the set of finitely-determined matrices is the whole space $\Si$.
In terms of singularity theory the finite determinacy means the finiteness of the Tjurina number, for a given $(\Si,G,A)$.
Thus the genericity of finite determinacy means:
the stratum of matrices whose Tjurina number is infinite, $\Si^G_{\tau=\infty}\sset \Si$, is itself of infinite codimension.
\subsection{The criteria for the $G_r$, $G_l$, $G_{lr}$ actions of example \ref{Ex.Intro.Typical.Groups}}\label{Sec.Results.Criteria.Glr}
In this subsection $\Si=\Mat$, with $m\le n$.
\bthe\label{Thm.Results.Annihilators.for.GlGrGlr} Let $A\in\Si=\Mat$, with $m\le n$.
\bee
\item $T^1_{(\Si,G_r,A)}=\Big(coker(A)\Big)^{\oplus n}$ and $ann(T^1_{(\Si,G_r,A)})=ann.coker(A)$.
\item If $m<n$ then $ann\big(T^1_{(\Si,G_l,A)}\big)=\{0\}$. If $m=n$ then $ann\big(T^1_{(\Si,G_l,A)}\big)=ann.coker(A)$.
\item $ann\big(T^1_{(\Si,G_{lr},A)}\big)\supseteq ann.coker(A)$.
If $R$ is Noetherian then $\overline{ann\big(T^1_{(\Si,G_{lr},A)}\big)}\sseteq \overline{I_m(A)}:\overline{I_{m-1}(A)}$.
\eee
\ethe
In particular, when $R$ is UFD and matrices are square ($m=n$), we get:
$ann\big(T^1_{(\Si,G_{lr},A)}\big)=ann.coker(A)$, see equation \eqref{Eq.Background.ann.coker.in.terms.of.Fitting.ideals}.
For $m<n$ the bounds are often close. In fact, in \S\ref{Sec.Proofs.Glr} we prove a slightly stronger statement:
\beq
\overline{ann\big(T^1_{(\Si,G_{lr},A)}\big)\cdot I_{m-1}(A)}\sseteq \overline{ I_m(A)}.
\eeq
\
Theorem \ref{Thm.Results.Annihilators.for.GlGrGlr} bounds $ann\big(T^1_{(\Si,G_{lr},A)}\big)$ in terms of
$ann.coker(A)$ and $\overline{I_m(A)}:\overline{I_{m-1}(A)}$. These ideals are usually rather small, thus finite-$G_{lr}$-determinacy places severe restrictions on the ring $R$.
For the applications to the (in)finite determinacy we assume:
\beq\label{Eq.Intro.assumption.R.Noetherian.or.surjects}
\text{either $R$ is Noetherian or the $\cm$-adic completion map is surjective, $R\twoheadrightarrow\hR$, and $\hR$ is Noetherian.}
\eeq
(The ring $C^\infty(\R^p,0)$ satisfies this condition, see \S\ref{Sec.Background.Approximation.Properties}.)
\bprop\label{Thm.Results.Fin.Det.for.GlGrGlr} Suppose $R$ satisfies condition \eqref{Eq.Intro.assumption.R.Noetherian.or.surjects}. Fix $\Si=\Mat$.
\bee
\item
If $ann.coker(A)\supseteq\cm^\infty$ then $ll_R\Big(ann.coker(A)\Big)-1 \le ord^\Si_{G_r}(A)\le ll_R\Big(ann.coker(A|_{\cm R^{\oplus n}})\Big)-1.$
In particular:
{\bf i.} If $dim(R)>(n-m+1)$ then no matrix in $\Matm$ is finitely-$G_r$-determined.
{\bf ii.} If $dim(R)\le(n-m+1)$ then the generic $G_r$-finite determinacy holds.
\item If $m<n$ then no matrix $A\in\Matm$ is finitely-$G_l$-determined.
If $m=n$ then $ord^\Si_{G_r}(A)=ord^\Si_{G_l}(A)$.
\item \
{\bf i.} If $dim(R)>(n-m+1)$ then no matrix in $\Matm$ is finitely-$G_{lr}$-determined,
{\bf ii.} If $R$ is Noetherian then
$ll_R\Big(\overline{I_m(A)}:\overline{I_{m-1}(A)}\Big)-1\le ord^\Si_{G_{lr}}(A)\le ll_R\Big(ann.coker(A|_{\cm R^{\oplus n}})\Big)-1.$
{\bf iii.} If $R$ satisfies \eqref{Eq.Intro.assumption.R.Noetherian.or.surjects} and $ann.coker(A)\supseteq\cm^\infty$ then
\[ll_\hR\Big(\overline{I_m(\hA)}:\overline{I_{m-1}(\hA)}\Big)-1\le ord^\Si_{G_{lr}}(A)\le ll_R\Big(ann.coker(A|_{\cm R^{\oplus n}})\Big)-1.\]
\eee
\eprop
Here $ann.coker(A|_{\cm R^{\oplus n}})$ denotes the annihilator-of-cokernel of the restricted map $\cm\cdot R^{\oplus n}\stackrel{A}{\to}R^{\oplus m}$.
Note that the bounds are rather tight:
$ll_R\big(ann.coker(A)\big)\le ll_R\big(ann.coker(A|_{\cm R^{\oplus n}})\big)\le ll_R\big(ann.coker(A)\big)+1$ and
also the ideals $\overline{I_m(A)}:\overline{I_{m-1}(A)}$, $ann.coker(A)$ are quite close and often coincide.
The proofs, examples and other corollaries are given in \S\ref{Sec.Proofs.Glr}.
\
We emphasize that part 1 of the proposition implies also criteria for the infinite-$G_r$-determinacy, see corollary \ref{Thm.Proofs.Infinite.Determinacy.Gr} for the $C^\infty(\R^p,0)$ case.
\subsection{The criteria for $G_{congr}$}\label{Sec.Results.Congruence}
Through this subsection we assume $m>1$.
Recall that for skew-symmetric matrices of even size $I_m(A)=(det(A))=(Pf(A))^2$, the square of the Pfaffian ideal of $A$. Moreover, $ann.coker(A)\supseteq Pf(A)$, see proposition \ref{Thm.Background.AntiSymmetric.Matrices.Pfaffians}.
For skew-symmetric matrices of odd size: $det(A)=0$ and $ann.coker(A)=\{0\}$. As the measure of non-degeneracy of such matrices we use the $(m-1)$'st Pfaffian ideal
\beq
Pf_{m-1}(A):=\suml_{i=1}^m (Pf(A_{\hi})).
\eeq
Here $A_{\hi}$ is the $(m-1)\times(m-1)$ block of $A$, obtained by erasing the $i$'th row and column.
(Note that $A_{\hi}$ is skew-symmetric and $Pf(A_{\hi})^2=I_{m-1}(A_{\hi})$.)
See \S\ref{Sec.Background.Pfaffians} for more detail.
\bthe\label{Thm.Results.Annihilator.T1.for.Gcongr} Suppose $I_{m-1}(A)\neq\{0\}$.
\bee
\item Let $\Si=Mat_{m\times m}(R)$, where $R$ satisfies the condition \eqref{Eq.Intro.assumption.R.Noetherian.or.surjects}. Suppose $dim(R)>0$, i.e. $R$ is not Artinian.
Then $ann(T^1_{(\Si,G_{congr},A)})\sseteq \cm^\infty$.
In particular, if $\cm^\infty=\{0\}$ then $ann(T^1_{(\Si,G_{congr},A)})=\{0\}$.
\item Suppose $R$ is Noetherian and $A\in \Si=Mat^{sym}_{m\times m}(R)$, then
\[
I_{m}(A):I_{m-1}(A)\sseteq ann(T^1_{\Si,G_{congr},A)})\sseteq\overline{I_{m}(A)}:\overline{I_{m-1}(A)}.
\]
\item Suppose $R$ is Noetherian and $A\in \Si=Mat^{skew-sym}_{m\times m}$, then
{\bf i.} If $m$ is even then
$I_{m}(A):I_{m-1}(A)\sseteq ann(T^1_{(\Si,G_{congr},A)})\sseteq\overline{I_{m}(A)}:\overline{I_{m-1}(A)}$.
{\bf ii.} If $m$ is odd then
$Pf_{m-1}(A)\sseteq ann(T^1_{(\Si,G_{congr},A)})\sseteq\overline{I_{m-1}(A)}:\overline{I_{m-2}(A)}$.
\eee
\ethe
As in the case of $G_{lr}$, if $R$ is UFD then in parts 2 and 3.i. holds:
$ann(T^1_{(\Si,G_{congr},A)})=I_m(A):I_{m-1}(A)$.
\
The bounds on $ann(T^1_{(\Si,G_{congr},A)})$ give immediate applications to the finite determinacy:
\bprop\label{Thm.Results.Finite.Determinacy.G_congr}
{\bf 1.} If $dim(R)>1$ then no $A\in Mat^{sym}_{m\times m}(\cm)$ is finitely-$(Mat^{sym}_{m\times m}(R),G_{congr})$-determined.
Similarly, no $A\in Mat^{skew-sym}_{m\times m}(\cm)$ with $m$-even is finitely-$(Mat^{skew-sym}_{m\times m}(R),G_{congr})$-determined.
{\bf 1'.} If $m$ is odd and $dim(R)>3$ then no $A\in Mat^{skew-sym}_{m\times m}(\cm)$ is finitely-$(Mat^{skew-sym}_{m\times m}(R),G_{congr})$-determined.
{\bf 1''.} If $dim(R)>0$ then no $A\in Mat_{m\times m}(\cm)$ is finitely-$(Mat_{m\times m}(R),G_{congr})$-determined.
{\bf 2.} Suppose $R$ is Henselian and satisfies condition \eqref{Eq.Intro.assumption.R.Noetherian.or.surjects}.
\quad {\bf i.} Suppose $dim(R)=1$ and either $\Si=Mat^{sym}_{m\times m}(R)$ or $\Si=Mat^{skew-sym}_{m\times m}(R)$ for $m$-even.
Let $A\in\Si$ and suppose $I_m(A)\supseteq\cm^\infty$. Then
\[
ll_\hR\Big(\overline{I_m(\hA)}:\overline{I_{m-1}(\hA)}\Big)-1\le ord^{\Si}_{G_{congr}}(A)\le ll_R\Big(I_m(A):I_{m-1}(A)\Big).
\]
In particular, $A$ is finitely-$(\Si,G_{congr})$-determined iff $det(A)$ is not a zero divisor in $R$.
In particular, the generic finite determinacy holds.
\quad {\bf ii.} Suppose $dim(R)\le3$ and $A\in \Si=Mat^{skew-sym}_{m\times m}(R)$, for $m$-odd.
Suppose moreover $I_{m-1}(A)\supseteq\cm^\infty$.
Then
\[ll_\hR\Big(\overline{I_{m-1}(\hA)}:\overline{I_{m-2}(\hA)}\Big)-1\le ord^{\Si}_{G_{congr}}(A)\le ll_R(Pf_{m-1}(A)).
\]
In particular, $A$ is finitely-$(\Si,G_{congr})$-determined iff either $I_{m-1}(A)=R$ or
$I_{m-1}(A)\sset R$ is of height$=dim(R)$, as expected.
In particular, the generic finite determinacy holds.
\eprop
The proofs, examples and further corollaries are in \S\ref{Sec.Proofs.Congruence}.
For the criteria of infinite
determinacy of $C^\infty(\R^p,0)$-valued (skew-)symmetric forms see corollary \ref{Thm.Proofs.Infinite.Determinacy.G_congr}.
\subsection{Upper-block-triangular matrices}
For an upper-block-triangular matrix $A\in Mat^{up}_{\{m_i\}\times\{n_j\}}(R)$ take the block-structure, $\{A_{ij}\}_{1\le i,j\le k}$,
as in example \ref{Ex.Intro.Typical.Groups}.
The context/motivation for this case, the definitions of $T^1_{(\Si,G^{up}_r,A)}$, $T^1_{(\Si,G^{up}_{lr},A)}$,
their roles, the proofs
and corollaries are in \S\ref{Sec.Proofs.Upper.Triangular}.
\bthe\label{Thm.Results.Annihilator.T1.Upper.Triang}
Let $R$ be a local ring and $A\in \Si=Mat^{up}_{\{m_i\}\times\{n_j\}}(R)$.
\bee
\item $ann(T^1_{(\Si,G^{up}_r,A)})=ann.coker(A)$. In particular:
i. $\prodl_{i=1}^k ann.coker(A_{ii})\sseteq ann(T^1_{(\Si,G^{up}_r,A)})\sseteq\capl_{i=1}^k ann.coker(A_{ii})$.
ii. If $m=n$ then $ann(T^1_{(\Si,G^{up}_r,A)})=I_m(A):I_{m-1}(A)$.
iii. If $n>m$ and $height(I_m(A))=(n-m+1)$ then $ann(T^1_{(\Si,G^{up}_r,A)})=I_m(A)$.
\item Suppose $R$ is Noetherian then
$\prodl_{i=1}^k ann.coker(A_{ii})\sseteq ann.coker(A)\sseteq ann(T^1_{(\Si,G^{up}_{lr},A)})\sseteq\capl_{i=1}^k \Big(\overline{I_{m_i}(A_{ii})}:\overline{I_{m_i-1}(A_{ii})}\Big)$.
\eee
\ethe
\subsection{The non-finite determinacy for conjugation}\label{Sec.Results.Conjugation}
The conjugation action is defined in example \ref{Ex.Intro.Typical.Groups}.
\bthe\label{Thm.Results.T1.Conjugation} Suppose $R$ satisfies condition \eqref{Eq.Intro.assumption.R.Noetherian.or.surjects} and let $A\in Mat_{m\times m}(R)$.
\bee
\item $ann(T^1_{(Mat_{m\times m}(R),G_{conj},A)})\sseteq\cm^\infty$.
\item In particular, if $dim(R)>0$ then there are no finitely-$G_{conj}$-determined matrices.
\eee
\ethe
\subsection{Finite determinacy of chains of free modules}
As an immediate application of our methods we consider bounded chains (or complexes) of free modules.
Take such a chain, $\cdots\stackrel{\phi_{i+1}}{\to} F_i\stackrel{\phi_i}{\to}F_{i-1}\stackrel{\phi_{i-1}}{\to}\cdots$,
fix some bases of $\{F_i\}$, so that the maps are represented by some matrices. The deformations of such a chain are taken up to isomorphisms, induced by the action of the product of groups $\prodl_i GL(F_i,R)$. We obtain the bounds on the determinacy in \S\ref{Sec.Proofs.Complexes}, quantifying and generalizing theorem 5.8 of \cite{Cutkosky-Srinivasan}.
\subsection{Relative determinacy/admissible deformations}\label{Sec.Results.Relative.Determinacy}
As we see in theorems \ref{Thm.Results.Annihilators.for.GlGrGlr}, \ref{Thm.Results.Annihilator.T1.for.Gcongr}, \ref{Thm.Results.Annihilator.T1.Upper.Triang}, the condition $ann(T^1_{(\Si,G,A)})\supseteq\cm^N$ can be quite restrictive. For example:
\bei
\item for the action $G_{lr}\circlearrowright\Mat$ it means:
$ann.coker(A)\supseteq\cm^N$, i.e. the module $coker(A)$ is of finite length. Geometrically: the module is supported only at one point, the origin, i.e. is a skyscraper.
\item
for the action $G_{congr}\circlearrowright Mat^{sym}_{m\times m}(R)$ it means: $A$ and its associated quadratic form are non-degenerate
on the punctured neighborhood of the origin.
\eei
However, one often needs to deform $A$ not inside the whole $\Mat$, or $\Si$, but only by elements of $Mat_{m\times n}(J)$ (for some ideal $J\sset R$) or
$\Si^{(J)}:=\Si\cap\Big(\{A\})+Mat_{m\times n}(J)\Big)$.
For example, one studies the determinacy for the action $G_{lr}\circlearrowright Mat_{m\times n}(J)$ or $G_{congr}\circlearrowright Mat^{sym}_{m\times m}(J)$.
Similarly, for a given group action, $G\circlearrowright\Mat$, one often considers the subgroup of transformations that are identities modulo some (prescribed) ideal $I\subsetneq R$. More precisely:
\beq\label{Def.Unipotent.Subgroup}
G^{(I)}:=\Big\{g\in G|\ g\cdot Mat_{m\times n}(I)=Mat_{m\times n}(I)\ and\ [g]=[Id]\circlearrowright\quotient{\Mat}{Mat_{m\times n}(I)}\Big\}.
\eeq
For example, for $I=\cm^{k}$ the group $G^{(\cm^k)}$ consists of elements that are identities up to the order
$(k-1)$.
\
Working with $\Si^{(J)}$ and $G^{(I)}$ leads to the notion of relative determinacy/admissible deformations.
For complex analytic non-isolated hypersurface singularities this was studied in \cite{Pellikaan88}, \cite{Pellikaan}, \cite{Siersma83}, \cite{Siersma},
\cite{de Jong-van Straten1990}, \cite{Jiang}.
For the $C^\infty(\R^p,0)$-version see \cite{Grandjean00}, \cite{Grandjean04}, \cite{Thilliez}, \cite{Cutkosky-Srinivasan}.
Our criteria directly adapt to this formulation, see proposition \ref{Thm.Proofs.Corollary.For.Rel.Determinacy} in \S\ref{Sec.Proofs.Relative.Determinacy}.
\subsection{Remarks}\label{Sec.Results.Remarks}
\subsubsection{}
Theorems \ref{Thm.Results.Annihilators.for.GlGrGlr}, \ref{Thm.Results.Annihilator.T1.for.Gcongr}, \ref{Thm.Results.Annihilator.T1.Upper.Triang}
and their corollaries
are absolutely explicit, ready-to-use criteria. Their role is similar to the classical Mather theorem (for $\k=\bar\k$), the transition from:
$\bullet$ ``the hypersurface germ is finitely determined iff its miniversal deformation is finite dimensional"
\\to
$\bullet$ ``the hypersurface germ is finitely determined iff it has at most an isolated singularity".
Moreover, we do not just establish the finite determinacy, but give rather tight bounds for the order of determinacy.
To emphasize, in most cases the module $T^1_{(\Si,G,A)}$ does not admit any explicit/short presentation, e.g. for $T^1_{(\Mat,G_{lr},A)}$
one cannot write anything more explicit than $Ext^1_R(coker(A),coker(A))$. But the bounds we get are quite
explicit and tight, the upper and lower bounds often have the same integral closure.
\subsubsection{``Negativity" of the results} As our results read, for many actions there are no finitely determined matrices in $\Matm$.
For example, there are no finitely determined $\cm$-valued quadratic forms when $dim(R)>1$ (proposition \ref{Thm.Results.Finite.Determinacy.G_congr}).
For the action $G_{conj}\circlearrowright Mat_{m\times m}(R)$ the situation is even worse, theorem \ref{Thm.Results.T1.Conjugation}.
Sometimes one insists on having enough finitely determined objects, e.g. one wants the generic finite determinacy in the sense of definition \ref{Def.Generic.Finite.Determinacy}. The two ways to achieve this are:
\bei
\item Either to enlarge the group, e.g. to use also the local coordinate changes/automorphisms of the ring, $Aut_\k(R)$. Then one considers the action of $G_{lr}\rtimes Aut(R)$, $G_{congr}\rtimes Aut(R)$, etc.
\item Or to restrict the possible deformations, e.g. to deform in a way that preserves the determinantal ideals, $\{I_j(A)\}$, or to preserve the characteristic polynomial for square matrices.
\eei
We do this in \cite{Belitski-Kerner3}.
\subsubsection{A ``theoretical" remark}
Fix some action $G\circlearrowright\Si\sseteq Mat_{m\times n}(R)$, we assume that $T_{(\Si,A)}$ is a free $R$-module. (This holds in our cases.)
Take the generating matrix $\cA_{G,A}$ of the module $T_{(GA,A)}$ so that $Im(\cA_{G,A})=T_{(GA,A)}\sseteq T_{(\Si,A)}$. Suppose $\cA_{G,A}$ is of size $m_1\times n_1$, here $m_1=rank(T_{(\Si,A)})$, then
\beq
T^1_{(\Si,G,A)}=\quotient{T_{(\Si,A)}}{T_{(GA,A)}}=
coker(\cA_{G,A}),\quad
\text{while}\quad T^1_{(Mat_{m_1\times n_1}(R),G_r,\cA_{G,A})}=coker(\cA_{G,A})^{\oplus n_1}.
\eeq
(Here on the right $G_r=GL(n_1,R)$.) In particular, $ann(T^1_{(\Si,G,A)})=ann(T^1_{(Mat_{m_1\times n_1}(R),G_r,\cA_{G,A})})$.
Therefore the study of $T^1$ for any $G$ action is ``embedded" into the study of
$T^1$ for the $G_r$ action, on the space of bigger size matrices. And the same ``embedding" holds for the determinacy problems.
\subsubsection{Finite determinacy of (non-)embedded modules and (skew-)symmetric forms}
The notion of finite determinacy and our results can be stated more algebraically as follows.
The projection $R\stackrel{jet_k}{\to}\quotients{R}{\cm^{k+1}}$ induces the functor, $mod(R)\to mod({\quotients{R}{\cm^{k+1}}})$,
defined by $M\to \quotients{M}{\cm^{k+1} M}$. This functor is trivially surjective on the objects, by the restriction of scalars. But the functor is not injective on the objects, e.g. if the presentation matrix of $M$ has all
its entries in $\cm^{k+1}$ then $\quotients{M}{\cm^{k+1} M}$ is a free $\quotients{R}{\cm^{k+1}}$ module. Our results give the regions of parameters (the size of presentation matrix, $dim(R)$) for which the functor is
"generically injective on the objects".
\
Let $M\sseteq R^{\oplus m}$ be a finitely
generated $R$-submodule. Fix some set of generators of $M$, combine them into the matrix $A$.
\bei
\item
The finite-$G_r$-determinacy of $A$ means:
``{\em if $M,N\sseteq R^{\oplus m}$ and $jet_k M=jet_k N$ for $R\stackrel{jet_k}{\to}\quotients{R}{\cm^{k+1}}$, then $M=N\sseteq R^{\oplus m}$}".
\item The finite-$G_{lr}$-determinacy of $A$ means:
``{\em if $M,N\sseteq R^{\oplus m}$ and $jet_k M=jet_k N$ for $R\stackrel{jet_k}{\to}\quotients{R}{\cm^{k+1}}$, then $M= \phi(N)\sseteq R^{\oplus m}$ for some $\phi\in GL(m,R)$}".
\eei
\
Similarly for (skew-)symmetric forms the finite determinacy means that the form is determined (up to the change of generators) by its $k$'th jet for $k\gg0$.
\subsubsection{Algebraization}\label{Sec.Remarks.Algebraization}
A trivial consequence of finite determinacy is the algebraization:
the $G$-orbit of a given object contains many algebraic representatives and they are obtained by just cutting the
``Taylor-tails". Moreover, the order of determinacy gives an upper bound on the possible degrees of polynomials.
As we show in \S\ref{Sec.Background.Fin.Determin.vs.Stable.Algebraizability}, finite determinacy is equivalent to stable algebraizability: any higher order deformation of
a given object is $G$-equivalent to an algebraic family.
Even if $A$ is not finitely determined, i.e. $ann(T^1_{(\Si,G,A)})$ does not contain any $\cm^N$, one can consider
some ideal $J$ such that $J+ann(T^1_{(\Si,G,A)})$ contains a power of the maximal ideal. Then one gets the ``algebraization modulo $J$", or the ``algebraization with respect to some subset of variables". Therefore our results generalize those of \cite{Elkik73}, \cite{Kucharz} and many others.
\subsection{Relation to Singularity Theory}\label{Sec.Results.Relation.to.Singularity.Theory}
The determinacy is a classical notion of Singularity Theory, well studied for functions and maps over the ``classical" rings, when $\k=\R$ or $\C$, and $R=\k[[\ux]]$, $\k\{\ux\}$, $\k[\ux]_{(\ux)}$, $C^{\infty}(\R^p,0)$,
\cite{AGLV}, \cite{Damon1984}, \cite{Wall-1981}, etc.
The generic finite determinacy has been considered in \cite{Wall-1979}, \cite{du Plessis1982}, \cite{du Plessis1983}.
(For a short discussion of the results on the determinacy of functions/maps see \cite[\S2]{Belitski-Kerner}.)
\subsubsection{Some known results on ``matrix singularities"}\label{Sec.Results.Previous.Results.Singular.Theory}
Some necessary and sufficient conditions for finite determinacy of modules over the ring $C^\infty(\R^p,0)$ were obtained in \cite{Tougeron1968}.
The square matrices over $R=\k\{x_1,\dots,x_p\}$, for $\k\in\R,\C$, and $G=G_{lr}\rtimes Aut(R)$,
were considered in \cite{Haslinger}, \cite{Bruce2003}, \cite{Bruce-Tari04}, and further studied in \cite{Bruce-Goryun-Zakal02},
\cite{Bruce2003}, \cite{Goryun-Mond05}, \cite{Goryun-Zakal03}. In particular, the generic finite
determinacy was established and the simple types were classified.
In \cite{Cutkosky-Srinivasan} they study the finite determinacy of matrices for the actions of $G_r$, $G_{lr}$, $G_r\rtimes Aut_\k(R)$, $G_{lr}\rtimes Aut_\k(R)$ on matrices over complete local rings.
They obtain the qualitative results for the finite determinacy.
In \cite{Damon.Pike.1}, \cite{Damon.Pike.2} they study the vanishing topology of matrix singularities, relating these to the free divisors. In \cite{Fr.Kr-Za.2015} they study the vanishing topology of determinantal singularities.
The study of determinacy in positive characteristic has been initiated in \cite{Greuel-Pham}, \cite{Pham}.
\
We emphasize that most of the previous works addressed only the general criteria for the finite determinacy. In our work we directly compute (or at least bound) the order(s) of determinacy.
\subsubsection{}
In our results we see the standard dichotomy: either the finite determinacy is the generic property or there are no finitely determined matrices in $\Matm\cap\Si$.
The later case is the analogue of the ``bad dimensions" for the determinacy of maps $Maps\Big((\k^n,0),(\k^m,0)\Big)$, \cite[III.1]{AGLV}.
\section{Background and Preparations}\label{Sec.Background}
We denote the zero matrix by $\zero$, the identity matrix by $\one$.
\subsection{The relevant canonical forms for matrices}\label{Sec.Background.Canon.Forms.Matrices}
Let $(R,\cm)$ be a local ring over a field $\k$, $char(\k)=0$. (As a trivial case we allow $R=\k$, with $\cm=\{0\}$.)
Fix the embedding $\k\sset R$ by the decomposition of vector spaces $R=\k\oplus\cm$. Any element of $R\smin\cm$ is invertible.
Given $A\in Mat_{m\times n}(R)$ we sometimes ``compute it at the origin", i.e. take its image over the residue field, $A|_0\in Mat_{m\times n}(\k)$.
In particular, $A|_0=\zero$ means $A\in Mat_{m\times n}(\cm)$. As $A|_0$ is a matrix over a field, we take its classical rank, $r=rank(A|_0)$.
Recall that the rank of a skew-symmetric matrix is necessarily even.
We denote by $E=\oplusl_i \bpm 0&1\\-1&0\epm$ the canonical skew-symmetric matrix, the size (i.e. the number of summands) should be clear from the context.
\bprop\label{Thm.Background.Chip.off.Unit}
\bee
\item Let $A\in\Mat$ then $A\stackrel{G_{lr}}{\sim}\one_{r\times r}\oplus\tA$, where $r=rank(A|_0)$ and
$\tA\in Mat_{(m-r)\times(n-r)}(\cm)$.
\item Let $A\in Mat^{sym}_{m\times m}(R)$ then $A\stackrel{G_{congr}}{\sim}(\oplusl_i \la_i\one_i)\oplus\tA$,
where $\{ \la_i\in R\smin\cm\}_i$ and $\tA\in Mat^{sym}_{(m-r)\times(m-r)}(\cm)$.
\item Let $A\in Mat^{skew-sym}_{m\times m}(R)$ then $A\stackrel{G_{congr}}{\sim}\Big(\oplusl_{i} \la_i E_i\Big)\oplus\tA$,
where $\{ \la_i\in R\smin\cm\}_i$ and $\tA\in Mat^{skew-sym}_{(m-r)\times(m-r)}(\cm)$.
\eee
\eprop
Part 1 is known in Commutative Algebra e.g. as ``passing to the minimal resolution of the module $coker(A)$".
Similarly, in parts 2,3 we split a (skew-)symmetric form into its regular part (in the canonical form) and the complement - the `purely degenerate' part.
(Part 2. is proved e.g. in \cite[theorem 3, page 345]{Birkhoff-MacLane}, for part 3 see \cite[exercise 9, page 347]{Birkhoff-MacLane}.)
\
Let $R$ be a discrete valuation ring (DVR) over $\k$. In particular $R$ is local, regular and $dim(R)=1$.
The simplest examples of such rings are the formal power series, $R=\k[[t]]$, or the analytic series, $R=\k\{t\}$, when $\k$ is a normed field.
Note that the ring of real-valued smooth function-germs, $C^\infty(\R^1,0)$, is not a DVR.
Matrices over DVR often have good canonical forms.
\bprop\label{Thm.Background.Canonical.Form.Matrix.over.DVR} Suppose $R$ is a DVR over $\k$.
\bee
\item For any $A\in \Mat$ holds: $A\stackrel{G_{lr}}{\sim}\bpm \la_1&0&\dots&\dots&0\\0&\la_2&0&\dots&\dots\\\dots&\dots&\dots&\dots&\dots\\0&\dots&&\la_{m}&0\epm$, where $R\supseteq(\la_1)\supseteq(\la_2)\supseteq\cdots\supseteq(\la_m)$
\item For any $A\in Mat^{sym}_{m\times m}(\cm)$ holds: $A\stackrel{G_{congr}}{\sim}\oplusl_i \la_i\one_i$,
where $R\supseteq(\la_1)\supseteq(\la_2)\supseteq\cdots\supseteq(\la_m)$.
\item For any $A\in Mat^{skew-sym}_{m\times m}(\cm)$ holds: $A\stackrel{G_{congr}}{\sim}\Big(\oplusl_{i} \la_i E_i\Big)\oplus \zero$, where $R\supseteq(\la_1)\supseteq(\la_2)\supseteq\cdots$, while $\zero$ is the zero matrix of an appropriate size.
\eee
\eprop
Here part 1 is the Smith normal form.
Considering $A$ as a presentation matrix of the $R$-module $coker(A)$, the statement is:
``every module over a DVR is a direct sum of cyclic modules". Parts 2,3 read: every (skew-)symmetric form over a DVR splits into the
direct sum of rank-one forms.
While parts 2,3 are standard facts, we did not find an exact reference, thus we sketch a proof. Let $t$ be a uniformizing parameter
of $R$. Denote by $p$ the minimal among the vanishing orders of the entries of $A$. Then $t^{-p}A$ is a (skew-)symmetric matrix over
$R$ and $t^{-p}A|_0\neq\zero$. Apply proposition \ref{Thm.Background.Chip.off.Unit} to get, in the symmetric case,
$A\stackrel{G_{congr}}{\sim}t^{p}\Big((\oplus_i\la_i\one_i) \oplus \tA\Big)$,
(or $A\stackrel{G_{congr}}{\sim}t^{p}\Big((\oplus_i\la_i E_i) \oplus \tA\Big)$ in the skew-symmetric case).
Thus we get the reduction in size and one repeats for $\tA$.
\
The ideals $\{(\la_i)\}$ are called {\em the invariant ideals} of the matrix. They are expressible through the determinantal ideals of $A$ in the obvious way.
\subsection{Determinantal ideals and the annihilator-of-cokernel}\label{Sec.Background.Fitting.Ideals.ann.coker}
\cite[\S 20]{Eisenbud}
For $0< j\le m$ and $A\in\Mat$ denote by $I_j(A)\sset R$
the $j$'th determinantal ideal, generated by all the $j\times j$ minors of $A$.
By definition $I_{j\le 0}(A)=R$ and $I_{j>m}(A)=\{0\}$. When the size of $A$ is not stated explicitly we denote the ideal of maximal minors by $I_{max}(A)$.
The chain of ideals
$R=I_0(A)\supseteq I_1(A)\supseteq\cdots\supseteq I_m(A)$ is {\em invariant} under the $G_{lr}$-action.
Determinantal ideals behave nicely under homomorphisms of rings, i.e. for any $R\stackrel{\phi}{\to}S$ and the induced
$\Mat\stackrel{\phi}{\to}Mat_{m\times n}(S)$ one has: $I_j(\phi(A))=\phi(I_j(A))$.
\
If $A|_0=\zero$, i.e. $A\in\Matm$, then
the height of $I_j(A)$ is at most $min\Big((m+1-j)(n+1-j),dim(R)\Big)$ and this bound is achieved generically.
More precisely, for any $A\in\Matm$, any $N>0$ and any generic $B\in Mat_{m\times n}(\cm^N)$ the height of the ideal $I_j(A+B)$ equals
$min\Big((m+1-j)(n+1-j),dim(R)\Big)$.
For $A\in Mat^{sym}_{m\times m}(\cm)$ the expected height of $I_j(A)$ is $min\Big(\bin{m-j+2}{2},dim(R)\Big)$. For $A\in Mat^{skew-sym}_{m\times m}(\cm)$ and $j$ even the expected height of $I_j(A)$ is
$min\Big(\bin{m-j+2}{2},dim(R)\Big)$.
For $A\in Mat^{skew-sym}_{m\times m}(\cm)$ and $j$ odd the expected height of $I_j(A)$ is
$min\Big(\bin{m-j+1}{2},dim(R)\Big)$.
\
A matrix $A\in\Matm$ can be considered as a homomorphism of free modules, its cokernel is an $R$-module as well:
\beq
R^{\oplus n}\stackrel{A}{\to} R^{\oplus m}\to coker(A)\to0.
\eeq
Then one takes the annihilator-of-cokernel ideal:
\beq
ann.coker(A)=\{f\in R|\ f\cdot R^{\oplus m}\sseteq Im(A)\}=ann\quotients{R^{\oplus m}}{Im(A)}\sset R.
\eeq
This ideal is $G_{lr}$-invariant and is a refinement of the determinantal ideal $I_m(A)$.
\
The annihilator-of-cokernel is a rather delicate invariant, for example it does not transform nicely under the ring homomorphisms, $ann.coker(\phi(A))\neq \phi(ann.coker(A))$. Yet, this ideal is controlled by
the determinantal ideals as follows:
\beq\label{Eq.Background.ann.coker.in.terms.of.Fitting.ideals}\ber
ann.coker(A)^m\sseteq I_m(A)\sseteq ann.coker(A)\sseteq\sqrt{I_m(A)},
\\
\text{and}\ \forall\ m>j>1:\quad ann.coker(A)\cdot I_{j-1}(A)\sseteq I_j(A)
\text{ \cite[proposition 20.7]{Eisenbud}}
\\
\text{If $m=n$ then $ann.coker(A)=I_m(A):I_{m-1}(A)$.}
\\
\text{If $m<n$ and $height(I_{m})=(n-m+1)$ then $ann.coker(A)=I_m(A)$, \cite[exercise 20.6]{Eisenbud}}
\eer\eeq
In particular, for one-row matrices, $m=1$, or when $I_m(A)$ is a radical ideal, $I_m(A)=ann.coker(A)$.
We use the following properties:
\bel\label{Thm.Background.Ann.Coker.Properties}
\bee
\item (Block-diagonal case) $ann.coker(A\oplus B)=ann.coker(A)\cap ann.coker(B)$.
\item If $m=n$, i.e. $A$ is a square matrix, then $ann.coker(A)=ann.coker(A^T)$.
If moreover $R$ is a unique factorization domain (UFD) then $ann.coker(A)$ is a principal ideal.
\item If $m>n$ then $ann.coker(A)=\{0\}$.
\item Let $A_\Box\sset A$ be a submatrix of $A$ obtained by erasing some rows of $A$. Then $ann.coker(A)\sseteq ann.coker(A_\Box)$.
\eee\eel
Here statements {\bf 1} and {\bf 4} are immediate.
The first part of Statement {\bf 2} follows directly, e.g. from $ann.coker(A)=I_m(A):I_{m-1}(A)$ of equation
(\ref{Eq.Background.ann.coker.in.terms.of.Fitting.ideals}).
For the second part of {\bf 2} note that for a square matrix the height of $ann.coker(A)\sset R$ is one. If $R$ is UFD then $ann.coker(A)$ is generated by just one element.
Statement {\bf 3}: in this case the submodule $Im(A)\sset R^{\oplus m}$ is of rank$\le n<m$.
Thus the module $coker(A)$ is supported on the whole $Spec(R)$ and is not annihilated by any elements of $R$.
\subsection{Skew-symmetric matrices and their Pfaffian ideals}\label{Sec.Background.Pfaffians}
Suppose $A$ is a square matrix then $I_m(A)=(det(A))$. If moreover $A\in Mat^{skew-sym}_{m\times m}(R)$, with $m$-even,
then the determinant is a full square, $det(A)=Pf(A)^2$. Here the Pfaffian polynomial is explicitly written in terms of the entries of $A$,
\cite{Heymans}, \cite[exercise A.2.11]{Eisenbud}, \cite{Greub}.
In particular, the Pfaffian polynomial transforms under the homomorphism of rings: $Pf(\phi(A))=\phi(Pf(A))$.
The transformation of the Pfaffian under the congruence is: $Pf(UAU^T)=det(U)Pf(A)$.
If $m$ is odd then $det(A)=0$. In this case, as a measure of degeneracy, we need the following refinement of the Pfaffian ideal:
\beq
Pf_{m-1}(A):=\suml_{i=1}^m \big(Pf(A_{\hi})\big).
\eeq
Here $A_{\hi}$ is the $(m-1)\times(m-1)$ block of $A$, obtained by erasing the $i$'th row and column.
(Note that $A_{\hi}$ is skew-symmetric and $I_{m-1}(A_{\hi})=\big(Pf(A_{\hi})^2\big)$.)
We use the following properties of these ideals:
\bel\label{Thm.Background.AntiSymmetric.Matrices.Pfaffians}
Let $R$ be a local ring.
\\1. Let $A\in Mat^{skew-sym}_{m\times m}(R)$ for $m$-even. There exists the ``Pfaffian-adjugate" matrix, $adj^{Pf}(A)\in Mat^{skew-sym}_{m\times m}(R)$, that satisfies $A\cdot adj^{Pf}(A)=Pf(A)\one=adj^{Pf}(A)\cdot A$.
In particular, $ann.coker(A)\supseteq Pf(A)$.
\\1'. Moreover: $Span_R(AU+U^TA)_{U\in Mat_{m\times m}(R)}\supseteq Mat^{skew-sym}_{m\times m}(Pf(A))$.
\\2. Let $A\in Mat^{skew-sym}_{m\times m}(R)$ for $m>1$, odd. Then
$Span_R(AU+U^TA)_{U\in Mat_{m\times m}(R)}\supseteq Mat^{skew-sym}_{m\times m}(Pf_{m-1}(A))$.
\eel
\bpr
{\bf 1.} We recall the relevant exterior algebra, \cite[exercise A.2.11]{Eisenbud}. Let $V$ be a free $R$-module of rank $n$. Its dual, $V^*$, is the module of one-forms.
Associate to $A$ the skew-symmetric form $w_A:=\suml_{1\le i<j\le m}a_{ij}dx_i\wedge dx_j\in\overset{2}{\wedge} V^*$.
The top exterior power of $w_A$ gives the Pfaffian:
\beq
\overset{\frac{m}{2}}{\wedge}w_A=m!\cdot Pf(A)dx_1\wedge\dots\wedge dx_m\in\overset{m}{\wedge} V^*.
\eeq
Now, using the pairing $\overset{2}{\wedge} V^*\otimes \overset{m-2}{\wedge} V^*\to \overset{m}{\wedge} V^*$, we associate to the form $\overset{\frac{m}{2}-1}{\wedge}w_A$ an element of $\overset{2}{\wedge} V^*$, which we call $adj^{Pf}(A)$. Then the relation $w_A\wedge (\overset{\frac{m}{2}-1}{\wedge}w_A)=Pf(A)dx_1\wedge\dots\wedge dx_m$ transforms into $A\cdot adj^{Pf}(A)=Pf(A)\one$.
{\bf 1'.} This is now immediate, one takes $U$ in the form $adj^{Pf}(A)\tilde{U}$, where $\tilde{U}\in Mat^{skew-sym}_{m\times m}(R)$.
{\bf 2.} By proposition \ref{Thm.Background.Chip.off.Unit} $A$ is congruent to the block-diagonal matrix $E\oplus\tA$, where $E$ is a non-degenerate skew-symmetric matrix over $R$ (i.e. $det(E)$ is invertible), while $\tA|_0=\zero$. Moreover, this splitting is stable, any family $A_t$ is congruent to
$E_t\oplus\tA_t$. Therefore it is enough to prove the statement for $\tA$ and from now on we assume $A\in Mat^{skew-sym}_{m\times m}(\cm)$.
As the Pfaffian, $Pf(A)$, is an explicit polynomial in the entries of the matrix $A^T=-A$, it is functorial under the change of rings. Therefore it is enough to prove the statement for the skew-symmetric matrix of indeterminates, $A=(a_{ij})$, over the ring $R=\k[[\{a_{ij}\}]]$. In this case the ideal $Pf_{m-1}(A)$ is Gorenstein and of height $3$. Therefore we can use the classical result, \cite[page 457]{Buchsbaum-Eisenbud} (for $m$ odd):
\beq\label{Eq.inside.proof.Pfaffians}
\text{If $B\in Mat^{skew-sym}_{m\times m}(R)$ and $Pf_{m-1}(B)=Pf_{m-1}(A)$, then $A\stackrel{G_{congr}}{\sim}B$.}
\eeq
Our statement is just the infinitesimal version of this result, and the proof goes as follows.
For $m=3$ the statement is checked directly, thus we assume $m>3$.
Let $\De\in Mat^{skew-sym}_{m\times m}(Pf_{m-1}(A))$. We claim that the two ideals coincide: $(Pf_{m-1}(A))=(Pf_{m-1}(A+\De))$. Indeed, for any block $A_\hi$ take the standard expansion of the Pfaffian polynomial $Pf(A_\hi)$ in the entries of $A_\hi$. Then we get:
\beq
Pf(A_\hi+\De_\hi)=Pf(A_\hi)+\underbrace{(\dots)}_{\in\cm\cdot (Pf(A_\hi))}.
\eeq
Therefore for the ideals we have: $\Big(Pf(A_\hi+\De_\hi)\Big)\sseteq \Big(Pf(A_\hi)\Big)$. The converse inclusion follows by the expansion
\beq
Pf\big((A_\hi-\De_\hi)+\De_\hi\big)=Pf(A_\hi-\De_\hi)+\cdots.
\eeq
Thus
$\Big(Pf(A_\hi+\De_\hi)\Big)=\Big(Pf(A_\hi)\Big)$.
As this holds for any $i$ we get:
$\Big(Pf_{m-1}(A+\De)\Big)=\Big(Pf_{m-1}(A)\Big)$.
Now, by \eqref{Eq.inside.proof.Pfaffians}, we get $A+\De\stackrel{G_{congr}}{\sim}A$. Finally, take the infinitesimal family of such congruences,
$\{A+\ep\De\stackrel{G_{congr}}{\sim}A\}_{\ep\in\k[\ep]/(\ep^2)}$. And take the first order terms in $\ep$, to get:
$\De\in T_{(G_{congr}A,A)}=Span_R(UA+AU^T)_{U\in Mat_{m\times m}(R)}$.
\epr
\subsection{Integral closure of ideals}\label{Sec.Background.Integral.Closure.Ideals}
The integral closure of an ideal $I\sset R$ is defined as
\beq
\bI=\Big\{f\in R\big| \ f^d+\suml^d_{j=1}a_j f^{d-j}=0,\ \text{for some }d\in\N \text{ and some } \{a_j\in I^j\}_{j=1,\dots,d}\Big\}.
\eeq
This subset is an ideal in $R$ and is itself integrally closed, i.e. $\bar{\bI}=\bI$.
\bex \cite[\S1.4]{Huneke-Swanson}
Fix a monomial ideal $I\sset R=\k[\![x_1,\dots,x_p]\!]$. Take the set of all its exponents:
$\cS_I=\{\ud|\ \ux^\ud\in I\}$. This leads to the diagram of initial exponents (the Newton diagram),
$\Ga_I:=\{\ud|\ \ud\in Conv(\cS_I+\R^p_{\ge0})\}$. Then the integral closure is also a monomial ideal, $\bI=\{\ux^\ud|\ \ud\in \Ga_I\}$.
\eex
\bex
Let $R=\quotients{\k[[\ux]]}{f(\ux)}$, where $f$ is in the Weierstra\ss\ form, $f(x_1,\dots,x_p)=x^d_p+x^{d-1}_pa_1+\cdots+a_d$, with the coefficients $a_j=a_j(x_1,\dots,x_{p-1})\in(x_1,\dots,x_{p-1})^j$. Then obviously $x_p\in\overline{(x_1,\dots,x_{p-1})}$, i.e.
$\overline{(x_1,\dots,x_{p-1})}=\cm$. In particular, for the plane curve case, $p=2$, we have $x_2\in\overline{(x_1)}$.
\eex
We mention several useful properties:
\bei
\item if the height of $I$ is defined then $height(I)=height(\overline{I})$;
\item $\overline{I:J}\sseteq\overline{I}:J$, \cite[page 7]{Huneke-Swanson};
\item if $R$ is a domain and $I\neq\{0\}$ then $\overline{I\cdot J}:\overline{I}=\overline{J}$, \cite[Corollary 6.8.7]{Huneke-Swanson};
\item if $R$ is local, Noetherian and regular then $\overline{f\cdot J}=(f)\cdot \overline{J}$.
\eei
\
In general the computation of $\bI$ is rather involved. Things simplify due to the following ``geometric" criterion,
Theorem 6.8.3 of \cite{Huneke-Swanson}.
\beq\label{Eq.Integr.Closur.Val.Criter}
\text{If $R$ is Noetherian then }
\bI=\bigcapl_{R\stackrel{\phi}{\to}S_{DVR}}\phi^{-1} S\phi(I)
\eeq
(Here the intersection goes over all the homomorphisms to the discrete valuation domains.)
In words:
\beq
\text{$f\in \overline{I}$ iff for any homomorphism to a DVR, $R\stackrel{\phi}{\to}S,\ \phi(f)\in S\phi(I)$.}
\eeq
Geometrically this criterion (initially for $R=\C\{\ux\}$, \cite[1.3.4]{Teissier}) reads:
\beq
f\in \overline{I} \text{ iff for any map of the smooth curve-germ, } (C,0)\stackrel{\nu}{\to}Spec(R),
\text{ the pullbacks satisfy: } \nu^*(f)\in\nu^*(I).
\eeq
\subsection{Integral closure of modules}\label{Sec.Background.Integral.Closure.Modules}
For the general definition of the integral closure of modules see \cite[definition 16.1.1]{Huneke-Swanson}. In our case all the modules are embedded into $R^{\oplus m}$, the ring is Noetherian over a field, $char(\k)=0$, thus the definition simplifies. Note that any ring homomorphism $R\stackrel{\phi}{\to}S$ induces the homomorphism of embedded modules $R^{\oplus m}\supseteq M\stackrel{\phi}{\to}\phi(M)\sseteq S^{\oplus m}$.
\bed\cite[\S16.1]{Huneke-Swanson}
The integral closure of a module $M\sset\R^m$ is
$\overline{M}:=\bigcapl_{R\stackrel{\phi}{\to}S_{DVR}}\phi^{-1} S\phi(M)$.
\eed
In words: $z\in\overline{M}$ iff for any homomorphism to a discrete valuation domain, $R\stackrel{\phi}{\to}S_{DVR}$, there holds $\phi(z)\in S\phi(M)$.
\
We will use the property, \cite[Remark 1.3.2]{Huneke-Swanson}:
\beq\label{Eq.Background.integral.closure.product.of.ideals}
J_1\cdot\overline{J_2}\sseteq \overline{J_1}\cdot\overline{J_2}\sseteq \overline{J_1\cdot J_2},
\eeq
More generally:
\bel\label{Thm.Background.Integral.Closure.Lemma}
Let $R$ be local, Noetherian, $N\sseteq M\sseteq R^{\oplus m}$ embedded modules, $J\sset R$ an ideal.
\bee
\item $J\cdot\overline{M}\sseteq \overline{J}\cdot\overline{M}\sseteq \overline{J\cdot M}$.
\item $ann\quotients{\overline{M}}{\overline{N}}\supseteq\overline{ann\quotients{M}{N}}$.
\eee
\eel
\bpr
1. The first inclusion is trivial. For the second it is enough to prove: for any $f\in \overline{J}$, $z\in\overline{M}$ holds $fz\in \overline{J\cdot M}$.
Let $R\stackrel{\phi}{\to}S_{DVR}$ then $\phi(f)\in S\cdot \phi(J)$ and $\phi(z)\in S\cdot\phi(M)$.
Thus $\phi(fz)=\phi(f)\phi(z)\in S\cdot \phi(J)\cdot\phi(M)=S\cdot \phi(J\cdot M)$. As this holds for any such $\phi$ we get $fz\in \overline{J\cdot M}$.
2. Let $f\in \overline{ann\quotients{M}{N}}$, i.e. for any $R\stackrel{\phi}{\to}S_{DVR}$ holds $\phi(f)\in \phi(ann\quotients{M}{N})$. Thus $f\in ann\quotients{M}{N}+ker(\phi)$, i.e. $f\cdot M\sseteq N+ker(\phi)$.
But then $\phi(f)\phi(M)\sseteq\phi(N)$. As this holds for any $\phi$, we get: $\overline{f\cdot M}\sseteq\overline{N}$. Finally, by part (1.), $f\cdot \overline{M}\sseteq\overline{f\cdot M}\sseteq\overline{N}$, i.e. $f\in ann\quotients{\overline{M}}{\overline{N}}$.
\epr
\bex
The ideals in part (2) can differ significantly. For example, let $R=\k[[x_1,\dots,x_p]]$, with $p\ge3$, and let $N=(x^d_1,\dots,x^d_p)\subsetneq M=\cm^d$.
Then $\overline{N}=\overline{M}=\cm^d$, hence $ann\quotients{\overline{M}}{\overline{N}}=R$. But $\overline{ann\quotients{M}{N}}=\cm^{d}$.
Proof: obviously $N\sseteq ann\quotients{M}{N}$,
thus $\overline{ann\quotients{M}{N}}\supseteq\overline{N}=\cm^{d}$. Furthermore, $\overline{ann\quotients{M}{N}}$ is a monomial ideal.
Suppose $i_1+\cdots +i_p=d-1$ and
$x^{i_1}_1\cdots x^{i_p}_p\in \overline{ann\quotients{M}{N}}\smin \cm^d$.
Take the monomial $x^{d-1-i_1}\cdots x^{d-1-i_p}_p\in \cm^{p(d-1)-(d-1)}\overset{p\ge3}{\sset}\cm^d=M$.
Then
\[
(x^{d-1-i_1}\cdots x^{d-1-i_p}_p)\cdot (x^{i_1}_1\cdots x^{i_p}_p)=x^{d-1}_1\cdots x^{d-1}_p\not\in N.
\]
Hence $(x^{i_1}_1\cdots x^{i_p}_p)\cdot M\not\subseteq N$, contradicting the assumption.
\eex
The following lemma compares the bounds in part three of theorem \ref{Thm.Results.Annihilators.for.GlGrGlr}.
\bel\label{Thm.Background.Comparing.ann.coker.vs.integral.closures}
If $R$ be a local Noetherian ring then $\overline{ann.coker(A)}\sseteq \overline{I_m(A)}:\overline{I_{m-1}(A)}$.
\eel
\noindent Indeed, by equation \eqref{Eq.Background.ann.coker.in.terms.of.Fitting.ideals} we have $ann.coker(A)\cdot I_{m-1}(A)\sseteq I_m(A)$.
Thus the statement follows from part 1 of lemma \ref{Thm.Background.Integral.Closure.Lemma}.
\subsection{The largest ``reasonable" group is $G_{lr}$ (matrix preservers over rings)}\label{Sec.Background.Largest.Reasonable.Group}
When looking for the possible groups acting on matrices, $G\circlearrowright\Mat$,
the guiding principle is that the action of $G$ should, at least, distinguish between degenerate and
non-degenerate matrices. Groups with such properties are restricted by the following proposition.
For the general introduction to the theory of preservers, i.e. self-maps of $Mat_{m\times n}(\k)$, that preserve
some properties/structures see \cite{Molnar}, \cite{Li-Pierce}.
\bprop
Let $R$ be a local ring over a field $\k$, $dim(R)>0$, and suppose $R$ is a unique factorization domain. Let $T$ be an invertible map acting $R$-linearly on the module $Mat_{m\times n}(R)$, for $m\le n$, possibly violating the row/column structure.
Suppose $T$ preserves the set of degenerate matrices, i.e. $I_m(A)=\{0\}$ iff $I_m(T(A))=\{0\}$.
Then either $T(A)=UAV$, for some $U\in GL(m,R)$, $V\in GL(n,R)$, or, when $m=n$, $T(A)=UA^tV$.
\eprop
\bpr
We pass to the field of fractions $Frac(R)$. As $R$ has no zero-divisors, it is naturally embedded, $R\sset Frac(R)$.
Thus $T$ extends to an invertible operator that acts $Frac(R)$-linearly on
the vector space of matrices $Mat_{m\times n}(Frac(R))$
and such that $I_m(A)=\{0\}$ if and only if $I_m(T(A))=\{0\}$.
In other words, $T$ preservers the degenerate matrices, of $rank<m$, and also preserves the non-degenerate matrices of $rank=m$.
Now we use the classical results on the preservers over a field:
\bee[i.]
\item(the case $m=n$) \cite[Theorem 3]{Dieudonne}: Let $\k$ be an arbitrary field. Any invertible $\k$-linear transformation of $Mat_{m\times m}(\k)$ that preserves the set of degenerate matrices has either the form $T(A)=UAV$ or the form $T(A)=UA^tV$, for some $U,V\in GL(m,\k)$.
\item(the case $m<n$) \cite[Theorem 7]{Beasley-Laffey} Suppose the field $\k$ contains at least four elements. Any invertible $\k$-linear transformation of $Mat_{m\times n}(\k)$ that preserves the set of $rank=r$ matrices (for some fixed $0<r\le m$) is of the form $T(A)=UAV$ for some $U\in GL(m,\k)$, $V\in GL(n,\k)$.
\eee
Therefore we get: either $T(A)=UAV$, for some $U\in GL(m,Frac(R))$, $V\in GL(m,Frac(R))$, or, for $m=n$, $T(A)=UA^tV$. It remains to show that in fact $U\in GL(m,R)$, $V\in GL(n,R)$. (We check the case $m<n$, the case $m=n$ is similar.)
Substitute for $A$ all the elementary matrices (with 1 at the i'th row, j'th column and zeros otherwise), $e_{ij}$.
The $p,q$-entry of the matrix $Ue_{ij}V$ equals $u_{pi}v_{jq}$ and it belongs to $R$. Suppose some element of $V$ is not in $R$, we can assume
it is $v_{11}$. Present $v_{11}$ as a fraction of elements of $R$, $v_{11}=\frac{f}{g}$. As $R$ is UFD we can assume that $f,g$ are co-prime, i.e. $(f)\cap (g)=(fg)$ and $g$ is not invertible in $R$. As $u_{pi}v_{11}\in R$, for any $p,i$, we get: $U$ is divisible by $g$. Thus we pass from $(U,V)$ to $(\frac{U}{g},gV)$ and proceed by induction.
\epr
Therefore, in our work the group $G\circlearrowright\Mat$ is always a subgroup of $G_{lr}=GL_R(m)\times GL_R(n)$.
\beR The last statement is very sensitive to the particular form of condition ``$T$ preserves degenerate matrices".
For example, as the ring is local, one could ask for a condition: $T(A)\in Mat_{m\times n}(\cm^q)$ iff $A\in Mat_{m\times n}(\cm^q)$.
But this condition is satisfied just by $R$-linearity of $T\circlearrowright\Mat$, without any additional restrictions.
\eeR
\subsection{Unipotent subgroups}\label{Sec.Background.Unipotent.Subgroups}
Fix a group action on an $R$-module $G\circlearrowright M$ and a decreasing filtration $M=M_0\supsetneq M_1\supsetneq\cdots$. We get the induced filtration of $G$ by the
(normal) subgroups
\beq
G^{(i)}:=\{g\in G|\ \forall\ j\ge0\quad \forall\ z_j\in M_j:\ g(z_j)-z_j\in M_{j+i}\}.
\eeq
In particular, the subgroup $G^{(0)}\sseteq G$ preserves the filtration, while the subgroup $G^{(1)}\sseteq G$ acts unipotently. We call $G^{(1)}$ the ``unipotent" subgroup of $G$. If $G^{(1)}=G$ then $G$ itself is called ``unipotent for the given filtration".
The general properties of the filtration $G\supseteq G^{(0)}\supseteq G^{(1)}\supseteq\cdots$ are studied in \cite[\S3]{Belitski-Kerner}.
For the goals of the current paper we usually need the filtration by the powers of the maximal ideal,
\beq
\Mat\supsetneq \Matm\supsetneq Mat_{m\times n}(\cm^2)\cdots.
\eeq
Therefore we denote the unipotent subgroup by $G^{(\cm)}\sseteq G$. In particular:
\beq
G^{(\cm)}_r=\{U=\one+\tilde{U}|\ \tilde{U}\in Mat_{n\times n}(\cm)\},\quad
G^{(\cm)}_l=\{U=\one+\tilde{U}|\ \tilde{U}\in Mat_{m\times m}(\cm)\},\quad
G^{(\cm)}_{lr}=G^{(\cm)}_l\times G^{(\cm)}_r.
\eeq
The last equality is a statement, but the proof is straightforward, the conditions $\{Ue_{ij}V-e_{ij}\in Mat_{m\times n}(\cm)\}_{i,j}$ force $U\in G^{(\cm)}_l$, $V\in G^{(\cm)}_r$.
Furthermore, for any subgroup $G\sset G_{lr}$ we have: $G^{(m)}=G\cap G^{(\cm)}_{lr}$.
\subsection{The tangent spaces}\label{Sec.Background.Tangent.Spaces}
The tangent spaces to the groups and group-orbits are defined and studied in great generality in \cite[\S 3.7]{Belitski-Kerner}.
Here we record them for our group actions.
We use the isomorphism of $R$-modules, $T_{(\Mat,A)}\isom\Mat$, to identify $T_{(GA,A)}$, $T_{(\Si,A)}$ with their images in $\Mat$.
\bex\label{Ex.Tangent.Spaces.to.the.Actions} (See \cite[\S 3.8]{Belitski-Kerner}.)
\bei
\item
$G_{lr}:\ A\to UAV^{-1}$.
Here $T_{(G_{lr}A,A)}=Span_R\{uA,Av\}_{(u,v)\in Mat_{m\times m}(R)\times Mat_{n\times n}(R)}$. Similarly
for $G_l$ and $G_r$.
\item
$G_{congr}:\ A\to UAU^T$. Here $T_{(G_{congr}A,A)}=Span_R\{uA+Au^t\}_{u\in Mat_{m\times m}(R)}$.
\item
$G_{conj}:\ A\to UAU^{-1}$. Here $T_{(G_{conj}A,A)}=Span_R\{uA-Au\}_{u\in Mat_{m\times m}(R)}$.
\item
$G^{up}_{lr}:\ A\to UAV$, where $U,V$ are upper-block-triangular. Here
\[T_{(G_{lr}A,A)}=Span_R\{uA,Av\}_{(u,v)\in Mat^{up}_{\{m_i\}\times \{m_j\}}(R)\times Mat^{up}_{\{n_i\}\times \{n_j\}}(R)}.\]
Similarly for $G^{up}_r$.
\eei
\eex
In all these cases holds: $T_{(G^{(\cm)}A,A)}=\cm\cdot T_{(GA,A)}$.
In general, for all the ``reasonable" group actions there holds $\cm\cdot T_{(GA,A)}\subseteq T_{(G^{(\cm)}A,A)}\subseteq T_{(GA,A)}$,
see \cite[\S3.4]{Belitski-Kerner} for more detail.
\subsection{Transition to the completion}
When $R$ contains flat functions, i.e. $\ca^\infty\neq\{0\}$ for some proper ideal $\ca\ssetneq R$, we often take the $\ca$-adic completion, $R\stackrel{\hat{\psi}}{\to}\hR$, and compute over $\hR$. To translate the results to $R$ we need the following standard facts.
Given two embedded modules, $M_1\sset M_2\sset R^{\oplus p}$, take their images under completion,
$\hat{\psi}(M_1)\sseteq\hat{\psi}(M_2)\sseteq \hR^{\oplus p}$.
We assume that the completion map is surjective, $R\twoheadrightarrow\hR$, therefore the images are naturally $\hR$-modules: $\hat{\psi}(M_i)=\hR\cdot\hat{\psi}(M_i)$. Note that in general $\hat{\psi}(M_i)$ differs from the ``non-embedded" completion $\widehat{M_i}$. (For example, fix a non-zero divisor $f\in\ca^\infty$ and consider $I=(f)$. Then $\hat{\psi}(I)=\{0\}\sset\hR$, while $\widehat{I}\approx\hR$.)
\bel\label{Thm.Background.Completion.Lemma}
Suppose that the $\ca$-adic completion of a local ring is surjective, $R\stackrel{\hat\psi}{\twoheadrightarrow}\hR$. Take two modules
$M_1\sseteq M_2\sseteq R^{\oplus p}$ with $M_1\supseteq M_2\cap(\ca^\infty\cdot R^{\oplus p})$.
\\1. $\hat{\psi}(ann\quotients{M_2}{M_1})=ann\quotients{\hat{\psi}(M_2)}{\hat{\psi}(M_1)}$.
\\2. $ann\quotients{\hat{\psi}(M_2)}{\hat{\psi}(M_1)}\supseteq\hat{\psi}(\ca^k)$ iff
$ann\quotients{M_2}{M_1}\supseteq \ca^k$.
\\2'. In particular, for $\ca=\cm$ holds: $ll_R(ann\quotients{M_2}{M_1})=ll_\hR(\hat{\psi}(ann\quotients{M_2}{M_1}))$.
\eel
\bpr
1. The part $\sseteq$. Any element of $\hat{\psi}(ann\quotients{M_2}{M_1})$ has the form $\hat{\psi}(f)$, where $f\in ann\quotients{M_2}{M_1}$.
Then $fM_2\sseteq M_1$, hence $\hat{\psi}(f)\hat{\psi}(M_2)\sseteq\hat{\psi}(M_1)$ giving $\hat{\psi}(f)\in ann\quotients{\hat{\psi}(M_2)}{\hat{\psi}(M_1)}$.
The part $\supseteq$.
By the surjectivity $R\twoheadrightarrow\hR$ any element of $ann\quotients{\hat{\psi}(M_2)}{\hat{\psi}(M_1)}$ is of the form $\hat{\psi}(f)$ for some $f\in R$. As $\hat{\psi}(f)\hat{\psi}(M_2)\sseteq\hat{\psi}(M_1)$ we get: $fM_2\sseteq M_1+(M_2\cap\ca^\infty R^{\oplus p})=M_1$. Thus $f\in ann\quotients{M_2}{M_1}$.
2. $\Rrightarrow$
By part one we have $\hat{\psi}(ann\quotients{M_2}{M_1})\supseteq\hat{\psi}(\ca^k)$, thus $ann\quotients{M_2}{M_1}+\ca^\infty\supseteq\ca^k$. As $M_1\supseteq M_2\cap(\ca^\infty\cdot R^{\oplus p})\supseteq\ca^\infty M_2$ we have: $ann\quotients{M_2}{M_1}\supseteq\ca^\infty$, thus $ann\quotients{M_2}{M_1}\supseteq\ca^k$.
$\Lleftarrow$ If $ann\quotients{M_2}{M_1}\supseteq \ca^k$ then $\ca^k M_2\sseteq M_1$, thus $\hat{\psi}(\ca^k)\hat{\psi} (M_2)\sseteq\hat{\psi}(M_1)$, hence $ann\quotients{\hat{\psi}(M_2)}{\hat{\psi}(M_1)}\supseteq \hat{\psi}(\ca^k)$.
2'. As $M_1\supseteq M_2\cap(\cm^\infty\cdot R^{\oplus p})$ we get: $ann\quotients{M_2}{M_1}\supseteq\cm^\infty$ and
$\hat{\psi}(ann\quotients{M_2}{M_1})\supseteq\hat{\psi}(\cm^\infty)$. Thus the Loewy lengths of both sides in the statement are well defined. And the statement is immediate by part 2 for $\ca=\cm$.
\epr
\be
Fix two ideals, $I,J\sset R$, with $I\supseteq\Big((I+J)\cap\ca^\infty\Big)$, apply part one to the embedding $I\sseteq I+J\sset R$.
We get: $\hat{\psi}\Big(ann\quotient{I+J}{I}\Big)=ann\Big(\quotient{\hat{\psi}(I+J)}{\hat{\psi}(I)}\Big)$. Note that $I:J=ann\quotient{I+J}{I}$. Therefore we get: $\hat{\psi}(I:J)=\hat{\psi}(I):\hat{\psi}(J)$.
\eex
\subsection{The ``relevant approximation" property}\label{Sec.Background.Approximation.Properties}
For the criteria in the next subsection we need the ``relevant approximation property of the ring $R$". The particular property depends on the type of equations participating in the condition $gz=z+w$, $g\in G$.
While the general theory is developed in \cite{Belitski-Kerner},
for the subgroups of $G_{lr}$ we need only two types of approximation property:
\bee[i.]
\item If $G\sseteq G_{lr}$ is defined by $R$-linear equations and the condition $gA=A+B$ can be written as a system of linear equations on $g=(U,V)$ then
we say that $R$ has the relevant approximation property if
the assumption \eqref{Eq.Intro.assumption.R.Noetherian.or.surjects} together with the condition $ann.coker(A)\supseteq\cm^\infty$ hold.
The equations are linear for the groups $G_r$, $G_l$, $G_{lr}$, $C_{conj}$.
They are: $UA=A+B$, $AV=A+B$, $UA=(A+B)V$ and $UA=(A+B)U$.
\
Recall that
the ring $C^\infty(\R^p,0)$ satisfies condition \eqref{Eq.Intro.assumption.R.Noetherian.or.surjects}. Its completion is
Noetherian, $\R[\![\ux]\!]$, and the completion map, $C^\infty(\R^p,0)\twoheadrightarrow\R[\![\ux]\!]$, is surjective
by Borel's lemma, \cite[pg. 284, exercise 12]{Rudin-book}.
For this ring the condition $ann.coker(A)\supseteq\cm^\infty$ can checked using the following property:
\begin{multline}\label{Eq.Lojasiewicz.Ineq}
\text{If $f\in C^\infty(\R^p,0)$ satisfies $|f|\ge C|\ux|^\de$, for some $C>0$, $\de>0$,}
\\\text{then $f$ divides any function flat at the origin, and thus $(f)\supseteq\cm^\infty$.}
\end{multline}
The search for such an element is often simplified by the classical {\L}ojasiewicz inequality, \cite[\S IV.4]{Malgrange}:
\begin{multline}
\text{if $f\in C^\infty(\R^p,0)$ is analytic at the origin then $|f(x)|\geq C(dist (x,Z))^\de$,}
\\
\text{where $dist$ is the usual distance to the set of zeros,$Z=f^{-1}(0)$.}
\end{multline}
In particular, if $f$ has an isolated zero at the origin then $f$ satisfies $|f|\ge C|\ux|^\de$.
\item If $G\sseteq G_{lr}$ is defined by polynomial/analytic equations and the condition $gA=A+B$ can be written as a system of polynomial/analytic
equations on $g=(U,V)$,
we say that $R$ has the relevant approximation property provided $R$ is Henselian ring.
This is the case for $G_{congr}$, the equations being quadratic, $UAU^T=A+B$.
\eee
\
We mention that for $0<r<\infty$ the ring $C^r(\R^p,0)$ of r-times differentiable germs does not satisfy condition \eqref{Eq.Intro.assumption.R.Noetherian.or.surjects}. Its completion is not Noetherian, as the ideal
$\cm^{r+0^+}:=\{g\cdot\cm^r|\ g\in C^0(\R^p,0),\ g(0)=0\}$
is not finitely generated, and neither is its image in $\widehat{C^r(\R^p,0)}$.
\subsection{The transition from $ann(T^1_{(\Si,G,A)})$ to the finite determinacy and the main results of \cite{Belitski-Kerner}}\label{Sec.Background.BK1}
Fix a finitely generated $R$-module with descending filtration, $M=M_0\supset M_1\supset\cdots$.
We assume that the filtration is ``essentially descending", i.e. for any $j$ there exists $k_j$ such that
$M_{k_j}\sseteq \cm^j\cdot M$.
Fix a (filtered) group action, $G\circlearrowright M$, and a deformation space $\Si\sseteq M$. (In our case $M=\Mat$ or $Mat^{sym}_{m\times m}(R)$
or $Mat^{skew-sym}_{m\times m}(R)$.)
To understand the ``essential" deformations of an element $z\in \Si$, we need to know
``how small/large" is the orbit $Gz$ as compared to $\Si$. In many cases this question reduces to the linear version:
\beq\label{Eq.question}
\text{what is the biggest $R$-submodule $\La\sset M$ satisfying: $z+\La\sseteq Gz$?}
\eeq
(The precise statement is \cite[Lemma 2.1]{Belitski-Kerner}.)
The main result of \cite{Belitski-Kerner} is the transition of this linear version to the comparison of the
tangent spaces, $T_{(Gz,z)}\sseteq T_{(M,z)}$. We assume that the germ $(Gz,z)$ is ``good enough", in particular
it possesses the well defined tangent space. The precise condition is:
\beq\label{Eq.assumptions.of.kpd.fs.etc}\ber
\text{\em the subgroup $G\sset GL_\k(M)$ and its completion $\widehat{G}\sseteq GL_\k(\hM)$ are \kpd;}\\
\text{\em their unipotent parts, $G^{(1)}$, $\widehat{G}^{(1)}$, are of Lie type}.
\eer\eeq
(The condition ``$G$ is of Lie type" ensures that the tangent space $T_{(Gz,z)}$ ``approximates" the germ $(Gz,z)$.)
These properties hold for all our groups, $G_l,G_r,G_{lr}, G_{congr}, G_{conj},\dots$, see the details in \cite[\S3.7]{Belitski-Kerner}.
Furthermore, we use the isomorphism of $R$-modules $T_{(M,z)}\isom M$ to identify $T_{(Gz,z)}$ with its embedding.
Accordingly, we have the embedded module: $T_{(Gz,z)}\sseteq M$, as in \S\ref{Sec.Background.Tangent.Spaces}.
\bthe\label{Thm.Background.Linearization}\cite[Theorem 2.2]{Belitski-Kerner}
Suppose the (filtered) action $G\circlearrowright\{M_i\}$ satisfies assumptions \eqref{Eq.assumptions.of.kpd.fs.etc}. Suppose that $G$ is unipotent for the filtration $\{M_i\}$.
\bee
\item If $M_i\sseteq T_{(Gz,z)}$ and $R$ has the relevant approximation property then $\{z\}+M_i\sseteq Gz$.
\item Suppose $T_{(Gz,z)}\sseteq T_{(M,z)}$ is an $R$-submodule. If $\{z\}+M_i\sseteq Gz$ then $M_i\sseteq T_{(Gz,z)}$.
\eee
\ethe
While theorem \ref{Thm.Background.Linearization} is quite general, to compute/bound the order of determinacy we use a more specific criterion:
\bprop\label{Thm.Intro.ord.of.det.via.Loewy.length}
Suppose $\Si\sseteq\Mat$ is a free direct summand, i.e. $\Si\oplus\Si^\bot=\Mat$ for a free submodule $\Si^\bot\sset\Mat$.
\bee
\item \cite[Corollary 2.6]{Belitski-Kerner} Suppose $ann(T^1_{(\Si,G,A)})\supseteq\cm^\infty$ and $R$ has the relevant approximation property. Then
\[
ll_R\Big(ann(T^1_{(\Si,G,A)})\Big)-1\le ord^\Si_G(A)\le
ll_R\Big(ann(T^1_{(\Si,G^{(\cm)},A)})\Big)-1.
\]
\item If $ann(T^1_{(\Si,G,A)})\not\supseteq\cm^\infty$, then $A$ is not infinitely-$(\Si,G)$-determined.
\eee
\eprop
For all our groups holds $\cm\cdot T_{(GA,A)}\sseteq T_{(G^{(\cm)}A,A)}\sseteq T_{(GA,A)}$, thus
$\cm\cdot ann(T^1_{(\Si,G,A)})\sseteq ann(T^1_{(\Si,G^{(\cm)},A)})\sseteq ann(T^1_{(\Si,G,A)})$. So, the bounds in the corollary differ at most by one. Moreover, in many cases $ann(T^1_{(\Si,G,A)})=ann(T^1_{(\Si,G^{(\cm)},A)})$, see \S\ref{Sec.Background.Tangent.Spaces}.
This proposition reduces the determinacy question to the study of $T^1_{(\Si,G,A)}$ and its annihilator.
\beR
We mention the relation to the Castelnuovo-Mumford regularity of a module. Recall, \cite[exercise 20.18]{Eisenbud}:
if $M=\oplusl_i M_i$ is a graded $R$-module of finite length then its regularity equals $reg(M)=max\{i|\ M_i\neq\{0\}\}$.
Therefore, when $T^1_{(\Si,G,A)}$ is graded and of finite length, we can rewrite:
\beq
reg(T^1_{(\Si,G,A)})\le ord^\Si_G(A)\le reg(T^1_{(\Si,G^{(\cm)},A)}).
\eeq
\eeR
\subsection{Finite determinacy vs stable algebraizability}\label{Sec.Background.Fin.Determin.vs.Stable.Algebraizability}
Let $\quotients{\k[\ux]}{I}\subseteq R\subsetneq \hR=\quotients{\k[[\ux]]}{\k[[\ux]]\cdot I}$, where the typical examples of $R$
are: $\quotients{\k[\ux]}{I}$, $\quotients{\k\bl\ux\br}{I}$ (algebraic power series), $\quotients{\k\{\ux\}}{I}$ (analytic power series).
In this section we do not use any matrix structure, thus instead of the modules $\Mat$, $Mat_{m\times n}(\hR)$ we work with just free modules $F$, $\hF$.
As the completion is injective we have $F\sset \hF$.
Fix a group action $G\circlearrowright \hF$, $\k$-linear but not necessarily $\hR$-linear.
\bed
\bee[i.]
\item An element $v\in \hF$ is called $R$-algebraizable if $v\in G\tv$ for some $\tv\in F$.
\item An element $v\in \hF$ is called stably $R$-algebraizable if there exists $N\in\N$ such that for any deformation $v+u(t)$, with
$u(t)\in \cm^N\cdot \hF[[t]]$,
there exists a family $g(t)\in G[[t]]$ satisfying: $g(t)(v+u(t))\in F[[t]]$.
\eee
\eed
These algebraizability notions depend on the choice of $R$, the strongest notion is for $R=\quotients{\k[\ux]}{I}$.
Finite determinacy is stronger than $\quotients{\k[\ux]}{I}$-algebraizability.
However, for many choices of $R$, the finite determinacy is equivalent to the stable $R$-algebraizability.
Consider the ring $R$ as a $\k$-vector space, its dimension, $dim_\k(R)$, is infinite.
Recall that the vector spaces $R=\quotients{\k[\ux]}{I}$ and $R=\quotients{\k\bl\ux\br}{I}$ are of countable dimension.
But $\k[[x_1]]$ is of uncountable dimension.
\bprop
Suppose $dim_\k(R)< dim_\k\k[[x]]$.
Suppose $G\circlearrowright \hF$ is a Lie-type group (see \S\ref{Sec.Background.BK1}) and for any $v\in \hF$ the tangent space $T_{(Gv,v)}$ is an $\hR$-module.
Then $v\in\hF$ is $G$-finitely determined iff $v$ is stably $R$-algebraizable.
\eprop
\bpr
$\Rrightarrow$ As $v$ is finitely determined, we can cut off its Tailor tails, thus we assume $v\in F$.
Choose any $N$ bigger than the $G$-order of determinacy and take a deformation $u(t)=\suml_{i\ge 1}u_it^i$, where $u_i\in\cm^N\cdot \hF$.
Then exists $g_1(t)\in G[[t]]$ satisfying:
$g_1(t)(v+u_1t)=v\in F$. Thus $g_1(t)(v+u(t))=v+\suml_{i\ge2}\tu_i t^i$, with $\tu_i\in \cm^N\cdot \hF$. Then fix $g_2(t)$ satisfying $g_2(t)(v+\tu_2 t)=v$.
Thus
\beq
g_2(t^2)g_1(t)(v+tu_1+t^2u_2)=v+\suml_{i\ge3}\tilde{\tu}_it^i.
\eeq
In this way fix the sequence $\{g_i(t)\}$. Finally note that the
product $(\cdots g_k(t^k)g_{k-1}(t^{k-1})\cdots g_1(t))$ converges in $G[[t]]$, as it is a Cauchy sequence.
Thus exists $g(t)\in G[[t]]$ satisfying: $g(t)(v+u(t))=v\in F$.
$\Lleftarrow$ {\bf Step 1.} We prove $T_{(Gv,v)}\supseteq \cm^N\cdot \hF$, by theorem \ref{Thm.Background.Linearization}
this implies the finite determinacy.
By the assumption, for any $u(t)\in \cm^N\cdot\hF[[t]]$ there exists $g(t)\in G[[t]]$ and $v(t)\in F[[t]]$
satisfying $g(t)v(t)=v+u(t)$. We assume $v\in F$. Then, as $u(0)=0$, we assume $g(0)=Id$ and $v(0)=v$. Differentiate to get:
\beq
T_{(Gv,v)}+R^n\ni g'(0)v(0)+g(0)v'(0)=u'(0)\in \cm^N\cdot \hF.
\eeq
Here $u'(0)$ can be any element of $\cm^N\cdot \hF$, therefore we get the condition: $T_{(Gv,v)}+F\supseteq \cm^N\cdot \hF$. As the right hand side obviously
contains $F$ we get:
$T_{(Gv,v)}+F=\hF$.
Take some basis $\{e_i\}$ of $\hF$, and consider the i'th coordinate to get: $\Big(T_{(Gv,v)}\cap \hR(e_i)\Big)+R(e_i)=\hR(e_i)$.
Note that $T_{(Gv,v)}\cap \hR(e_i)$ is an ideal in $\hR$. Therefore it is enough to prove the general algebraic statement:
\beq
\text{
If an ideal $J\sset\hR$ satisfies $J+R=\hR$ then $J\supset \cm^N$
for some $N$.}
\eeq
{\bf Step 2.}
Consider the condition $J+R=\hR$ as an equality of $\k$-vector spaces and take the quotient of vector spaces: $\quotients{J+R}{J}=\quotients{\hR}{J}$.
Suppose $J$ does not contain any power of $\cm^N$, then, as $\hR$ is Noetherian, the Krull dimension of $\quotients{\hR}{J}$ is positive, denote it by $r\ge1$.
Then there exists a set of parameters of $\quotients{\hR}{J}$, i.e. some elements $y_1,..,y_r\in\quotients{\hR}{J}$ satisfying:
the ring $\quotients{\hR}{J+(y_1,..,y_r)}$ is Artinian and $\quotients{\hR}{J}\supseteq \k[[y_1,\dots,y_r]]$.
Now compare the dimensions of the vector spaces:
\beq
dim_\k\quotients{\hR}{J}\ge dim_\k\k[[y_1,\dots,y_r]]\ge dim_\k\k[[y_1]]>dim_\k(R).
\eeq
On the other hand, $dim_\k\quotients{J+R}{J}\le dim_\k(R)$, giving the statement.
\epr
\section{Proofs, examples and corollaries}\label{Sec.Proofs}
\subsection{The case of $G_l,G_r,G_{lr}\circlearrowright \Mat$}\label{Sec.Proofs.Glr}
\subsubsection{Approximation of $ann(T^1_{(\Si,G,A)})$} \
{\em Proof of theorem \ref{Thm.Results.Annihilators.for.GlGrGlr}}
{\bf 1.} Note that $T_{(G_rA,A)}=Span_R(AU)_{U\in Mat_{n\times n}(R)}=\oplusl^n_{i=1} Im(A)$, where $Im(A)\sset R^{\oplus m}$.
In addition, $T_{(\Si,A)}=\oplusl^n_{i=1} R^{\oplus m}$. Finally, the two decompositions are compatible, i.e.
$\quotients{T_{(\Si,A)}}{T_{(G_rA,A)}}=\oplusl^n_{i=1}\quotients{R^{\oplus m}}{Im(A)}=\oplusl^n_{i=1} coker(A)$.
{\bf 2.} The $(G_l,A)$ case is identical to the $(G_r,A^T)$ case. If $m<n$ then $ann.coker(A^T)=\{0\}$, by lemma \ref{Thm.Background.Ann.Coker.Properties}.
Similarly for $m=n$: $ann(T^1_{(\Si,G_l,A)})=ann(T^1_{(\Si,G_r,A^T)})=ann.coker(A^T)=ann.coker(A)$, again, lemma \ref{Thm.Background.Ann.Coker.Properties}.
{\bf 3.}
As $G_{lr}\supset G_r$, we have $T_{(G_{lr}A,A)}\supseteq T_{(G_rA,A)}$, thus part 1 gives the bound
\beq
ann\big(T^1_{(\Si,G_{lr},A)}\big)\supseteq ann\big(T^1_{(\Si,G_r,A)}\big)=ann.coker(A).
\eeq
Suppose $f\in ann\big(T^1_{(\Si,G_{lr},A)}\big)$,
then
\beq
f\cdot \Mat\sseteq Span_R\Big(UA+AV|\ (U,V)\in Mat_{m\times m}(R)\times Mat_{n\times n}(R)\Big).
\eeq
For any projection onto a discrete valuation domain, $R\stackrel{\phi}{\to}S$, we get:
\beq
\phi(f)\cdot Mat_{m\times n}(S)\sseteq Span\Big(U\phi(A)+\phi(A)V|\ (U,V)\in Mat_{m\times m}(S)\times Mat_{n\times n}(S)\Big).
\eeq
Now we use the canonical form, proposition \ref{Thm.Background.Canonical.Form.Matrix.over.DVR}, $\phi(A)=PD_{\phi(A)}Q$, where $D_{\phi(A)}$ has zeros off the main diagonal.
We get:
\beq
P^{-1}\phi(f)\cdot Mat_{m\times n}(S)Q^{-1}\sseteq Span\Big(UD_{\phi(A)}+D_{\phi(A)}V|\ (U,V)\in Mat_{m\times m}(S)\times Mat_{n\times n}(S)\Big).
\eeq
On the right hand side one has the space of matrices whose $(m,m)$'th component lies in the ideal $(\la_m)$.
On the left hand side one has $Mat_{m\times n}(\phi(f)$.
Therefore
\beq
\phi(f)\in (\la_m)=I_m(\phi(A)):I_{m-1}(\phi(A))=\phi\big(I_m(A)\big):\phi\big(I_{m-1}(A)\big),
\eeq
equivalently:
$\phi(f\cdot I_{m-1}(A))\sseteq\phi(I_m(A))$.
As this holds for any $\phi$, we get: $\overline{f\cdot I_{m-1}(A)}\sseteq\overline{I_m}$. As this holds for any $f\in ann\big(T^1_{(\Si,G_{lr},A)}\big)$, we have:
$\overline{I_{m-1}(A)\cdot ann\big(T^1_{(\Si,G_{lr},A)}\big)}\sseteq\overline{I_m(A)}$.
Finally, by lemma \ref{Thm.Background.Integral.Closure.Lemma}
we have:
$\overline{ann\big(T^1_{(\Si,G_{lr},A)}\big)}\sseteq \overline{I_m(A)}:\overline{I_{m-1}(A)}$.
\epr
\beR
As one sees from the end of this proof, we have a stronger property: $\overline{ann\big(T^1_{(\Si,G_{lr},A)}\big)I_{m-1}(A)}\sseteq\overline{I_m(A)}$.
One would like to strengthen it further, to achieve from $\phi(f)\in\phi\big(I_m(A)\big):\phi\big(I_{m-1}(A)\big)$ that $\phi(f)\in\phi\big(I_m(A):I_{m-1}(A)\big)$. Then the final conclusion would be: $ann\big(T^1_{(\Mat,G_{lr},A)}\big)\sseteq \overline{I_m(A):I_{m-1}(A)}$.
But in general: $\phi(I:J)\neq\phi(I):\phi(J)$.
\eeR
\bex
Let $R$ be a regular, local, Noetherian ring and $A\in Mat_{m\times m}(R)$. Suppose that the ideal $I_m(A)=(det(A))$ is radical, i.e. $det(A)$ is square-free, and $height(I_{m-1}(A))>1$. (The later condition implies in particular: $dim(R)>1$.)
Then
\beq
ann.coker(A)=(det(A))=I_m(A)=\overline{I_m(A)}=\overline{I_m(A)}:\overline{I_{m-1}(A)}.
\eeq
Therefore
$ann(T^1_{(\Si,G_{lr},A)})=(det(A))$.
\eex
\bex
Suppose $R$ is a DVR. Then for any $A\in\Mat$ we have \beq
ann.coker(A)=I_m(A):I_{m-1}(A)=\overline{I_m(A)}:\overline{I_{m-1}(A)}.
\eeq
(For example, bring $A$ to the Smith normal form.) Therefore $ann(T^1_{(\Si,G_{lr},A)})=I_m(A):I_{m-1}(A)=ann.coker(A)$.
\eex
\subsubsection{Applications to finite determinacy}
\
{\em Proof of proposition \ref{Thm.Results.Fin.Det.for.GlGrGlr}.}
We use proposition \ref{Thm.Intro.ord.of.det.via.Loewy.length}. First we remark that the assumption \eqref{Eq.Intro.assumption.R.Noetherian.or.surjects} ensures the needed approximation property of $R$, see \S\ref{Sec.Background.Approximation.Properties}.
Indeed, in all the cases the equations $UA=A+B$, $AV^{-1}=A+B$, $UAV^{-1}=A+B$ are linear in the entries of $U,V$.
(In the later case present the equation in the form $UA=(A+B)V$.)
{\bf 1.} Note that $ann(T^1_{(\Si,G^{(m)}_r,A)})=ann\quotients{R^{\oplus m}}{\cm\cdot Im(A)}=
ann\quotients{R^{\oplus m}}{Im(A|_{\cm R^{\oplus n}})}$. Thus proposition \ref{Thm.Intro.ord.of.det.via.Loewy.length} gives the stated bound on $ord^\Si_{G_r}(A)$.
{\bf i.} If $dim(R)>(n-m+1)$ then for any $A\in Mat_{m\times n}(\cm)$ the ideal $ann.coker(A)$, having height at most $(n-m+1)$, see equation (\ref{Eq.Background.ann.coker.in.terms.of.Fitting.ideals}), cannot contain any power of the maximal ideal $\cm$.
This implies the statement.
{\bf ii.} Vice versa, the height of the ideal $ann.coker(A)$ is generically the expected one, see \S\ref{Sec.Background.Fitting.Ideals.ann.coker}.
Thus for any $A$ and the generic $B\in Mat_{m\times n}(\cm^N)$, the ideal $ann.coker(A+B)$
is of height $min\Big((n-m+1),dim(R)\Big)$. Thus if $dim(R)\le (n-m+1)$ then $ann.coker(A+B)$
contains a (finite) power of the maximal ideal.
{\bf 2.} Follows immediately from part 2 of theorem \ref{Thm.Results.Annihilators.for.GlGrGlr}.
{\bf 3.} Note that $height\big(\overline{I_m(A)}:\overline{I_{m-1}(A)}\big)=height(\overline{I_m(A)})=height(I_m(A))=height(ann.coker(A))$.
Thus the statement i. follows immediately from the first part and part 3 of theorem \ref{Thm.Results.Annihilators.for.GlGrGlr}.
{\bf ii.} and {\bf iii.}
Note that $G_{lr}\supset G_r$, thus $ord^\Si_{G_{lr}}(A)\le ord^\Si_{G_r}(A)\le ll_R(ann.coker(A|_{\cm R^p}))-1$, by part 1.
If $R$ is Noetherian then the lower bound on $ord^\Si_{G_{lr}}(A)$ is immediate.
To prove the lower bound in the non-Noetherian case we use the surjectivity of the completion, $R\stackrel{\hat{\psi}}{\twoheadrightarrow}\hR$.
Now by lemma \ref{Thm.Background.Completion.Lemma} (parts 2' and 1) we have the equality of Loewy lengths:
\beq
ll_R(ann(T^1_{(\Si,G_{lr},A)}))=ll_\hR\Big(\hat{\psi}(ann(T^1_{(\Si,G_{lr},A)}))\Big)=
ll_\hR\Big(ann\quotients{\hat{\psi}(T_{(\Si,A)})}{\hat{\psi}(T_{(G_{lr}A,A)})}\Big)=
ll_\hR\Big(\overline{I_m(\hA)}:\overline{I_{m-1}(\hA)}\Big).
\eeq
\epr
\bex
Consider the trivial case: $A$ is a ``numerical" matrix (with entries in $\k$), $m\le n$.
Then $ann.coker(A)$ is either $R$ or $\{0\}$. Hence
$A$ is finitely $G_{lr}$-determined
iff at least one of its maximal minors is a non-zero constant, i.e. $A$ is of the full rank.
(In other words, $A$ is invertible from the left, i.e. it has no left-kernel.) In this case, for $m\le n$,
$A$ is 0-determined \wrt $G_r$.
\eex
\bex
Let $R=\k[[x,y]]$ and $A=\bpm x^5&0&y^3\\0&y^4&x^3\epm$. Then $I_1(A)=(x^3,y^3)$, $I_2(A)=(y^7,x^5y^4,x^8)$
and $ann.coker(A)=(y^7,x^5y^4,x^8)$. (To compute $ann.coker$ one can notice that $height(I_2(A))=2$, as expected,
thus $ann.coker(A)=I_2(A)$, by equation \eqref{Eq.Background.ann.coker.in.terms.of.Fitting.ideals}.)
Therefore $\overline{I_1(A)}=\cm^3$ and $\overline{I_2(A)}=\cm^8+(y^7)$.
Thus
\beq
(y^7,x^5y^4,x^8)\sseteq ann T^1_{(\Si,G_{lr},A)}\sseteq \overline{I_2(A)}:\overline{I_1(A)}=\cm^5.
\eeq
Note that $ann.coker(A|_{\cm\cdot R^{\oplus 3}})=(y^7,x^5y^4,x^8)$.
Finally, $ll_R(\cm^5)=5$ and $ll_R(y^7,x^5y^4,x^8)=12$.
We get: $4\le ord^\Si_{G_{lr}}(A)\le 11$.
\eex
\bex\label{Thm.Finite.Determinacy.Glr.Low.Dimensional.Cases} Suppose $R$ satisfies the assumption \eqref{Eq.Intro.assumption.R.Noetherian.or.surjects} and $dim(R)>0$. Let $\Si=\Mat$.
1. If $m=n$ then $A\in Mat_{m\times m}(\cm)$ is finitely-$G_{lr}$-determined iff $dim(R)=1$, $det(A)\in R$ is not a
zero divisor and $det(A)\not\in\cm^\infty$. This generalizes Part 1 of \cite[Theorem 1.1]{Bruce-Tari04}.
Proof: for square matrices $height\Big(ann.coker(A)\Big)=height(det(A))\le 1$. Thus it can contain a power of the maximal ideal only when $dim(R)\le1$. Now invoke corollary \ref{Thm.Results.Fin.Det.for.GlGrGlr}.
2. Suppose $dim(R)=2$ and $m<n$. Suppose $A$ has at least two $m\times m$ blocks whose determinants are not zero divisors, neither belong to $\cm^\infty$ and moreover are relatively prime,
i.e. if $\De_i=a_i h\in R$ then $h\in R$ is invertible. Then A is finitely $G_r$-determined.
Proof: Let $\De_1,\De_2$ be two such minors then the height of the ideal $(\De_1)+(\De_2)$ is two.
Thus $height(ann.coker(A))=2$ and $ann.coker(A)$ contains a power of $\cm$.
\eex
\bex
Suppose $R$ satisfies the assumption \eqref{Eq.Intro.assumption.R.Noetherian.or.surjects} and $A\in\Mat$.
Then part 3 of theorem \ref{Thm.Results.Annihilators.for.GlGrGlr}
and theorem \ref{Thm.Background.Linearization} imply:
\beq
\text{if $\tA\equiv A\ mod(\cm\cdot ann.coker(A))$ then $\tA\stackrel{G_{lr}}{\sim}A$}.
\eeq
This both strengthens (quantifies) and generalizes theorem 5.2 of \cite{Cutkosky-Srinivasan}.
\eex
\subsubsection{Applications to the infinite determinacy}
Part 1 of proposition \ref{Thm.Results.Fin.Det.for.GlGrGlr} implies the following criterion.
\bcor\label{Thm.Proofs.Infinite.Determinacy.Gr}
Let $A\in Mat_{m\times n}(C^\infty(\R^p,0))$. If there exists an element $f\in I_m(A)$ satisfying {\L}ojasiewicz inequality, equation \eqref{Eq.Lojasiewicz.Ineq}, then $A$ is infinitely-$G_r$-determined.
\ecor
Indeed, by proposition \ref{Thm.Results.Fin.Det.for.GlGrGlr} it is enough to check that the ideal $I_m(A)$ contains $\cm^\infty$, in other words: any function flat at the origin is divisible by some element of $I_m(A)$. And this is ensured by {\L}ojasiewicz inequality.
\subsection{(Anti-)symmetric matrices and the congruence}\label{Sec.Proofs.Congruence}
\subsubsection{Computation of $ann(T^1_{(\Si,G_{congr},A)})$} \
{\em Proof of theorem \ref{Thm.Results.Annihilator.T1.for.Gcongr}.}
{\bf 1.} First we assume that $R$ is Noetherian.
To prove that $ann(T^1_{(\Si,G_{congr},A)})=\{0\}$ it is enough to show that this module is not a torsion module and is supported at the generic point
of any component of $Spec(R)$. Thus we assume that $Spec(R)$ is irreducible and localize at the generic point of $Spec(R)$. (Note that $dim(R)>0$.) At this
point we compare the modules $(T_{(GA,A)})_{(0)}$ and $(T_{(\Si,A)})_{(0)}$.
Note that $R_{(0)}$ is a field, thus we can bring the symmetric part of $A$ to the canonical form of proposition \ref{Thm.Background.Canonical.Form.Matrix.over.DVR}:
$A\stackrel{G_{congr}}{\sim} Diag+A_-$, where $Diag$ is diagonal while $A_-\in Mat^{skew-sym}_{m\times m}(R)$.
We can assume that $A+A^T$ is generically non-degenerate on $Spec(R)$, therefore $Diag$ is invertible over $R_{(0)}$.
To understand $(T_{(GA,A)})_{(0)}$, which is spanned by $UA+AU^T$, we re-scale the matrices, $\tU:= U\cdot Diag$, and expand them into the (skew-)symmetric parts, $\tU=\tU_++\tU_-$.
Then
$(T_{(GA,A)})_{(0)}=Span_R(2\tU_+,\tU A_-+A_-\tU^T)$. To prove that
$(T_{(GA,A)})_{(0)}\subsetneq(T_{(\Si,A)})_{(0)}=Mat^{sym}_{m\times m}(R_{(0)})\oplus Mat^{skew-sym}_{m\times m}(R_{(0)})$
we consider the map
\beq
Mat_{m\times m}(R_{(0)})\stackrel{\phi}{\to}Mat^{sym}_{m\times m}(R_{(0)})\oplus Mat^{skew-sym}_{m\times m}(R_{(0)}),\quad
\tU\to(2\tU_+,\tU A_-+A_-\tU^T)
\eeq
Note that the domain and the target are $R_{(0)}$-vector spaces of the same (finite) dimension. Thus it is enough to demonstrate
a non-trivial kernel of $\phi$. Indeed, the space $ker(\phi)=\{\tU_+=\zero,\ \tU_- A_-=A_-\tU_-\}$ is non-zero. It contains all the
matrices $\tU_-$ that commute with $A_-$, e.g.
$ker(\phi)\supseteq Span_R(A_-,A^3_-,A^5_-,\dots)$.
\
Now, if $R$ is non-Noetherian, but satisfies condition \eqref{Eq.Intro.assumption.R.Noetherian.or.surjects}, we use lemma \ref{Thm.Background.Completion.Lemma} to get: $\hat\psi(ann(T^1))=\{0\}$. Therefore $ann(T^1)\sseteq\cm^\infty$.
{\bf 2.} To prove that $ann(T^1_{(\Si,G_{congr},A)}) \supseteq I_m(A): I_{m-1}(A)$ we note that
$I_m(A): I_{m-1}(A)=ann.coker(A)$, see \S\ref{Sec.Background.Fitting.Ideals.ann.coker}.
Take any $B\in Mat^{sym}_{m\times m}(ann.coker(A))$. By the definition of the annihilator-of-cokernel there exists $U$ satisfying: $AU=\frac{B}{2}$. As $A$, $B$ are symmetric: $U^TA=\frac{B}{2}$. Therefore $UA+AU^T=B$.
But $UA+AU^T\in T_{(G_{congr}A,A)}$, see example \ref{Ex.Tangent.Spaces.to.the.Actions}.
Thus $ann.coker(A)\cdot T_{(\Si,A)}=Mat^{sym}_{m\times m}(ann.coker(A))\sseteq T_{(G_{congr}A,A)}$.
To prove that $ann(T^1_{(\Si,G_{congr},A)})\sseteq\overline{I_m(A)}:\overline{I_{m-1}(A)}$, let $f\in ann(T^1_{(\Si,G_{congr},A)})$, i.e. $f\cdot Mat^{sym}_{m\times m}(R)\sseteq Span_R(UA+AU^T|\ U\in Mat_{m\times m}(R)$. Then, for any $R\stackrel{\phi}{\to}S_{DVR}$:
\beq
\phi(f)\cdot Mat^{sym}_{m\times m}(S)\sseteq Span_R(U\phi(A)+\phi(A)U^T|\ U\in Mat_{m\times m}(S)).
\eeq
Now, bring $\phi(A)$ to the canonical form (using proposition \ref{Thm.Background.Canonical.Form.Matrix.over.DVR}), $\phi(A)=V\cdot D_{\phi(A)}V^T$, where $D_{\phi(A)}=\oplus\la_i\one$, with $(\la_1)\supseteq(\la_2)\supseteq\cdots\supseteq(\la_m)$.
Therefore:
\beq
\phi(f)\cdot Mat^{sym}_{m\times m}(S)\sseteq Span_R(UD_{\phi(A)}+D_{\phi(A)}U^T|\ U\in Mat_{m\times m}(S)).
\eeq
But the $(m,m)$-entry of any matrix of the form $(UD_{\phi(A)}+D_{\phi(A)}U^T)$ lies in the ideal
\beq
(\la_m)=ann.coker(\phi(A))=I_m(\phi(A)):I_{m-1}(\phi(A))=\phi(I_m(A)):\phi(I_{m-1}(A)).
\eeq
Therefore
$\phi(f\cdot I_{m-1}(A))\in \phi(I_m(A))$.
As this holds for any $R\stackrel{\phi}{\to}S_{DVR}$ we get:
$\overline{f\cdot I_{m-1}(A)}\sseteq\overline{I_m(A)}$. Finally, equation (\ref{Eq.Background.integral.closure.product.of.ideals}) implies
$f\in \overline{I_{m}(A)}:\overline{I_{m-1}(A)}$.
{\bf 3.} Let $f\in ann(T^1_{\Si,G_{congr},A})$, i.e. $f\cdot Mat^{skew-sym}_{m\times m}(R)\sseteq Span_R(UA+AU^T|\ U\in Mat_{m\times m}(R)$. Then, for any $R\stackrel{\phi}{\to}S_{DVR}$:
\beq
\phi(f)\cdot Mat^{skew-sym}_{m\times m}(S)\sseteq Span_R(U\phi(A)+\phi(A)U^T|\ U\in Mat_{m\times m}(S)).
\eeq
Bring $\phi(A)$ to the canonical form (using proposition \ref{Thm.Background.Canonical.Form.Matrix.over.DVR}), $\phi(A)=V\cdot E_{\phi(A)}V^T$, where $E_{\phi(A)}=\oplus_i \la_i E_i$
and $(\la_1)\supseteq(\la_2)\cdots\supseteq(\la_{\lfloor\frac{m}{2}\rfloor})$.
Therefore:
\beq
\phi(f)\cdot Mat^{skew-sym}_{m\times m}(S)\sseteq Span_R(UE_{\phi(A)}+E_{\phi(A)}U^T|\ U\in Mat_{m\times m}(S)).
\eeq
The $(m-1,m)$ and $(m,m-1)$ entries of $(UE_{\phi(A)}+E_{\phi(A)}U^T)$ belong to the ideal
$(\la_{\lfloor\frac{m}{2}\rfloor})$.
Therefore,
\beq\label{Eq.inside.proof}
\phi(f)\in (\la_{\lfloor\frac{m}{2}\rfloor})=I_{m-1}(\phi(A)):I_{m-2}(\phi(A))=\phi(I_{m-1}(A)):\phi(I_{m-2}(A)), \eeq
i.e. $\phi(f\cdot I_{m-2}(A))\sseteq\phi(I_{m-1}(A))$.
As this holds for any $R\stackrel{\phi}{\to}S_{DVR}$ we get:
$\overline{f\cdot I_{m-2}(A)}\sseteq\overline{I_{m-1}(A)}$. Finally, we get
$f\in \overline{I_{m-1}(A)}:\overline{I_{m-2}(A)}$, see equation (\ref{Eq.Background.integral.closure.product.of.ideals}).
Therefore $ann(T^1_{(\Si,G_{congr},A)})\sseteq \overline{I_{m-1}(A)}:\overline{I_{m-2}(A)}$.
\
{\bf 3.i.} If $m$ is even, then $\phi(I_{m}(A)):\phi(I_{m-1}(A))=\phi(I_{m-1}(A)):\phi(I_{m-2}(A))$. Thus
in equation (\ref{Eq.inside.proof}) we have: $\phi(f)\in \phi(I_{m}(A)):\phi(I_{m-1}(A))$ and therefore $ann(T^1_{(Mat_{m\times m}^{skew-sym}(R),G_{congr},A)})\sseteq \overline{I_{m}(A)}:\overline{I_{m-1}(A)}$.
The part $ann(T^1_{(Mat_{m\times m}^{skew-sym}(R),G_{congr},A)})\supseteq ann.coker(A)$ goes as in part 2.
{\bf 3.ii.} Suppose $m$ is odd, then by proposition \ref{Thm.Background.AntiSymmetric.Matrices.Pfaffians}:
$T_{(G_{congr}A,A)}\supseteq Mat^{skew-sym}_{m\times m}(Pf_{m-1}(A))$.
Thus
$ann(T^1_{(\Si,G_{congr},A)})\supseteq Pf_{m-1}(A)$.
\epr
\bex
In the low dimensional cases $T^1_{(\Si,G_{congr},A)}$ can be written down explicitly.
\\1. In the trivial case $A=\bpm 0&a\\-a&0\epm$ we get $Pf(A)=(a)\sset R$ and thus
$T^1_{(\Si,G_{congr},A)}\approx\quotients{R}{(a)}$.
\\2. For $3\times 3$ matrices take $A=\bpm 0&a&b\\-a&0&c\\-b&-c&0\epm$ we get $Pf_2(A)=(a)+(b)+(c)\sset R$ and thus
$T^1_{(\Si,G_{congr},A)}\approx\Big(\quotients{R}{(a)+(b)+(c)}\Big)^{\oplus 3}$.
\eex
\subsubsection{Finite determinacy for congruence}
\
{\em Proof} of Proposition \ref{Thm.Results.Finite.Determinacy.G_congr}.
The statements 1, 1', 1'' follow from theorem \ref{Thm.Results.Annihilator.T1.for.Gcongr} because for a square matrix $height(ann.coker(A))\le1$, while for a skew-symmetric matrix of odd-size $height(Pf_{m-1}(A))\le 3$, see \S\ref{Sec.Background.Fitting.Ideals.ann.coker}.
The statements 2.i. 2.ii. follow from theorem \ref{Thm.Results.Annihilator.T1.for.Gcongr} and proposition \ref{Thm.Intro.ord.of.det.via.Loewy.length}.
\bex
The natural question is the order of determinacy for a generic (skew-)symmetric matrix.
By proposition \ref{Thm.Results.Finite.Determinacy.G_congr} we must assume $dim(R)=1$.
For simplicity we assume that $R$ is a Henselian DVR.
\bee[i.]
\item Take the generic matrix $A\in Mat^{sym}_{m\times m}(\cm^k)$. Use proposition \ref{Thm.Background.Canonical.Form.Matrix.over.DVR} to
bring $A$ to the canonical form. (Note that the order of determinacy and $ann(T^1_{(\Si,G_{congr},A)})$ are invariant under the $G_{congr}$-action.)
Thus $A=\oplusl_j \la_j\one$ and by genericity the orders of all $\la_j$ are $k$. Thus the order of
determinacy is $k$. We get the rigidity in the following sense. Denote $\Si^{(k+1)}=\{A\}+Mat^{sym}_{m\times m}(\cm^{k+1})$,
then $T^1_{(\Si^{(k+1)},G_{congr},A)}=\{0\}$.
\item Similarly for skew-symmetric matrices of even size. Let $A\in Mat^{skew-sym}_{m\times m}(\cm^k)$ be generic, then
$A\stackrel{G_{congr}}{\sim}\oplusl_j \la_j E_j$, where $ord(\la_j)=k$. Thus the order of determinacy is $k$. If one takes
$\Si^{(k+1)}=\{A\}+Mat^{skew-sym}_{m\times m}(\cm^{k+1})$ then $T^1_{(\Si^{(k+1)},G_{congr},A)}=\{0\}$.
\eee\eex
\
Proposition \ref{Thm.Results.Finite.Determinacy.G_congr} implies the following criterion for the infinite determinacy of (skew-)symmetric forms
valued in the germs of smooth functions.
\bcor\label{Thm.Proofs.Infinite.Determinacy.G_congr}
1. Let either $A\in Mat^{sym}_{m\times m}(C^\infty(\R^p,0))$ or $A\in Mat^{skew-sym}_{m\times m}(C^\infty(\R^p,0))$, for $m$-even. If $det(A)$ satisfies the {\L}ojasiewicz inequality, equation \eqref{Eq.Lojasiewicz.Ineq},
then $A$ is infinitely-$G_{congr}$-determined.
\\2. Let $A\in Mat^{skew-sym}_{m\times m}(C^\infty(\R^p,0))$, for $m$-odd. If there exists an element $f\in I_{m-1}(A)$ that satisfies the {\L}ojasiewicz inequality, then $A$ is infinitely-$G_{congr}$-determined.
\ecor
The argument is the same as for corollary \ref{Thm.Proofs.Infinite.Determinacy.Gr}.
\subsection{The upper-block-triangular matrices and the action of $G^{up}_r$, $G^{up}_{lr}$}\label{Sec.Proofs.Upper.Triangular}
First we give two examples when the block-triangular matrices appear naturally.
\subsubsection{Morphisms of filtered free modules and their deformations}
Let $M$ be a free module and denote by $M_\bullet$ a filtration by free submodules which are direct summands, $\{0\}=M_0\ssetneq M_1\ssetneq \cdots\ssetneq M_k=M$.
Given two such filtered modules, $M_\bullet$, $N_\bullet$, consider their filtered homomorphisms,
$Hom_R(M_\bullet,N_\bullet)=\{\phi\in Hom_R(M,N)|\ \phi(M_i)\sseteq N_i\}$. Fix a filtered basis, i.e. a minimal set of generators
of $M$ compatible with the filtration:
\beq
(e_1,\dots,e_{i_1},e_{i_1+1},\dots,e_{i_2},\dots,e_{i_{k-1}+1},\dots,e_{i_k}), \quad \Big\{M_j=Span_R(e_1,\dots,e_{i_j})\Big\}_{j=1,\dots,k}.
\eeq
Similarly fix an $N_\bullet$-basis. For these chosen bases any homomorphism $\phi\in Hom_R(M_\bullet,N_\bullet)$ is presented by an upper-block-triangular matrix,
\beq
[\phi]=A=\bpm A_{11}&A_{12}&\dots\dots&\dots&A_{1k}\\\zero&A_{22}&\dots&\dots&A_{2k}\\\dots&\dots&\dots&\dots&\dots\\\zero&\zero&\dots&\zero&A_{kk}\epm.
\eeq
The change of bases in $M_\bullet$, $N_\bullet$ act on $A$ by elements of $G^{up}_{lr}=G^{up}_l\times G^{up}_r
=GL(M_\bullet,R)\times GL(N_\bullet,R)$.
Thus for the deformations of $\phi$ in $\Si=Hom_R(M_\bullet,N_\bullet)$ we have:
\beq
T^1_{(\Si,G^{up}_{lr},A)}= T^1_{\phi\in Hom_R(M_\bullet,N_\bullet)}=\quotient{T_{(Mat^{up},A)}}{T_{(G^{up}_{lr}A,A)}}.
\eeq
\subsubsection{Deformations of filtered submodules}
Fix a filtered module $\{0\}=M_0\ssetneq M_1\ssetneq \cdots\ssetneq M_k=M\sseteq R^{\oplus m}$, not necessarily free.
For each $M_i$ choose the minimal ``container", i.e. the minimal free direct summand $V_i$ satisfying: $R^{\oplus m}\supset V_i\supseteq M_i$.
(This ``container" module exists and is unique. Indeed, one begins with $M\sset R^{\oplus m}$ and notes that the intersection of two direct summands
is a direct summand. Moreover, the intersection process is finite as it occurs ``at $jet_0$ level". Namely, if $V,V'\sset R^{\oplus m}$ are
direct summands and $V\underset{R}{\otimes}\k=V'\underset{R}{\otimes}\k$ then $V=V'$.)
We assume $V_i\subsetneq V_{i+1}$. Choose a set
of generators of $R^{\oplus m}$ compatibly with the filtration $V_i\ssetneq V_{i+1}\ssetneq\dots R^{\oplus m}$.
Then the generating matrix $A_M$ of $M$ is upper-block-triangular:
\beq
A_M=\bpm A_{11}&A_{12}&\dots\dots&\dots&A_{1k}\\\zero&A_{22}&\dots&\dots&A_{2k}\\\dots&\dots&\dots&\dots&\dots\\\zero&\zero&\dots&\zero&A_{kk}\epm.
\eeq
A change of basis of $M_\bullet$ corresponds to the $G^{up}_r$-action. Therefore for the deformations of $M_\bullet$, inside the ambient filtration $V_\bullet$,
we have
\beq
T^1_{(\Si,G_r,A)}=T^1_{M_\bullet\sset V_\bullet}=\quotient{T_{(Mat^{up},A)}}{T_{(G^{up}_{r}A,A)}}.
\eeq
\subsubsection{}
{\em Proof of theorem \ref{Thm.Results.Annihilator.T1.Upper.Triang}.}
{\bf 1.} Denote $\Si=Mat^{up}_{\{m_i\}\times\{n_j\}}(R)$, thus $T_{(\Si,A)}\approx Mat^{up}_{\{m_i\}\times\{n_j\}}(R)$. We write down the tangent space to the group orbit:
\beq
T_{(G^{up}_rA,A)}=Span_R\Bigg(\bpm A_{11}U_{11}&A_{11}U_{12}+A_{12}U_{22}&\dots&\dots \suml^k_{i=1}A_{1i}U_{ik}\\
\zero&A_{22}U_{22}&\dots&\suml^k_{i=2}A_{2i}U_{ik}\\\dots&\dots&\dots&\dots&
\\
\zero&\dots&\zero&A_{kk}U_{kk}
\epm,\ \Big\{U_{ij}\in Mat_{m_i\times n_j}(R)\Big\}_{i\le j}\Bigg)
\eeq
Note that both $T_{(\Si,A)}$ and $T_{(G^{up}_rA,A)}$ decompose into the direct sums, according to the blocks of columns. Explicitly, $T_{(\Si,A)}=\oplusl^k_{l=1}\La_l$, where $\La_l$ consists of all matrices $B$ with the block-structure $B=\{B_{ij}\}$ satisfying: $B_{ij}=\zero$ if $j\neq l$ or if $i>j$.
Therefore, for a fixed $A$ we have:
$T^1_{(\Si,G^{up}_r,A)}=\oplusl^k_{q=1}(T^1_{(\Si,G^{up}_r,A)})_q$, where
\beq
(T^1_{(\Si,G^{up}_r,A)})_q=\frac{Mat_{(\suml^q_{j=1}m_j)\times n_q}(R)}{Span_R\Bigg(\bpm A_{11}&\dots&A_{1q}\\\zero&\dots&\\\zero&\dots&A_{qq}\epm\bpm U_{1q}\\\dots\\U_{qq}\epm,\ U_{jq}\in Mat_{m_j\times n_q}(R)\Bigg)}=
coker\bpm A_{11}&\dots&A_{1q}\\\zero&\dots&\\\zero&\dots&A_{qq}\epm.
\eeq
Accordingly $ann(T^1_{(\Si,G^{up}_r,A)})=\capl^k_{q=1}ann(T^1_{(\Si,G^{up}_r,A)})_q=\capl^k_{q=1}
ann.coker\bpm A_{11}&\dots&A_{1q}\\\zero&\dots&\\\zero&\dots&A_{qq}\epm=ann.coker(A)$. Here the last equality holds by part 4 of lemma \ref{Thm.Background.Ann.Coker.Properties}.
{\bf 1.i.} now is immediate. (For example, for each $l$ we have: $ann(T^1_{(\Si,G^{up}_r,A)})_q\sseteq ann.coker(A_{qq})$.)
{\bf 1.ii.} and {\bf 1.iii.} follow from equation \eqref{Eq.Background.ann.coker.in.terms.of.Fitting.ideals}.
{\bf 2.} The lower bound on $ann(T^1_{(\Si,G^{up}_{lr},A)})$ follows from part one because $G^{up}_{lr}\supset G^{up}_r$. For the upper bound we write down the tangent space
\beq\small
T_{(G^{up}_{lr}A,A)}=Span_R\Bigg(\bpm A_{11}U_{11}&\dots& \suml^k_{i=1}A_{1i}U_{ik}\\
\zero&\dots&\suml^k_{i=2}A_{2i}U_{ik}\\\dots&\dots&\dots
\\
\zero&\zero&A_{kk}U_{kk}\epm,
\bpm V_{11}A_{11}&\dots& \suml^k_{i=1}V_{1i}A_{ik}\\
\zero&\dots&\suml^k_{i=2}V_{2i}A_{ik}\\\dots&\dots&\dots
\\
\zero&\zero&V_{kk}A_{kk}\epm\Bigg)
\eeq
Thus $f\cdot T_{(\Si,A)}\sseteq T_{(G^{up}_{lr}A,A)}$ implies $Mat_{m_i\times n_i}((f))\sseteq Span_R(V_{ii}A_{ii}+A_{ii}U_{ii})$ for any $i$. But the later means:
\beq
f\in \capl_i ann(T^1_{(\Si_i,G_{lr},A_{ii})})\sseteq\capl_i \Big(\overline{I_{m_i}(A_{ii})}:\overline{I_{m_i-1}(A_{ii})}\Big).\text{\epr}
\eeq
\bex
\bee
\item Suppose $\{A_{ij}\}$ are mutually generic so that the ideals $ann.coker(A_{ii})$ are relatively prime. Then
we get the equality, $ann(T^1_{(\Si,G^{up}_r,A)})=\prodl^{k}_{i=1}ann.coker(A_{ii})$.
\item Suppose $A_{ij}=\zero$ for $1\le i<j\le k$. Then we have the obvious $ann(T^1_{(\Si,G^{up}_r,A)})=\capl^{k}_{i=1}ann.coker(A_{ii})$.
\eee
\eex
\subsubsection{Applications to finite determinacy}
Theorem \ref{Thm.Results.Annihilator.T1.Upper.Triang} implies the immediate
\bcor
\bee
\item If $dim(R)>(n_i-m_i+1)$ at least for some $i$, then no matrix $A\in Mat^{up}_{\{m_i\}\times\{n_j\}}(\cm)$ is finitely-$G^{up}_{lr}$-determined.
\item Suppose $R$ satisfies condition \eqref{Eq.Intro.assumption.R.Noetherian.or.surjects} and moreover $dim(R)\ge(n_i-m_i+1)$ for all $i$.
Then the generic finite-$G^{up}_{r}$-determinacy holds and
\begin{multline}
ll_R\Big(\capl_i ann.coker(A_{ii})\Big)-1\le ll_R(ann.coker(A))\le ord^{\Si}_{G^{up}_r}
\le\\\le ll_R(ann.coker(A|_{\cm\cdot R^{\oplus m}}))
\le
ll_R\Big(\prod_i ann.coker(A_{ii}|_{\cm\cdot R^{\oplus m}})\Big)-1.
\end{multline}
\eee
\ecor
\subsection{The case of conjugation}
{\em Proof of theorem \ref{Thm.Results.T1.Conjugation}} Denote $\Si=Mat_{m\times m}(R)$.
1. First we assume that $R$ is Noetherian.
As in the case of congruence, to prove that $ann(T^1_{(\Si,G_{conj},A)})=\{0\}$ it is enough to prove that $T^1_{(\Si,G_{conj},A)}$ is not a torsion on any component of $Spec(R)$. Thus we restrict to an irreducible component of $Spec(R)$ and check $T^1_{(\Si,G_{conj},A)}$ at the generic point of $Spec(R)$, algebraically we localize at the generic ideal $\{0\}$.
Therefore we compare the $R_{(0)}$-vector spaces:
\beq
Mat_{m\times m}(R_{(0)})\quad vs \quad (T_{(G_{conj}A,A)})_{(0)}=Span_{(0)}\Big(UA-AU|\ U\in Mat_{m\times m}(R_{(0)})\Big).
\eeq
Now the inequality $(T_{(G_{conj}A,A)})_{(0)}\ssetneq (T_{(\Si,A)})_{(0)}$ is immediate as $(T_{(G_{conj}A,A)})_{(0)}$ lies inside the hyperplane
$\{B|\ trace(B)=0\}\sset (T_{(\Si,A)})_{(0)}$.
Therefore $\Big(T^1_{(\Si,G_{conj},A)}\Big)_{(0)}\neq\{0\}$.
If $R$ is not Noetherian but satisfies condition \eqref{Eq.Intro.assumption.R.Noetherian.or.surjects}, then we take the completion $R\stackrel{\hat\psi}{\to}\hR$. We have $ann(\hat\psi(T^1_{(\Si,G_{conj},A)}))=\{0\}$, thus by lemma \ref{Thm.Background.Completion.Lemma} $\hat\psi(ann(T^1_{(\Si,G_{conj},A)}))=\{0\}$. Therefore
(as in the case of congruence) we get $ann(T^1_{(\Si,G_{conj},A)}))\sseteq\cm^\infty$.
2. Follows immediately from part one and proposition \ref{Thm.Intro.ord.of.det.via.Loewy.length}.
\epr
\subsection{Finite determinacy of chains of morphisms}\label{Sec.Proofs.Complexes}
Given a finite collection of free modules, $\{F_i\}_{i\in I}$, of ranks $\{m_i\}_{i\in I}$, and their morphisms, $\phi_i\in Hom_R(F_i,F_{i-1})$, we consider the corresponding chain $(F_\bullet,\phi_\bullet)$, not necessarily a complex. We assume $\{m_{i+1}\ge m_i\}$ and omit the terms $F_i$ with $m_i=0$.
The chains $(F_\bullet,\phi_\bullet)$, $(F_\bullet,\psi_\bullet)$ are called isomorphic if there exist automorphisms $U_i\circlearrowright F_i$ satisfying: $\psi_i\circ U_i=U_{i-1}\circ \phi_i$.
\bed
1. Denote by $\{J_{(F_i,\phi_i)}\}_{i\in I}$ the collection of the largest possible ideals satisfying:
\[\text{
if $\Big\{\psi_i\equiv\phi_i\ mod (J_{(F_i,\phi_i)})\Big\}_{i\in I}$, then the chains $(F_\bullet,\phi_\bullet)$ and $(F_\bullet,\psi_\bullet)$ are isomorphic.}
\]
2. Denote by $J_{(F_\bullet,\phi_\bullet)}$ the largest possible ideal satisfying:
\[
\text{if $\Big\{\psi_i\equiv\phi_i\ mod (J_{(F_\bullet,\phi_\bullet)})\Big\}_{i\in I}$, then the chains $(F_\bullet,\phi_\bullet)$ and $(F_\bullet,\psi_\bullet)$ are isomorphic.}
\]
\eed
\bel\label{Thm.Proofs.Lemma.Chains.Morphisms}
In both cases the largest object exists and $J_{(F_\bullet,\phi_\bullet)}=\capl_i J_{(F_i,\phi_i)}$.
\eel
\bpr
(We prove the existence in the first case, the second case is similar.)
Call a collection of ideals $\{J_i\}_{i\in I}$ admissible if the conditions $\{\psi_i\equiv \phi_i\ mod(J_i)\}_{i\in I}$ imply the isomorphism $(F_\bullet,\psi_\bullet)\approx (F_\bullet,\phi_\bullet)$.
The collection of zero ideals, $\{J_i=\{0\}\}_{i\in I}$ is obviously admissible.
Thus it suffices to prove: if $\{J_i\}_{i\in I}$, $\{\tJ_i\}_{i\in I}$ are two admissible collections then the collection $\{J_i+\tJ_i\}_{i\in I}$ is admissible.
Suppose $\{\psi_i=\phi_i+\De_i+\tilde{\De}_i\}_{i\in I}$, where the entries of $\De_i$ lie in $J_i$, while the entries of $\tilde{\De}_i$ lie in $\tJ_i$. Then there exists a collection of automorphisms, $\{U_i\in Aut_R(F_i)\}_{i\in I}$ satisfying: $\{U_i\psi_iU^{-1}_i=\phi_i+U_i\tilde{\De}_iU^{-1}_i\}_{i\in I}$. Now, the entries of $U_i\tilde{\De}_iU^{-1}_i$ still lie in the ideal $\tJ_i$. Therefore there exists a collection of automorphisms, $\{V_i\in Aut_R(F_i)\}_{i\in I}$ satisfying: $\{V_iU_i\psi_iU^{-1}_iV^{-1}_i=\phi_i\}_{i\in I}$. The composition $\{V_iU_i\}_{i\in I}$ provides the needed automorphism of the chains.
\
Finally, both inclusions $J_{(F_\bullet,\phi_\bullet)}\sseteq\capl_i J_{(F_i,\phi_i)}$ and $J_{(F_\bullet,\phi_\bullet)}\supseteq\capl_i J_{(F_i,\phi_i)}$ are obvious.
\epr
\bprop
Suppose $R$ is Noetherian.
\bee
\item $ann.coker(\phi_i|_{\cm\cdot F_i})\sseteq J_{(F_i,\phi_i)}\sseteq \overline{I_{m_i}(\phi_i)}:\overline{I_{m_i-1}(\phi_i)}$.
\item $\capl_i ann.coker(\phi_i|_{\cm\cdot F_i})\sseteq J_{(F_\bullet,\phi_\bullet)}\sseteq \Big(\capl_i\overline{I_{m_i}(\phi_i)}:\overline{I_{m_i-1}(\phi_i)}\Big)$.
\eee
\eprop
This strengthens (quantifies) and generalizes theorem 5.8 of \cite{Cutkosky-Srinivasan}.
\bpr
{\bf 1.} The isomorphism of the chains $(F_\bullet,\phi_\bullet)$, $(F_\bullet,\psi_\bullet)$ implies in particular the left-right equivalence of the representing matrices: $\phi_i\stackrel{G_{lr}}{\sim}\psi_i$ for each $i$. Thus the embedding $J_{(F_i,\phi_i)}\sseteq \overline{I_{m_i}(\phi_i)}:\overline{I_{m_i-1}(\phi_i)}$ follows immediately from part 3 of theorem \ref{Thm.Results.Annihilators.for.GlGrGlr} and theorem \ref{Thm.Background.Linearization}.
\
To prove the other inclusion observe that if for any $i$ holds:
$\phi_i \stackrel{G_r}{\sim}\psi_i$ then
$(F_\bullet,\phi_\bullet)\stackrel{ G_r}{\sim}(F_\bullet,\psi_\bullet)$.
Here $\prodl_i G_r$ acts on $\{F_i\}$ by a collection of automorphisms. Thus the statement follows directly from
part 1 of theorem \ref{Thm.Results.Annihilators.for.GlGrGlr} and theorem \ref{Thm.Background.Linearization}.
\
{\bf 2.} This now follows immediately from lemma \ref{Thm.Proofs.Lemma.Chains.Morphisms}.
\epr
\subsection{Criteria for relative determinacy}\label{Sec.Proofs.Relative.Determinacy}
We start from a preparatory result.
\bel
Fix two $R$-modules $N\sseteq M$ and an ideal $J\sset R$.
\bee
\item $ann\quotients{J\cdot M}{N}=ann\quotients{M}{N}:J$.
\item $ann\quotients{M}{N}\cdot ann\quotients{N}{J\cdot N}\sseteq ann\quotients{M}{J\cdot N} \sseteq ann\quotients{M}{N}\cap ann\quotients{N}{J\cdot N}$.
\eee
\eel
\bpr
1. The proof is just the chain of immediate equivalences:
\beq
\text{$\Big[f\in ann\quotients{M}{N}:J\Big]$ iff $\Big[f\cdot J\sseteq ann\quotients{M}{N}\Big]$ iff $\Big[f\cdot J\cdot M\sseteq N\Big]$ iff $f\in ann\quotients{J\cdot M}{N}$.}
\eeq
2. If $f\in ann\quotients{M}{N}$ and $g\in ann\quotients{N}{J\cdot N}$ then $gf\cdot M\sseteq g\cdot N\sseteq J\cdot N$, i.e. $gf\in ann\quotients{M}{J\cdot N}$.
If $f\in ann\quotients{M}{J\cdot N}$ then obviously $f\in ann\quotients{M}{N}$. As $N\sseteq M$ we get also: $f\cdot N\sseteq J\cdot N$, i.e. $f\in ann\quotients{N}{J\cdot N}$.
\epr
\bex
Both bounds in part 2 can be sharp.
\bee[i.]
\item If $dim(R)$ is large enough and $J$ is generic for the given $M,N$ then
$ann\quotients{M}{N}\cdot ann\quotients{N}{J\cdot N}=ann\quotients{M}{N}\cap ann\quotients{N}{J\cdot N}$, hence part two is an equality.
\item Let $R=\k[[x,y]]$, with $J=\cm^p$ and $N=(x^p,y^p)\sset M=R$. Then $J\cdot N=\cm^{2p}$,
$ann\quotients{N}{J\cdot N}=\cm^p$, $ann\quotients{M}{N}=(x^p,y^p)$. Therefore:
\beq
(x^p,y^p)= ann\quotients{M}{N}\cap ann\quotients{N}{J\cdot N}\supsetneq ann\quotients{M}{J\cdot N}=\cm^{2p}= ann\quotients{M}{N}\cdot ann\quotients{N}{J\cdot N}.
\eeq
\
\item Let $J=N=(f)\sset M=R$. Then
$ann\quotients{M}{N}\cdot ann\quotients{N}{J\cdot N}=ann\quotients{M}{J\cdot N}=(f^2)\ssetneq ann\quotients{M}{N}\cap ann\quotients{N}{J\cdot N}=(f)$
\eee
\eex
This lemma implies the immediate
\bcor\label{Thm.Proofs.Corollary.For.Rel.Determinacy}
Let $\Si^{(J)}$, $G^{(I)}$ be as in \S\ref{Sec.Results.Relative.Determinacy}.
Suppose $T_{(\Si^{(J)},A)}=J\cdot T_{(\Si,A)}$ and $T_{(G^{(I)}A,A)}=I\cdot T_{(GA,A)}$ . (This holds in all our examples.) Then:
\[
\Big(ann(T^1_{(\Si,G,A)})\cap ann\quotients{T_{(GA,A)}}{I\cdot T_{(GA,A)}}\Big):J
\supseteq ann(T^1_{(\Si^{(J)},G^{(I)},A)})\supseteq\Big(ann(T^1_{(\Si,G,A)})\cdot ann\quotients{T_{(GA,A)}}{I\cdot T_{(GA,A)}}\Big):J.
\]
\ecor
\bex
For $G=G_r$ and $I=R$ we get $ann(T^1_{(\Si^{(J)},G_r,A)})=ann.coker(A):J$.
If $R$ satisfies the condition \eqref{Eq.Intro.assumption.R.Noetherian.or.surjects}
and $J^q\sseteq ann.coker(A)$ then $A+Mat_{m\times n}(J^{q+k})\sseteq G^{(J^k)}_rA$, for any $k>0$.
This strengthens and generalizes \cite[Theorem 5.2]{Cutkosky-Srinivasan}: ``For $R$ a domain and complete \wrt $I_m(A_1)$-filtration, there exist $k_1,k_2$ such that if $A_1\equiv A_2\ mod(I_m(A_1))^j$ for $j\ge k_1$ then $A_1=UA_2V^{-1}$, with $U=\one\ mod(I_m(A_1))^{j-k_2}$, $V=\one\ mod(I_m(A_1))^{j-k_2}$."
\eex
\bex
Let $A\in Mat^{sym}_{m\times m}(\cm)$ where $R$ satisfies condition \eqref{Eq.Intro.assumption.R.Noetherian.or.surjects}
and is Henselian.
If $J^q\sseteq I_m(A):I_{m-1}(A)$ then $A+Mat^{sym}_{m\times m}(J^{q+k})\sseteq G^{(J^k)}_{congr}A$ for any $k>0$.
\eex
|
\section{Introduction\label{intro}}
The evolution of a galaxy is often strongly affected by the star-formation process. Numerous studies have characterized the relationship between star-formation activity, gas conditions, local environment, and galaxy morphology \citep[e.g.,][]{Kennicutt1998, Balogh2004, Kennicutt2012}. Volume-limited surveys \citep[e.g.,][]{Brinchmann2004, Lee2007, Bothwell2009} have mapped galaxy star-formation properties over stellar mass scales of $7\la$ log (M$_*$/M$_{\odot}$) $\ltsimeq11$. These surveys have found that the properties vary monotonically over a luminosity range $-19 < $M$_B < -15$ mag, but the distribution of star forming properties widens at fainter magnitudes roughly corresponding to a stellar mass of log (M$_*$/M$_{\odot}$) of 8.5. Specifically, galaxies at the lower mass ranges of these larger surveys show a broadening in their star-formation and gas properties, with less correlation between gas mass, stellar mass, and star-formation activity.
While these large surveys, as well as smaller, more detailed studies \citep[e.g., VLA-ANGST;][]{Ott2012}, have begun characterizing the star-formation and gas properties (and their dispersion) in lower mass galaxies, the faint end of the galaxy luminosity function remains largely unexplored. Outside of a dense group or cluster environment, these dwarf galaxies are expected to be gas rich, presumably evolving in relative quiescence due to inefficient star formation. Generally thought to have longer gas-cooling timescales and low, if not suppressed, star-formation activity, they play a key role in disentangling the factors driving star formation and star-formation driven galaxy evolution. Further highlighting the importance of low-mass galaxies, recent studies show that low-mass galaxies contribute significantly to both the total star-formation rate density at high redshift \citep[e.g.,][]{Alavi2014} and the ionizing background radiation \citep{Mostardi2013, Nestor2013}. While galaxies at the faint end of the luminosity function are predicted to be the most numerous class of galaxies, their intrinsic low luminosities and small angular sizes have made identifying and cataloging these systems through optical surveys difficult. Their high gas fractions make them better candidates for discovery through H{\sc i}\ surveys.
The Arecibo Legacy Fast ALFA (ALFALFA) survey \citep{Giovanelli2005, Haynes2011} is a blind extragalactic H{\sc i}\ survey to map the nearby H{\sc i}\ universe over 7000 deg$^2$ of high Galactic latitude sky. With an H{\sc i}\ mass detection limit as low as $10^6$ M$_{\odot}$\ for galaxies in the Local Group and $10^{9.5}$ M$_{\odot}$\ at the survey velocity limit of $z\sim0.06$, the ALFALFA survey was designed to populate the faint end of the H{\sc i}\ mass function over a cosmologically significant volume.
The Survey of H{\sc i}\ in Extremely Low mass Dwarfs \citep[SHIELD;][]{Cannon2011} was designed to examine the early release of the ALFALFA dataset in a systematic investigation of nearby galaxies with HI masses $\la 10^7$ M$_{\odot}$\ outside the Local Group. From the first $\sim10$\% of the processed ALFALFA survey data, 12 systems were selected for further study. $\textit{HST}$ optical imaging of these 12 systems have facilitated accurate distances based on the tip of the red giant branch (TRGB) method \citep{McQuinn2014}. The distances range from 5 to 12 Mpc; the H{\sc i}\ masses range from $4\times10^6-6\times10^7$ M$_{\odot}$\ based on these distances. All of the CMDs have red giant branch (RGB) sequences; a subset of the systems also show evidence of recent star formation based on their more populated upper main sequences (MS) and helium burning sequences \citep{Dohm-Palmer1997, McQuinn2011}. Most of the galaxies are detected in H$\alpha$ and spectroscopy of their H{\sc ii}\ regions reveal that the SHIELD galaxies are metal poor \citep{Haurberg2014}.
In addition to providing accurate distance measurements, resolved stellar populations provide a means to measure the star-forming properties of a system by reconstructing a star formation history (SFH) using stellar evolution isochrones and color-magnitude diagram (CMD) fitting techniques \citep[e.g.,][]{Tosi1989, Tolstoy1996, Gallart1996, Dolphin1997, Holtzman1999, Harris2001}. This approach has been used to measure the star-formation rate as a function of time (SFR(t)) in a significant number of nearby galaxies, putting constraints on star formation, star-formation feedback, and galaxy evolution \citep[e.g.,][]{Gallart2007, Dalcanton2009, McQuinn2010a}.
\begin{figure}
\epsscale{1.15}
\plotone{f1.eps}
\caption{$\textit{HST}$ composite image of AGC~110482, representative of the SHIELD sample, combining the F606W image (red), the average of F606W and F814W images (green), and the F814W image (blue). North is up and east is left. The field of view encompasses twice the optical major axis as determined by iteratively plotting CMDs with different elliptical parameters.}
\label{fig:image}
\end{figure}
Here, we use the optically resolved imaging of the SHIELD galaxies to measure the star-formation characteristics of gas-rich galaxies in a mass range that is not well probed in previous studies. In this work, we use the photometry from $\textit{HST}$ optical images measured in \citet{McQuinn2014} to derive the star-formation properties using a CMD fitting technique and stellar evolution isochrones. In Section~2, we describe the sample, observations, and data reduction. In Section~3, we explain the CMD fitting technique used to measure the star-formation activity. In Sections~4 and 5, we present the lifetime average SFRs, recent SFRs, and discuss whether the star-formation activity appears to be correlated with the local environment. In Section~6, we compare the star-formation and stellar mass properties with the H{\sc i}\ gas properties previously measured from \citet{Cannon2011} and compare the results in these low-mass galaxies to previously published works. Finally, in Section~7 we summarize our conclusions.
\section{The Galaxy Sample, Observations, and Photometry Method\label{data}}
Table~\ref{tab:galaxies} lists the 12 galaxies that comprise the SHIELD sample, their coordinates, observation details, and TRGB distances. The observations consist of optical imaging from the Advanced Camera for Surveys (ACS) Wide Field Channel \citep{Ford1998} aboard the $\textit{HST}$ in the F606W and F814W filters. New observations of 11 galaxies were obtained as part of the HST-GO-12658 SHIELD program (PI: Cannon). Archival observations were available for three galaxies from the previous HST-GO-10210 program (PI: Tully); two of these targets overlap with our new observations providing additional photometric depth to the final images. All observations were obtained with two filters during single $\textit{HST}$ orbits and were cosmic-ray split (CRSPLIT=2). Average integration times were 1000 s in the F606W filter and 1200 s in the F814W filter. The observations for the two galaxies that overlapped in the observing programs have longer final integration times of $\sim$1900 s in the F606W filter and $\sim$2300 s in the F814W filter. The observing programs and final integration times for each galaxy are noted in Table~\ref{tab:galaxies}.
\input{tab1}
Figure~\ref{fig:image} shows an example composite image from the data of AGC~110482, reproduced from the previous study measuring the distances to the galaxies \citep[][see their Figure~1]{McQuinn2014}. The images were cosmic-ray cleaned, processed by the standard ACS pipeline, corrected for charge transfer efficiency non-linearities, and combined using \textsc{MultiDrizzle}. The F606W filter image is shown in blue, F814W in red, and the average of the F606W and F814W images in green. Typical of the sample, and of gas-rich dwarf galaxies in general, the image shows both a young, blue stellar population and a red, older and somewhat more extended stellar population.
The sample has also been imaged in the narrowband H$\alpha$ filter from the WIYN 3.5 m telescope \citep{Haurberg2014}. These images reveal that 2 of the 12 galaxies lack H{\sc ii}\ regions. Specifically, AGC~749241 was not detected in H$\alpha$ and AGC~748778 had only weak, diffuse H$\alpha$ emission but no H{\sc ii}\ region. These H$\alpha$ data suggest that the star-formation activity over the most recent $\sim5$ Myr has been either below detection or very low (H$\alpha$-based SFR $< 10^{-5}$ M$_{\odot}$\ yr$^{-1}$). Thus, AGC~749241 and AGC~748778 meet the criteria of having both detectable amounts of H{\sc i}\ and no significant recent SFR as measured by H$\alpha$ and can be classified as transition dwarfs \citep[dTrans;][]{Mateo1998, Skillman2003, Weisz2011}. H$\alpha$-based SFRs for the rest of the sample will be presented in N.~C.~Haurberg (in preparation).
As described in detail in \citet{McQuinn2014}, photometry was performed on the pipeline processed, cosmic-ray cleaned images (CRJ files) using the ACS module of the DOLPHOT photometry package \citep{Dolphin2000}. Both quality and spatial cuts were performed on the photometry. Briefly, the photometry was filtered for well-fit point sources recovered with a signal-to-noise ratio $>5$ that were not severely affected by stellar crowding. Spatial cuts were determined iteratively by plotting the CMD of stars in concentric ellipses with the same ellipticity and position angles centered on the optical center of each galaxy. In the inner regions, the CMDs are dominated by stars from the galaxies, while in the outer regions, the CMDs are dominated by background sources. The semi-major and semi-minor axes of the ellipses were increased until the CMDs from larger annuli matched the distribution of point sources from a field region CMD. Artificial star tests using $\sim$500k stars were performed to measure the completeness limit of the images using the same photometry package and filtered on the same parameters.
Figure~\ref{fig:cmd} presents an example CMD for one of the farthest galaxies from the SHIELD sample, AGC~749237, plotted to the 50\% completeness level as determined by the artificial star tests, and one of the closest galaxies, AGC~111977. Representative photometric uncertainties per magnitude are shown for both. The photometry in the CMDs were corrected for Galactic absorption based on the dust maps of \citet{Schlegel1998} with recalibration from \citet{Schlafly2011}; these values are also noted in Table~\ref{tab:galaxies}. The depth of photometry shown in Figure~\ref{fig:cmd} brackets that of the overall sample. For the 5 galaxies that are located between $8-12$ Mpc, the depth of photometry is similar to AGC~749237, reaching approximately 1 mag below the TRGB, but populated by differing numbers of stars. For the 7 galaxies that are located between $5-8$ Mpc, the depth of photometry is similar to AGC~111977, reaching approximately 2 mag below the TRGB. CMDs of all 12 galaxies can be found in \citet[][see their Figure 2]{McQuinn2014}. In each CMD, the MS is identifiable as well as a populated RGB sequence. The presence of stars in the upper MS indicates that a system has experienced star formation at recent times (t$<$ few 100 Myr), while the presence of the RGB stars indicates star-formation activity at older times (t$>1$ Gyr). Some of CMDs also show somewhat populated helium burning sequences, including AGC~110482, AGC~111977, AGC~731457, AGC~749237, which is a further indication of recent star-formation activity. Asymptotic giant branch (AGB) populations, indicative of intermediate age star-formation activity, are also observed in a subset of the CMDs, most notably in AGC~110482, AGC~174605, and AGC~182595.
\begin{figure}
\epsscale{1.15}
\plotone{f2.eps}
\caption{CMDs of AGC~749237 (left) and AGC~111977 (right). Average uncertainties per magnitude are shown in each panel. The photometric depth of the CMD for AGC~749237 is typical of the 5 farthest galaxies, while the depth of the CMD for AGC~111977 is typical of the 7 closest galaxies. The number of stars in the CMDs across the sample varies between $\sim800$ and $\sim6500$.}
\label{fig:cmd}
\end{figure}
\section{Methodology for Deriving the Star-Formation Properties\label{sfhs}}
The star-formation characteristics of our sample of galaxies are reconstructed using stellar evolution isochrones to create a series of modeled CMDs from synthetic stellar populations of different ages and metallicities. The SFH of the best-fit modeled CMD to the observed CMD represents the most likely star-forming properties of the system. We use the numerical CMD fitting program, MATCH \citep{Dolphin2002}, with the stellar evolutionary models from \citet{Marigo2008} and updated AGB tracks from \citet{Girardi2010}. The primary inputs in reconstructing the past SFRs are the photometry and artificial star recovery fractions (i.e., quantitative observational uncertainties and incompleteness) coupled with the stellar evolutionary models. A Salpeter single sloped power law initial mass function with a spectral index of $-1.35$ from $0.1-120$~M$_{\odot}$\ \citep{Salpeter1955} is assumed as well as a binary fraction of $35\%$ with a flat secondary mass distribution. Distance is a free parameter fit by the SFH recovery program. We constrained the mean metallicity, $Z(t)$, to be a continuous, non-decreasing function with time. This physically motivated constraint guides the metallicity evolution in the absence of observational constraints that would be available with deeper photometric data and produces a more realistic metallicity evolution in the galaxies avoiding large jumps in metallicity over short time periods.
Photometric errors and extinction can broaden features in a CMD and are explicitly accounted for in the CMD fitting program. The photometric uncertainties and completeness are quantified using the artificial star tests. Extinction is a free parameter fit by the SFH recovery program. Thus, while the photometry shown in the CMDs in Figure~\ref{fig:cmd} has been corrected for foreground extinction, we used the uncorrected photometry in the CMD fitting program. Both foreground and internal extinction are expected to be low for this sample. The galaxies are located at high Galactic latitudes with foreground extinction estimated to range from $A_{F606W}$ of 0.04 to 0.26 mag (see Table~\ref{tab:galaxies}). Internal extinction is also expected to be low as dwarf galaxies have been shown to follow the mass-metallicity relation \citep{Berg2012}. Nonetheless, similar to the distance, extinction is an important parameter to fit in measuring the SFHs and provides a consistency check when compared to independently measured values. For our sample the distances and extinctions derived from the best-fit CMDs were within 0.1 mag and 0.05 mag, respectively, of the measured TRGB distances and foreground extinction estimates. Once the best fit has been verified, the final SFH solution is derived with the distance and extinction parameters fixed to the measured values from Table~\ref{tab:galaxies}.
Figure~\ref{fig:hess} presents two examples of the observed CMDs alongside the modeled CMD on the same axis scale for comparison as well as the residuals and weighted residuals between the observed and modeled CMD. As seen in Figure~\ref{fig:hess}, the models characterize all of the features of the CMDs quite well with only small discrepancies. The synthetic CMD of the galaxy is based on the most likely SFH and metallicity of the galaxy given our inputs and models \citep[e.g.,][]{Dolphin2003}. While the constraints on the metallicity evolution of the sample are limited by the photometric depth of the data, the best-fit values of the most-recent metallicities are in agreement with spectroscopic oxygen abundance measurements and the mass-metallicity relation for low-mass galaxies as discussed in \citet{Haurberg2014}, providing an additional consistency check on our results.
\begin{figure*}
\includegraphics[width=0.48\linewidth]{f3a.eps}
\includegraphics[width=0.48\linewidth]{f3b.eps}
\caption{$\textit{Left 4 panels:}$ AGC~749237: Hess diagrams of the observed CMD compared with the best-fit modeled CMD, residual of the data$-$model CMD with black and white points representing $\pm5\sigma$, and the residual significance CMD which weights the data$-$model by the variance in each Hess bin \citep{Dolphin2002}. The synthetic CMD reconstructs the different features seen in the observed CMD with minor discrepancies. The axes labels combine the name of the HST instrument (ACS WFC) with filter number. $\textit{Right 4 panels:}$ AGC~111977: Same sequence of hess diagrams as shown for AGC~749237.}
\label{fig:hess}
\end{figure*}
Uncertainties on the SFHs take into account systematic and random uncertainties. The systematic uncertainties measure the uncertainty in the SFHs due to variations of the real data from the theoretical stellar evolution models \citep{Dolphin2012}. These uncertainties are estimated by applying shifts to the models in both bolometric luminosity and temperature using Monte Carlo simulations. Random uncertainties were estimated using a new method of applying a hybrid Markov Chain Monte Carlo simulation as described by \citet{Dolphin2013}. This technique samples the probability distribution of the parameters used in the original solution to directly estimate confidence intervals. The inclusion of this additional source of uncertainty is particularly important for time bins where the SFR is low or zero, and for data whose photometric depth is above the horizontal branch or red clump as is the case for the current galaxy sample. The final uncertainties in the SFHs are calculated by adding the uncertainties in quadrature.
Reconstructing detailed SFHs (i.e., SFRs(t)) requires sufficient information on the ages and metallicity of the stellar populations from a CMD. Therefore, the photometric depth of the data is the main determinant on the temporal resolution achievable in a SFH at older ages. For a discussion on the systematic effects of photometric depth on measuring a SFH see \citet{McQuinn2010a}. A secondary determinant on the temporal resolution is the number of stars populating a CMD, which becomes a more significant factor to consider in low-mass galaxies. The small number of stars in the SHIELD galaxies compared to more massive dwarfs \citep{Dalcanton2009, McQuinn2010a} is therefore one of the dominant limitations in determining recent SFHs with high time resolution. Through extensive tests of MATCH using different time binning schemes, we found that measuring the SFRs over two timescales, recent (t$\sim$200 Myr; primarily measured by the upper part of an optical CMD) and lifetime-average SFRs (primarily measured from the RGB stars), resulted in the most robustly measured SFRs across the sample. The recent timescale of 200 Myr is an appropriate choice for the entire sample, regardless of distance and photometric depth, because of the generally low star-formation activity in low-mass galaxies. Further, the $\sim$200 Myr timescale for the recent SFRs is comparable to the timescale measured by near UV studies of star-formation rates, in particular from GALEX images \citep[e.g.,][]{Hao2011}. Thus, the recent SFR measurements reported here will enable comparisons with a large number of galaxies that have SFRs measured from integrated UV light.
\input{tab2}
Note that the data for a few of the galaxies allow only upper limits to be placed on the SFRs. For the lifetime SFRs, this is mainly due to limits in photometric depth and impacts two of the three farthest galaxies. For the more recent SFRs, this is a result of the low number of young stars available to fit in the CMD and impacts the two galaxies with the lowest recent SFRs. The recent and lifetime SFRs, as well as the uncertainties, are reported in Table~\ref{tab:sf}. For the galaxies whose measured SFRs are upper limits, we report only the upper bound of the uncertainties.
Stellar mass estimates can be derived from the lifetime-average SFRs. Integrating the lifetime-average SFR over the life of a galaxy yields the total stellar mass created in a system. This total stellar mass can be converted to a measurement of the current stellar mass by applying a correction for the amount of mass returned to a galaxy over the lifetime of the stellar populations. We used the recycling model from \citet{Kennicutt1994} who estimate that 30\% of the mass in a stellar generation is returned to a galaxy over its lifetime based on a Salpeter IMF. These values are reported in Table~\ref{tab:sf}. As a consistency check on our stellar mass values, we compared our results with preliminary estimates of the stellar masses based on mass-to-light (M/L) ratio analysis using \textit{Spitzer} Space Telescope IRAC imaging (J.~M.~Cannon et al., in preparation). Our stellar mass estimates are a factor of 2 higher than the majority of the stellar mass estimated from the M/L ratios, but are still in general agreement with one another given the large uncertainties. The exception is AGC~748778 whose stellar mass estimate from the M/L ratio is significantly lower than our value. Future work on the SHIELD sample will include a larger discussion of the two stellar mass estimates including a description of the infrared imaging, analysis, and uncertainties (J.~M.~Cannon et al., in preparation).
\begin{figure}
\epsscale{1.15}
\plotone{f4.eps}
\caption{Recent (t$<200$ Myr) SFRs compared with average lifetime SFRs for the SHIELD sample. To provide further context, the scale on the top x-axis is stellar mass based on the lifetime average SFRs and the age of the Universe and assuming a 30\% recycling fraction. Arrows denote cases where upper limits were placed on the SFRs. The solid line represents equivalent recent and lifetime SFRs. Overall, the recent SFRs are comparable to, or slightly higher than, the lifetime average SFRs. The birthrate parameters fall in a narrow range with a mean of 1.4 and a dispersion of 0.7. The level of recent star-formation activity in each system does not appear to be correlated with distance from the nearest known galactic neighbors or local environments.}
\label{fig:sfrs_stellar_mass}
\end{figure}
\begin{figure}
\epsscale{1.15}
\plotone{f5.eps}
\caption{An expanded view of Figure~\ref{fig:sfrs_stellar_mass} with results from 3 different surveys over-plotted. From \citet{Ott2012}, typical uncertainties for the VLA-ANGST recent SFRs are of order $\pm0.2$ dex; typical uncertainties for stellar mass measurements were not provided. From \citet{Huang2012}, typical uncertainties are of order $\pm0.3$ dex for the recent SFRs, and $\pm0.2$ dex for the stellar mass measurements. The shaded region for the SDSS survey covers the general range of properties that overlap with the other surveys plotted. The recent and lifetime star-formation measurements are consistent with those measured by these three surveys in the overlapping mass range. Further, the dTrans are indistinguishable from the dIrr galaxies. These low-mass galaxies follow the general trends that lower-mass galaxies have lower SFRs.}
\label{fig:sfrs_angst}
\end{figure}
\section{Star-Formation Rates and Comparison with Other Surveys\label{sfrs}}
Figure~\ref{fig:sfrs_stellar_mass} presents a comparison between the recent (t$<200$ Myr) SFRs with the average lifetime SFRs for the SHIELD sample. dIrrs are plotted as filled, blue cirlces; dTrans are plotted as unfilled, black cirlces. The recent and lifetime SFRs are low, ranging from $-3.5 \la$ log (SFR) $\la -1.8$, with SFRs in units of M$_{\odot}$\ yr$^{-1}$. The top axis in Figure~\ref{fig:sfrs_stellar_mass} shows the stellar mass for each system, which provides a complementary view of the results. Seen in this way, it is clear that the galaxies follow the well-established correlation that higher mass galaxies have higher recent SFRs. This trend reflects that while gas fractions tend to decline with increasing galaxy mass, total gas masses increase as a function of mass and can thus support higher SFRs \citep[e.g.,][]{Boselli2001}.
In Figure~\ref{fig:sfrs_stellar_mass}, the solid line marks where the recent to lifetime SFR ratio \citep[i.e., the birthrate parameter $b \equiv SFR_{recent}/SFR_{lifetime}$;][]{Scalo1986} is equal to 1. The birthrate parameters measured for the sample are listed in Table~\ref{tab:sf}. The range in the b parameter is narrow with a mean value of 1.4 and a dispersion of 0.7. To test whether the dispersion is due to measurement errors or to intrinsic scatter in star-formation activity, we fit a linear relationship to the recent and lifetime SFRs using the IDL mpfitexy routine \citep{Williams2010}. The mpfitexy routine depends on the mpfit package \citep{Markwardt2009}. We find that less than 35\% of the dispersion is due to intrinsic scatter while greater than 65\% is due to the measured uncertainties. Overall the SHIELD sample shows recent star-formation activity that is comparable to, or slightly higher than, the lifetime average SFR. Further, none of the galaxies show star-formation activity that is significantly lower than their lifetime average (i.e., no quenched star formation), as might be expected from their relatively low-density environments \citep{Hodge1971, Einasto1974, vandenBergh1994a, vandenBergh1994b, Mateo1998, McConnachie2012, Geha2012}.
Figure~\ref{fig:sfrs_angst} expands the plot from Figure~\ref{fig:sfrs_stellar_mass} to include results from three other surveys. We add the recent SFRs and stellar masses from both the VLA-ANGST sample \citep{Ott2012} and the sub-sample of gas-rich ALFALFA dwarf galaxies from \citet{Huang2012}, and highlight the general range of properties measured in the SDSS survey from \citet{Brinchmann2004} that overlaps with the parameter space in Figure~\ref{fig:sfrs_angst}. All stellar masses were normalized to a Salpeter IMF to facilitate the comparison. Note however, that the stellar masses from these surveys were derived using very different techniques including mass-to-light ratios \citep{Ott2012}, stellar absorption-line indices \citep{Brinchmann2004}, and spectral energy distribution (SED) fitting \citep{Huang2012}. In addition to the $\sim50$\% uncertainties in stellar mass from these various techniques, there are also systematic uncertainties between techniques that are not considered. However, these uncertainties are less than the broad parameter space probed in Figure~\ref{fig:sfrs_angst}, making the comparison between the surveys of value. The recent SFRs have also been measured using different techniques. The SFRs from \citet{Brinchmann2004} were based on emission line model fits and integrated H$\alpha$ emission measurements scaled to a SFR. The SFRs from \citet{Ott2012} were derived by scaling integrated ultraviolet luminosities. These UV-based SFRs benefit from tracing the star-formation activity on comparable timescale to our CMD-based recent SFRs. The SFRs in \citet{Huang2012} were derived from SED fitting and were shown to agree reasonably well with SFRs derived by scaling integrated ultraviolet luminosities in systems with low extinction.
\begin{figure}
\epsscale{1.15}
\plotone{f6.eps}
\caption{The birthrate parameter is plotted against the distance to the nearest known neighbor from \citet{McQuinn2014}. While the overall range in the b parameter is rather narrow, the galaxies with the closest known neighboring system have the lowest values of the birthrate parameter, while the galaxies that are located at larger distances from their nearest neighbor have higher values of the birthrate parameter, albeit with large uncertainties.}
\label{fig:neighbor_b}
\end{figure}
There are a number of correlations to note in Figure~\ref{fig:sfrs_angst}. First, while the dTrans (unfilled black points) are at the low-mass end of the SHIELD sample, their star-formation properties are consistent with the rest of the SHIELD sample and are indistinguishable from the properties of the dIrrs galaxies measured from the other surveys. Second, the trend that gas-rich low-mass galaxies generally have higher recent SFRs compared with lifetime average SFRs (i.e., b $\ga$ 1) is more apparent across the combined samples. Third, while the properties of the SHIELD galaxies are consistent with these other studies, the SHIELD galaxies have a narrower range in birthrate parameters, particularly compared to the ALFALFA dwarfs from \citet{Huang2012}. One possible explanation is that the broader range in the birthrate parameter in \citet{Huang2012} is due to higher uncertainties in comparing the stellar masses and SFRs by the aforementioned different techniques. On the other hand, the SHIELD sample is small; future work will expand the SHIELD sample to $\sim$30 galaxies and may help determine whether the narrow range in recent-to-lifetime SFRs is typical in this very low-mass regime (K.~B.~W.~McQuinn et al. in preparation). Note that the opposite trend (i.e., lower recent to historical average SFRs) was found in \citet{Skillman2003} for a sample of gas-rich dwarf galaxies in the Sculptor group. However, the SFRs were based on integrated H$\alpha$ emission measurements which may not accurately represent the recent star-formation activity. SFRs based on H$\alpha$ emission trace only the most recent ($\sim$5 Myr) star-formation activity and depend both on constant levels of star formation and on a fully-populated upper-end of the IMF. In this low SFR regime, not only can the star formation fluctuate on short timescales, but the IMF can be stochastically populated \citep{Boselli2009, Lee2009, Goddard2010, Koda2012}. Thus, the H$\alpha$-based SFRs can be an unreliable tracer of star-formation activity, introducing scatter in the measurements and potentially biasing SFR measurements low in dwarf galaxies.
\section{Recent SFRs: Influenced by Environment or Regulated Internally?\label{recent_sfrs}}
One of the outstanding questions concerning the evolution of relatively isolated, low-mass galaxies is the comparative importance of environmentally driven star-formation activity versus star-formation activity regulated by internal conditions within a system. The former is directly correlated with the number of, and proximity to, neighboring systems. A full investigation of the role environment plays requires a complete census of a galaxy's environment including positional and velocity information of neighboring systems and SFHs with fine temporal resolution. However, to first order, by comparing the recent SFRs of galaxies derived here with the proximity of their known neighboring systems determined from our previous study \citep[][see their Figures~5 and 6]{McQuinn2014}, we can probe whether there is an obvious environmental factor influencing their recent evolution.
In Figure~\ref{fig:neighbor_b} we present a comparison of the b values for the sample with the distances to the nearest known neighbors from \citet{McQuinn2014}. All six galaxies whose recent star-formation activity is somewhat higher than their lifetime average lie in low-density environments. Specifically, AGC~748778, AGC~174605, and AGC~749237 are truly isolated with no known neighbors identified within a 1 Mpc radius. AGC~174585 and AGC~731457 are both located $\sim0.9$ Mpc from their nearest known neighbors, namely the $14+19$ association and the low-mass galaxy DDO~83 respectively. AGC~111946 is located at the end of a linear structure comprised of two known galaxy associations, the NGC~784 and NGC~672 groups \citep{Zitrin2008, McQuinn2014}. While AGC~111946 appears to be part of this elongated 4 Mpc structure, it is separated by a distance of $\sim1$ Mpc from the more tightly clustered galaxies in the NGC~672 group. Overall, the isolated nature of these six galaxies suggests that the recent star-formation activity is not driven by gravitational interactions. Note however that we cannot rule out the possibility of possible undetected gas-poor companions outside of the HST fields of view.
Further, the three galaxies that show somewhat lower levels of recent star-formation activity compared with their lifetime averages (AGC~111164, AGC~111977, and AGC~112521) are part of the galaxy associations of the NGC~784 and NGC~672 groups mentioned above. This linear structure of galaxies stretches $\sim4$ Mpc and is thought to trace the local matter distribution. In contrast to the location of the aforementioned system AGC~111946, these three SHIELD systems lie at closer distances to the members of these groups. The nearest neighbor to AGC~111164 is the more massive dwarf starburst galaxy NGC~784 at a distance of $\sim0.14$ Mpc. The SFR in NGC~784 has been noted to be elevated over the last few 100 Myr \citep{McQuinn2010b}. However, despite being separated by only $\sim0.14$ Mpc, the recent SFR in AGC~111164 is half its lifetime average. AGC~111977 is located $\sim0.65$ Mpc from AGC~112521, which is also separated by a comparable distance in the opposite direction from four members of the NGC~672 group, namely NGC~672, IC~1727, AGC~111945, and the SHIELD galaxy AGC~110482. Even though these galaxies reside in a more populated environment, their recent star-formation activity is lower than their lifetime values. Note that the H{\sc i}\ masses for these galaxies are on the low end of our sample ($4\times10^6$ M$_{\odot}$\ for AGC~111164 and $7\times10^6$ M$_{\odot}$\ for AGC~111977 and AGC~112521), which may provide a partial explanation for the low recent star-formation activity.
Finally, we examine the three galaxies whose recent star-formation activity is approximately equivalent to their lifetime averages, namely AGC~110482, AGC~182595, and AGC~749241. The first system, AGC~1110482, is located in the galaxy association described above, $\gtsimeq0.65$ Mpc from other members of the NGC~672 group. AGC~182595 is truly isolated with no known neighbors with 1 Mpc. AGC~749241 lies in a more populated region within 9 other galaxies which form a linear structure that extends 1.6 Mpc from end to end. Given the distribution of galaxies in this region, it is likely that AGC~794241 has been influenced by the external gravitational perturbations from other systems. This is supported by the irregular H{\sc i}\ morphology whose centroid is not co-spatial with ground-based optical imaging of the stellar population \citep[][see their Figure 2]{Cannon2011}. However, the relatively low-level of recent star-formation activity in AGC~749241 indicates that any such gravitational interaction has not had a dramatic impact on the star forming properties over the past 200 Myr.
\begin{figure}
\epsscale{1.15}
\plotone{f7.eps}
\caption{Specific SFRs as a function of stellar mass. In contrast to the star-forming properties in more massive galaxies, there is a high degree of dispersion in these very low-mass galaxies. This dispersion likely reflects small fluctuations in star-formation activity superposed on already low average SFRs.}
\label{fig:ssfr_stellar_mass}
\end{figure}
\begin{figure}
\epsscale{1.15}
\plotone{f8.eps}
\caption{An expanded view of Figure~\ref{fig:ssfr_stellar_mass} with results from 3 different surveys over-plotted. Uncertainties for the SSFR measurements are greater than $\pm0.2$ dex for VLA-ANGST and $\pm0.3$ dex for ALFALFA; uncertainties for the stellar mass were not provided for the VLA-ANGST measurements and are of order $\pm0.2$ dex for ALFALFA. The shaded region from \citet{Bothwell2009} covers the range of points included in their study; the extension to lower SSFRs includes both measurements and upper limits. Overall, for low-mass galaxies none of the samples show a high degree of correlation between the SSFRs and stellar mass.}
\label{fig:ssfr_angst}
\end{figure}
In summary, while the range in b parameters is fairly small, the somewhat higher levels of recent star-formation activity in these low-density galaxy environments do not correlate with smaller distances to the nearest neighboring galaxies, or vice versa, suggesting that the recent activity is governed by internal conditions and not environmentally driven. Similarly, \citet{Hunter2004} did not find a correlation between star-formation activity and distances to nearest neighbors in a larger sample of gas-rich dwarf galaxies. However, these authors note the possibility that this could be a selection effect as there could be undetected low-mass neighbors or H{\sc i}\ clouds near the galaxies with higher measured recent SFRs. While we cannot rule out the possibility of undetected neighboring systems, particularly if they are gas-poor, there are no systems detected nearby in position or velocity in the ALFALFA catalog \citep{Haynes2011}. Further, no nearby systems have been found in the VLA data of the sample \citep{Cannon2011}, nor in the HST images which cover a field of view extending to $\sim0.5$ Mpc for the farther systems \citep{McQuinn2014}.
For the galaxies with higher recent SFRs, an alternative explanation to star formation driven by gravitational perturbations may be that these low-mass galaxies have experienced a delayed onset to star formation, similar to 3 low-mass galaxies found in the Local Group (Leo~A, Leo~T, and Aquarius). In a detailed study of the ancient star formation history of Leo~A, \citet{Cole2007} found that over 90\% of the stars in the galaxy were formed over the most recent 8 Gyr. In a similar study of Leo~T, \citet{Weisz2012} found $\sim50$\% of the stars in the galaxy were formed over the same timescale. The Aquarius dIrr also shows delayed onset of star-formation with 90\% of the stars forming in the last 10 Gyr \citep{Cole2014}. While a limited comparison based only on three galaxies, these examples of delayed or suppressed star formation in low-mass galaxies lend support to the idea of galaxy downsizing. Higher recent star-formation activity in low-mass galaxies may not always be tied to gravitational interactions, but may instead be the result of a combination of internal properties and conditions governed by, for example, gas cooling timescales or reionization, and stochastic phenomena such as mergers.
\section{Comparison of Star-Formation with Global Gas Properties\label{properties}}
The recent star-formation activity of the sample can also be analyzed by normalizing the SFR by the stellar mass (M$_*$) of each galaxy. Figure~\ref{fig:ssfr_stellar_mass} shows the specific star-formation rates (SSFR $\equiv$ recent SFR / M$_*$) as a function of stellar mass. There is no apparent correlation between the SSFR and the stellar mass of the galaxies. This is in agreement with previous surveys of low-mass galaxies which have noted a high degree of scatter in the SSFRs as a function of stellar mass in systems with SFRs $<10^{-3}$ M$_{\odot}$~yr$^{-1}$ \citep[e.g.,][]{Bothwell2009, Huang2012}.
Figure~\ref{fig:ssfr_angst} expands on Figure~\ref{fig:ssfr_stellar_mass} by adding results from the VLA-ANGST survey \citep{Ott2012} and a different sub-sample of gas-rich dwarf galaxies from the ALFALFA survey \citep{Huang2012}, as well as adding the range of galaxy properties measured in the survey by \citet{Bothwell2009}. The scatter is the greatest in the study by \citet{Bothwell2009}, particularly below M$_* < 10^7$ M$_{\odot}$, which includes a number of upper limits on galaxies with no detectable star formation. Similarly to the SFRs from \citet{Skillman2003}, the SFRs in \citet{Bothwell2009} were based on integrated H$\alpha$ emission which can be unreliable in the low-SFR regime. As a result, \citet{Bothwell2009} was unable to determine whether the scatter in the SSFR-stellar mass relation was intrinsic, or due to the H$\alpha$ emission not adequately tracing the SFR. \citet{Huang2012} also find a large scatter in SSFRs for low-luminosity galaxies, including a similar quenching of star formation for some galaxies with $M_* < 10^8$ M$_{\odot}$. However, since the SFRs in this study were based on SED fitting, \citet{Huang2012} conclude that the range in SSFR, including the departure to low SSFRs for some low-mass galaxies, is real. In comparison, the SHIELD and VLA-ANGST samples show a narrower scatter in the SSFR-stellar mass relationship, and none of the galaxies in the SHIELD sample show quenched star formation (i.e., none have outlying low values of SSFR).
\begin{figure}
\epsscale{1.15}
\plotone{f9.eps}
\caption{Fraction of H{\sc i}\ to stellar mass vs. stellar mass for the 12 SHIELD galaxies. The stellar mass was determined from the lifetime average SFRs of the galaxies as described in the text. The H{\sc i}\ mass was calculated from the H{\sc i}\ fluxes reported in \citet{Cannon2011}. The range in gas fractions is consistent with more massive gas-rich dIrrs.}
\label{fig:gas_frac}
\end{figure}
\begin{figure}
\epsscale{1.15}
\plotone{f10.eps}
\caption{An expanded view of Figure~\ref{fig:gas_frac} with results from the FIGGS sample over-plotted. From \citet{Begum2008}, uncertainties on $M_{HI}$ are of order 10\%, while uncertainties on the stellar masses, which are estimated from mass-to-light ratios for the sample, are not well-quantified. The SHIELD sample overlaps with the FIGGS galaxies that have both lower H{\sc i}\ masses $\textit{and}$ lower stellar masses.}
\label{fig:gas_frac_figgs}
\end{figure}
Figure~\ref{fig:gas_frac} presents a comparison of the fraction of H{\sc i}\ gas mass to stellar mass (M$_{HI} / $M$_*$) as a function of stellar mass. These gas fractions were calculated using the stellar masses derived from the lifetime SFRs and the H{\sc i}\ masses based on H{\sc i}\ flux measurements from \citet{Cannon2011}. The H{\sc i}\ masses and gas fractions are listed in Table~\ref{tab:sf}. The gas properties show a broad distribution as a function of stellar mass with a general trend that gas fractions decrease at higher stellar masses. The trend and range in gas fraction agrees with those previously noted for low-mass galaxies \citep[e.g.,][]{vanZee1997, Schombert2001}. Note that while the fraction of H{\sc i}\ mass relative to stellar mass decreases as a function of stellar mass, more massive galaxies in our sample still have overall higher H{\sc i}\ masses than do their less massive counterparts. It is this overall larger reservoir of fuel that helps explain the correlation between increasing SFRs with increasing stellar mass in gas-rich systems seen Figure~\ref{fig:sfrs_angst}. Figure~\ref{fig:gas_frac_figgs} expands Figure~\ref{fig:gas_frac} to include results from the Faint Irregular Galaxies GMRT Survey \citep[FIGGS;][]{Begum2008}. We use the H{\sc i}\ masses from \citet{Begum2008} and estimate the stellar masses using the B-band luminosities of the galaxies and assuming a mass-to-light ratio of one. The SHIELD galaxies overlap with the low-mass end of the FIGGS systems (i.e., systems with both lower H{\sc i}\ masses $\textit{and}$ lower stellar masses).
Another way of comparing the SFR and gas content is to normalize both parameters by stellar mass. In Figure~\ref{fig:ssfr_gas_ratio}, we present a comparison of the SSFR with the gas fraction (from above, M$_{HI}$ / M$_*$). Consistent with the results presented in Figure~\ref{fig:ssfr_stellar_mass}, the star-formation activity for the SHIELD sample shows a high degree of dispersion. Figure~\ref{fig:gas_ratio_angst} expands the plot from Figure~\ref{fig:ssfr_gas_ratio} to include the results from the VLA-ANGST survey from \citet{Ott2012} and a sub-sample of the ALFALFA survey with overlapping mass ranges from \citet{Huang2012}. The SHIELD sample lies at the low end of the gas ratio ranges probed by these other surveys. This may be a selection effect as the SHIELD sample was chosen based on low H{\sc i}\ masses, and hence may preferentially probe the lower range of gas and stellar mass properties. For gas fractions that overlap with the SHIELD sample, the galaxies from the other surveys span a larger range in SSFRs. This is simply a restatement of our previous finding that the SHIELD galaxies have a narrow range in birthrate parameter. The broadening in SSFR is in contrast with the tight correlation between SSFRs and gas fractions reported for more massive galaxies \citep{Brinchmann2004, Lee2007} and suggests that the stellar mass is not the dominant factor in regulating star formation in low-mass galaxies. Instead, results in this low-mass regime indicate a star-formation process that is dependent upon both the overall gas mass and gas fraction. The star-formation variability in gas-rich low-mass systems is sensitive not only to the overall galaxy mass, but also the star formation history of the galaxy due to intrinsically smaller gas reservoirs.
\begin{figure}
\epsscale{1.15}
\plotone{f11.eps}
\caption{Specific SFR as a function of the gas ratios (M$_{HI}$ / M$_*$) for the sample. Similar to Figure~\ref{fig:ssfr_stellar_mass}, there is little correlation between the gas content and star-formation activity once both quantities have been normalized by stellar mass. Despite the scatter, no galaxies show significantly lower levels of SSFRs for a given gas ratio.}
\label{fig:ssfr_gas_ratio}
\end{figure}
\begin{figure}
\epsscale{1.15}
\plotone{f12.eps}
\caption{An expanded view of Figure~\ref{fig:ssfr_gas_ratio} with results from 2 different surveys over-plotted. Uncertainties for the SSFR measurements are greater than $\pm0.2$ dex for VLA-ANGST and $\pm0.3$ dex for ALFALFA; uncertainties for the gas ratio are not quantified for VLA-ANGST and of order $\pm0.1$ dex for ALFALFA. There is a high degree of dispersion between the SFR and the H{\sc i}\ masses when both are normalized by stellar mass. The SHIELD sample overlaps with both surveys at the lower end of their gas ratios. This is not unexpected as the SHIELD sample was selected, in part, based on a criterion of low H{\sc i}\ masses.}
\label{fig:gas_ratio_angst}
\end{figure}
Assuming the star formation in each system persists at the recent levels of activity, the gas content in each system can be used to estimate how long a system can continue forming stars. This gas consumption timescale \citep[$\tau_{gas} \equiv$ M$_{gas}$ / SFR, where M$_{gas} \equiv 1.33 \times M_{HI}$ to correct for helium;][]{Roberts1963} assumes a 30\% recycle fraction of stellar mass \citep{Kennicutt1994} and ranges from $2 - 34$ Gyr; individual values are listed in Table~\ref{tab:sf}. Given the low baryonic content of the galaxies, it is unlikely that their masses have been significantly altered by mergers or gas in-fall. Rather, the longer gas consumption timescales for the SHIELD galaxies indicate that these low-mass galaxies are inefficient at turning their gas content into stars relative to more massive galaxies.
\section{Conclusions \label{conc} }
We have used optical imaging of resolved stellar populations obtained from the $\textit{HST}$ to measure the star-formation activity of 12 low-mass, gas-rich galaxies in the SHIELD program. The average lifetime SFRs range from $10^{-3.5}$ to $10^{-2.2}$ M$_{\odot}$~yr$^{-1}$. The stellar masses measured from the resolved stellar populations range from $3\times10^6 - 7\times10^7$ M$_{\odot}$, with gas fractions (M$_{HI}$ / M$_*$) between $0.16-1.6$. Overall, the galaxies have recent SFRs comparable to, or slightly higher than, their lifetime SFR averages; none show significantly lower levels of recent star-formation activity compared to historical values. The ratio of recent to lifetime average SFRs fall in a narrow range with a mean value of 1.4 and dispersion of 0.7; this is a much smaller range than reported in other studies of gas-rich, low-mass galaxies \citep[e.g.,][]{Brinchmann2004, Ott2012, Huang2012}. The larger scatter found in these other studies may reflect uncertainties in the measurements or in our comparison (see Section~4), or may be intrinsic to the galaxies. Future work expanding the SHIELD sample to include SFHs of 30 galaxies will provide a larger, homogeneous sample in which to study the birthrate parameter (K.~B.~W.~Mcquinn et al. in preparation).
We classify two galaxies in the sample, AGC~749241 and AGC~748778, as dTrans galaxies based on their gas content \citep{Cannon2011} and weak or non-detected H$\alpha$ emission \citep{Haurberg2014}. These two systems have star-formation and gas properties consistent with dIrrs both in the SHIELD sample and in previous studies \citep{Begum2008, Bothwell2009, Ott2012, Huang2012}. As both galaxies have measurable recent SFRs based on their resolved stellar populations, the lack of significant H$\alpha$ emission is more likely a result of either stochastic sampling of the upper-end of the IMF or a SFR changing on short timescales, rather than a fundamental difference in galaxy type or properties.
We used the 3-D distribution of galaxies mapped by \citet{McQuinn2014} around the SHIELD sample to look for correlations between the environments of the galaxies and their recent star-formation properties. Overall, we find no correlation between the distance to the nearest known neighbor and the recent versus lifetime SFR of a system. The more isolated galaxies tend to have higher than average SFRs and the galaxies located near other systems have lower than average SFRs. While we cannot rule out the possibility of an undetected system near the galaxies with higher recent SFRs, no other systems have been found in the ALFALFA data \citep{Haynes2011}, these HST images \citep{McQuinn2014}, nor ancillary VLA data \citep{Cannon2011}. Thus, the recent star-formation activity in these relatively low-density environments appears to be governed by internal evolutionary processes, or the galaxies in question may have experienced a delayed onset to star formation similar to the gas-rich low-mass galaxies Leo~A \citep{Cole2007}, Leo~T \citep{Weisz2012}, and Aquarius \citep{Cole2014}.
The star-formation activity in these very low-mass galaxies appears to be both inefficient and intrinsically variable. In support of an inefficient star-formation process, not only are the recent and lifetime SFRs low despite being gas-rich, but the oxygen abundances measured from optical spectroscopy are low, indicating an inefficient chemical evolution process \citep{Haurberg2014}. These authors have shown that 2 of the 12 SHIELD galaxies are classified as extremely metal-deficient galaxies \citep[XMD galaxies, e.g.,][]{Kunth2000} with oxygen abundances below 12 $+$ log(O/H) $\leq$ 7.65. In support of a variable, non-deterministic process, we find a high degree of dispersion between the SSFR, stellar mass, and gas fraction of the sample, although we do not report any significantly low values of SSFR for a given stellar mass. This idea of variable or ``flickering'' star-formation activity in low-mass galaxies is further supported by the lack of correlation between systems with higher recent SFRs and the distance to the nearest neighboring system. These results suggest that the star-formation process is governed by the changing internal conditions that are continuously shaped by the star-formation process itself.
Future work on the SHIELD sample will include a spatially resolved study of the H{\sc i}\ mass densities and a comparison of the gas properties to the SFRs, infrared emission from Spitzer warm-mission infrared, and ultraviolet emission from GALEX imaging (J.~M.~Cannon et al. in preparation). Additional analysis of the star-formation process from the ground-based H$\alpha$ observations is also forthcoming (N.~C.~Haurberg in preparation). Further, our results will be expanded to include the study of a larger sample of low-mass galaxies identified from the ALFALFA catalog in the SHIELD II survey.
\section{Acknowledgments}
Support for this work was provided by NASA through grant GO-12658 from the Space Telescope Institute, which is operated by Aura, Inc., under NASA contract NAS5-26555. JMC is supported by NSF grant AST-1211683. Partial support for publication charges was provided by the NRAO. The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc. The authors acknowledge the work of the entire ALFALFA collaboration team in observing, flagging, and extracting the catalogue of galaxies used to identify the SHIELD sample. The ALFALFA team at Cornell is supported by NSF grant AST-1107390 to R.G. and M.P.H. and by a grant to M.P.H. from the Brinson Foundation. This research made use of NASA's Astrophysical Data System and the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. The authors thank the anonymous referee for helpful and constructive comments.
{\it Facilities:} \facility{$\textit{HST}$}
|
\section{Introduction}
Quantum information and computation \cite{NC2011, Preskill97} is an interdisciplinary research field of applying fundamental principles
of quantum mechanics to information science and computer science. It represents a further development of quantum mechanics, and indeed
helps us to achieve deeper understandings on quantum physics. Quantum entanglement \cite{EPR35,Bell64} as a fundamental concept of distinguishing
classical physics from quantum physics has become a widely exploited resource in quantum information and computation. It has been experimentally
verified that, under the assistance of quantum entanglement, an unknown qubit can be transmitted from Alice to Bob without any non-local
physical interaction between them, using the quantum information protocol called quantum teleportation \cite{BBCCJPW93,Vaidman94,BDM00,Werner01}.
Hence both quantum entanglement and quantum teleportation oblige quantum physicists to think about the relationship between Einstein's locality
and quantum non-locality \cite{EPR35,Bell64}.
Fault-tolerant quantum computation \cite{NC2011, Preskill97} is required in practice to overcome decoherence, and
teleportation-based quantum computation \cite{GC99,Nielsen03, Leung04, ZP13, ZZ14} is a powerful approach to fault-tolerant
quantum computation in which universal quantum gate set is protected from noise using the teleportation protocol. Besides this,
teleportation-based quantum computation is an example of measurement-based quantum computation which exploits quantum measurement
as the main computing resource to determine which quantum gate is to be performed. In quantum mechanics, quantum measurement breaks
quantum coherence and is usually performed at the end of an experiment, so measurement-based quantum computation essentially changes
our conventional viewpoint on quantum measurement and also on the standard quantum circuit model \cite{Nielsen03} in which a coherent
unitary dynamics is mainly involved.
As the title of this paper claims, we apply the extended Temperley--Lieb diagrammatical approach and the Yang--Baxter gate approach to
the reformulation of teleportation-based quantum computation. The Temperley--Lieb algebra \cite{TL71} is a well known concept in
both statistical mechanics and low dimensional topology \cite{Kauffman02}, and the Yang--Baxter equation \cite{YBE67} arises
in exactly solving both some $1+1$-dimensional quantum many-body systems and vertex models in statistics physics. The fact \cite{Kauffman02}
that a type of solutions of the Yang--Baxter equation can be constructed using the Temperley--Lieb algebra motivates the authors to
explore teleportation-based quantum computation using the two related approaches.
The extended Temperley--Lieb diagrammatical approach \cite{Kauffman05, Zhang06} is an extension of the standard diagrammatical representation
of the Temperley--Lieb algebra, and with it, a quantum information protocol involving bipartite maximally entangled states, such as quantum
teleportation, has a very nice topological diagrammatical interpretation. The Yang--Baxter gates are nontrivial unitary solutions of the
Yang--Baxter equation, and the Yang--Baxter gate approach \cite{Dye03, KL04, ZKG04} to quantum information and computation
is an algebraic method originally motivated by the observation that topological entanglements (like braiding configurations \cite{Kauffman02})
and quantum entanglements may have a kind of connection. Therefore, our research is expected to be interesting for physicists in quantum information
and computation, which shows that teleportation-based quantum computation admits both topological and algebraic descriptions besides its standard description
in quantum circuit model \cite{GC99,Nielsen03, Leung04}.
As we have introduced, we are going to go directly into at least three different research subjects in this paper, including quantum information
and computation, the Temperley--Lieb algebra and the Yang--Baxter equation. The guiding principle of this kind of interdisciplinary research is
that we want to explore the nature of quantum entanglement (quantum non-locality \cite{EPR35,Bell64}). As a matter of fact, nowadays, nobody is able to
state that the nature of quantum entanglement has been fully understood \cite{NC2011, Preskill97}. In accordance with \cite{KL02}, we study the nature
of quantum entanglement by setting up a link between quantum entanglement (quantum non-locality ) and
topological entanglement (topological non-locality \cite{Kauffman02}).
Topological entanglement in the paper is represented by the extended Temperley--Lieb diagrammatical
configuration or the Yang--Baxter gate configuration (the braiding configuration), while quantum entanglement is represented by bipartite maximally
entangled two-qubit pure states, i.e., the Bell states \cite{NC2011, Preskill97}. In the extended Temperley--Lieb diagrammatical
approach \cite{Kauffman05, Zhang06}, both the Bell states and Bell measurements have the Temperley--Lieb diagrammatical configurations
on which quantum gates are allowed to move from this qubit to that qubit. In the Yang--Baxter gate approach \cite{Dye03, KL04, ZKG04}, the algebraic
formulation of teleportation-based quantum computation admits the braiding configuration on which quantum gates are permitted to move. Especially,
the relationship between such the two approaches will be clarified in this paper.
The complete scheme of teleportation-based quantum computation was proposed by Gottesman and Chuang \cite{GC99} in a very brief style, so it is necessary
firstly to make a detailed review on teleportation-based quantum computation, topics including quantum teleportation, the fault-tolerant construction of
universal quantum gate set and the fault-tolerant preparation of four-qubit entangled states. Then we present a topological diagrammatical description of
such the reviewed topics in the extended Temperley--Lieb diagrammatical approach. Afterwards, we concentrate on the Yang--Baxter gates which are the Bell
transform, a unitary basis transformation from the product basis to the Bell states, and with the help of our previous research on teleportation-based
quantum computation using the Bell transform \cite{ZZ14}, we go into the algebraic description of teleportation-based quantum computation in terms of
the Yang--Baxter gate. Finally, in view of the extended Temperley--Lieb diagrammatical representation of the Yang--Baxter gate, we set up a transparent link
between the extended Temperley--Lieb diagrammatical approach and the Yang--Baxter gate approach.
Our study in this paper is meaningful and useful in the following sense. First, as we have emphasized, we try to dig out the nature of quantum non-locality from
the viewpoint of topological non-locality. Second, the topological features in teleportation-based quantum computation make the fault-tolerant construction of
both universal quantum gate set and four-qubit entangled states more intuitive and simpler. Third, we develop the concept of teleportation operator \cite{ZZ14}
in the Yang--Baxter gate approach to catch the intrinsic characteristic of quantum teleportation, which is capable of including the algebraic formulations of
all possible teleportaiton processes. Fourth, the methodologies underlying our research are expected to be applied to other subjects in quantum information and
computation, see Section~\ref{Concluding remarks} for concluding remarks on further research. Fifth, our research is expected to shed a light on further research
in mathematical physics, see Appendix~A.
It is obvious that what we have done in this paper is an interdisciplinary research among quantum information and computation, the low-dimensional topology and
mathematical physics such as the Yang--Baxter equation. But we aim at introducing both the Temperley--Lieb algebra and the Yang--Baxter equation to physicists in
quantum information and computation, in other words, readers can go through the entire paper without the preliminary knowledge on concepts in mathematical physics.
All the results about both the Temperley--Lieb algebra and the Yang--Baxter equation are collected in Appendix A. Furthermore, for experts in mathematical
physics, Appendix~\ref{YBE_further_research} indeed presents a list of open problems for further research which are based on our reformulation of
teleportation-based quantum computation.
This paper is organized as follows. Section 2 is a review on teleportation-based quantum computation \cite{GC99}.
Section 3 focuses on the topological diagrammatical description of teleportation-based quantum computation in the extended Temperley--Lieb diagrammatical
approach. Section 4 presents the extended Temperley--Lieb configurations of two types of Yang--Baxter gates derived in Appendix A. Section 5 describes
the Yang--Baxter gate approach to teleportation-based quantum computation. Section 6 clarifies the relationship between the extended Temperley--Lieb
diagrammatical approach and the Yang--Baxter gate approach. To make the paper self-consistent, we present five appendices for readers' conveniences.
Appendix A reviews both the Temperley--Lieb algebra and the Yang--Baxter equation as well as presents a detailed construction on the Yang--Baxter gate
via the Temperley--Lieb algebra. Appendix B collects interesting properties of the permutation-like Yang--Baxter gates. Both Appendix C and Appendix D
present the topological diagrammatical construction of four-qubit entangled states respectively associated with the $CNOT$ gate and $CZ$ gate \cite{NC2011}.
Appendix E is about a method of calculating the extended Temperley--Lieb diagrammatical representation of the teleportation operator.
\section{Review on teleportation-based quantum computation}
\label{section review teleportation-based quantum computation}
In this section, we make a review on teleportation-based quantum computation \cite{GC99}. The key topics include
the standard description of quantum teleportation, the fault-tolerant construction of universal quantum gate set,
and the fault-tolerant construction of four-qubit entangled states. Meanwhile, we set up our notation and convention
for the study in the paper.
\subsection{Notation}
A single-qubit Hilbert space ${\mathcal H}_2$ has an orthonormal basis denoted by $|i\rangle$, $i=0,1$, and a single-qubit
state $|\alpha\rangle$ is given by $|\alpha\rangle= a|0\rangle +b |1\rangle$ with complex numbers $a$ and $b$.
The unit matrix $1\!\! 1_2$ and the Pauli gates $X$ and $Z$ take the form
\begin{equation}
1\!\! 1_2 =\left(\begin{array}{cc}
1 & 0 \\
0 & 1 \\
\end{array}\right), \quad X=\left(\begin{array}{cc}
0 & 1 \\
1 & 0 \\
\end{array} \right), \quad Z=\left(\begin{array}{cc}
1 & 0 \\
0 & -1 \\
\end{array} \right)
\end{equation}
with $Z |0\rangle=|0\rangle$ and $Z |1\rangle=-|1\rangle$, and the Pauli gate $Y$ is defined as $Y=ZX$.
A single-qubit gate is an element of the unitary group $U(2)$, and a typical single-qubit gate $W_{ij}$ in the present paper
has the form
\begin{equation}
\label{W ij}
W_{ij}=X^i Z^j, \quad i, j=0,1,
\end{equation}
satisfying $W_{ij}^T=W_{ij}^\dag$, in which the upper index $T$ denotes the matrix transpose conjugation and the upper index
$\dag$ denotes the matrix Hermitian conjugation.
A two-qubit Hilbert space ${\mathcal H}_2 \otimes {\mathcal H}_2$ has an orthonormal product basis denoted by $|ij\rangle, i,j=0,1$.
The EPR state $|\Psi\rangle$ \cite{EPR35,Bell64} takes the form
\begin{equation}
\label{EPR}
|\Psi\rangle =\frac 1 {\sqrt 2} (|00\rangle+|11\rangle)
\end{equation}
and it has a very nice algebraic property
\begin{equation}
\label{property of EPR}
(1\!\! 1_2 \otimes U)|\Psi\rangle =(U^T \otimes 1\!\! 1_2) |\Psi\rangle
\end{equation}
with $U$ denoting any single-qubit gate. The set of the orthonormal Bell states $|\psi(ij)\rangle$ given by
\begin{equation}
\label{Bell states}
|\psi(ij)\rangle = (1\!\! 1_2 \otimes W_{ij}) |\Psi\rangle, \quad i,j=0,1
\end{equation}
is called the Bell basis \cite{NC2011, Preskill97} of the two-qubit Hilbert space ${\mathcal H}_2 \otimes {\mathcal H}_2$. Obviously, $|\Psi\rangle=|\psi(00)\rangle$.
\subsection{Quantum teleportation}
\begin{figure}
\begin{center}
\includegraphics[width=5cm]{tele_circuit.eps}
\end{center}
\caption{\label{fig_tele_circuit} Quantum circuit for quantum teleportation as a diagrammatical representation of
the teleportation equation (\ref{tele_eq}). An unknown qubit $|\alpha\rangle$ is sent from Alice to Bob. The top two lines
represent Alice's system and the bottom line represents Bob's system. The single-lines denote qubits and
the double-lines denote classical bits. The triangle box represents the Bell measurement performed by Alice
and the square box represents the local unitary correction operator $W_{ij}^\dag$ performed by Bob to obtain
the transmitted state $|\alpha\rangle$.}
\end{figure}
Let Alice and Bob share the EPR state $|\Psi\rangle$, and Alice wants to transfer
an unknown quantum state $|\alpha\rangle$ to Bob. Alice and Bob prepare the quantum
state $|\alpha\rangle\otimes |\Psi\rangle$ which can be formulated as
\begin{equation}
\label{tele_eq}
|\alpha\rangle \otimes |\Psi\rangle =\frac 1 2 \sum_{i,j=0}^1 |\psi(ij)\rangle \otimes W_{ij} |\alpha\rangle
\end{equation}
which was called the teleportation equation in \cite{Zhang06}. Then, Alice performs the Bell measurements
$|\psi(ij)\rangle\langle\psi(ij)|\otimes 1\!\! 1_2$ on the prepared state
$|\alpha\rangle\otimes |\Psi\rangle$,
\begin{equation}
(|\psi(ij)\rangle\langle\psi(ij)|\otimes 1\!\! 1_2)(|\alpha\rangle\otimes |\Psi\rangle)
=\frac 1 2 |\psi(ij)\rangle \otimes W_{ij} |\alpha\rangle,
\end{equation}
and she informs Bob her measurement results labeled as $(i,j)$.
Finally, Bob applies the unitary correction operator $W_{ij}^\dag$ on his state
\begin{equation}
(1\!\! 1_2 \otimes 1\!\! 1_2 \otimes W^\dag_{ij} ) (|\psi(ij)\rangle \otimes W_{ij} |\alpha\rangle)
= |\psi(ij)\rangle \otimes |\alpha\rangle
\end{equation}
to obtain the transmitted quantum state $|\alpha\rangle$. See Figure \ref{fig_tele_circuit} for
the description of the teleportation protocol in the language of quantum circuit.
Note that the teleportation protocol of Bob transmitting an unknown qubit $|\alpha\rangle$ to Alice admits another
type of the teleportation equation
\begin{equation}
\label{tele_eq_transpose}
|\Psi\rangle \otimes |\alpha\rangle =\frac 1 2 \sum_{i,j=0}^1 W^T_{ij} |\alpha\rangle \otimes |\psi(ij)\rangle
\end{equation}
which was called the transpose teleportation equation in \cite{ZZ14} and is to be used
in the fault-tolerant construction of two-qubit gates in Subsubsection~\ref{subsection: two-qubit}.
\subsection{Fault-tolerant construction of universal quantum gate set}
\label{fault_tolerant_u_g_s}
Quantum gates \cite{Barenco95b} are defined as unitary transformation matrices acting on quantum states, and the set of all $n$-qubit gates forms
a representation of the unitary group $U(2^n)$.
\subsubsection{Universal quantum gate set}
An entangling two-qubit gate \cite{BB02} with single-qubit gates is called a universal quantum gate set which can perform universal quantum
computation in the circuit model \cite{NC2011} of quantum computation. All single-qubit gates can be generated by both the Hadamard gate $H$
given by
\begin{equation}
\label{Hadamard gate}
H=\frac 1 {\sqrt 2} (X+Z),
\end{equation}
and the $\pi / 8$ gate \cite{BMPRV00} given by
\begin{equation}
\label{pi/8}
T=\left(\begin{array}{cc}
1 & 0\\
0 & e^{i \frac \pi 4}\\
\end{array}\right).
\end{equation}
An entangling two-qubit gate \cite{BB02} is defined as a two-qubit gate capable of transforming a tensor product of two single-qubit states
into an entangling two-qubit state. For examples, the $CNOT$ gate
\begin{equation}
\label{CNOT gate}
CNOT = |0\rangle\langle 0| \otimes 1\!\! 1_2 + |1\rangle\langle 1| \otimes X,
\end{equation}
and the $CZ$ gate
\begin{eqnarray}
\label{CZ gate}
CZ = |0\rangle\langle 0| \otimes 1\!\! 1_2 + |1\rangle\langle 1| \otimes Z.
\end{eqnarray}
They are maximally entangling gates \cite{BG11}, namely, with single-qubit gates, they can generate the maximally entangling states such as the Bell
states (\ref{Bell states}) from the product states. The $CNOT$ gate is the well known two-qubit gate in quantum information and computation
which is the quantum analogue of the Exclusive OR gate in classical computation \cite{NC2011}. The $CZ$ gate is widely used in the one-way
quantum computation \cite{RB01}, the representative example of measurement-based quantum computation, and it is related to the $CNOT$ gate
in the way
\begin{equation}
CZ=(1\!\! 1_2 \otimes H) CNOT (1\!\! 1_2\otimes H)\equiv H_2\,\, CNOT\,\, H_2
\end{equation}
in which the subscript of the Hadamard gate $H$ means that it is acting on the second qubit. Hence the set of the $CNOT$ gate (\ref{CNOT gate}) (or
the $CZ$ gate (\ref{CZ gate})) with single-qubit gates $H$ (\ref{Hadamard gate}) and $T$ (\ref{pi/8}) is called a universal quantum gate set
to perform universal quantum computation \cite{NC2011}.
\subsubsection{Clifford gates and fault-tolerant quantum computation}
\label{clifford_gate_fault_tolerant}
The Pauli group gates $C_1$ are generated by tensor products of the Pauli matrices $X$, $Z$ and
the identity matrix $1\!\! 1_2$ with global phase factors $\pm 1, \pm i$. Quantum gates $U$
\cite{NC2011} are classified by
\begin{equation}
\label{C_k}
C_k \equiv \{ U | U C_{k-2} U^\dag \subseteq C_{k-1} \}
\end{equation}
where $C_2$ denotes the Clifford gates preserving
the Pauli group gates under conjugation.
Obviously, the Hadamard gate $H$ is a Clifford gate, namely $H \in C_2$, due to
\begin{equation}
H X H =Z, \quad H Z H =X,
\end{equation}
and the $\pi / 8$ gate $T$ is not a Clifford gate, $T\in C_3$, since
\begin{equation}
T X T^\dag =\frac {X- \sqrt{-1} Y} {\sqrt 2}, \quad T Z T^\dag = Z
\end{equation}
in which $ ({X- \sqrt{-1} Y})/ {\sqrt 2}$ is a Clifford gate \cite{NC2011, Gottesman97}. Here is another equivalent definition of the
Clifford gate \cite{NC2011, Gottesman97}. If a quantum gate can be represented as a tensor product of the Hadamard
gate (\ref{Hadamard gate}), the phase gate
\begin{equation}
\label{phase gate}
S=\left(
\begin{array}{cc}
1 & 0 \\
0 & i \\
\end{array}
\right),
\end{equation}
and the $CNOT$ gate (\ref{CNOT gate}), then it is a Clifford gate. Note that the phase gate $S$ is the square of
the $\pi / 8$ gate (\ref{pi/8}), $S=T^2$, and the square of the phase gate $S$ is the Pauli $Z$ gate, $S^2=Z$.
In fault-tolerant quantum computation \cite{Gottesman97}, the Pauli group gates $C_1$ and Clifford gates $C_2$ can be easily
performed in principle, but the $C_3$ gates may be difficultly realized. The teleportation-based quantum computation \cite{GC99}
is fault-tolerant quantum computation in the following sense: to perform a $C_3$ gate, it prepares a quantum state with the action
of $C_3$ gate and then applies $C_1$ or $C_2$ gates to such the quantum state using the teleportation protocol.
\subsubsection{Fault-tolerant construction of single-qubit gates}
\label{subsection: single-qubit}
\begin{figure}
\begin{center}
\includegraphics[width=6cm]{tele_single_gate.eps}
\end{center}
\caption{\label{fig_tele_single_gate} Quantum circuit for the fault-tolerant construction of the single-qubit gate $U$
associated with the teleportation equation (\ref{tele_single}). With the such diagrammatical representation, it is obvious that
fault-tolerantly implementing the single-qubit $U$ becomes how to fault-tolerantly perform the $R^{\dag}_{ij}$ gate and prepare
the preliminary quantum state $|\Psi_U\rangle$ (\ref{Psi_U}). }
\end{figure}
To perform a single-qubit gate $U\in C_k$ (\ref{C_k}) on the unknown state $|\alpha\rangle$, Alice prepares the quantum
state $|\Psi_U\rangle$ given by
\begin{equation}
\label{Psi_U}
|\Psi_U\rangle=(1\!\! 1_2\otimes U)|\Psi\rangle
\end{equation}
and reformulate $|\alpha\rangle\otimes |\Psi_U\rangle$ as
\begin{equation}
\label{tele_single}
|\alpha\rangle\otimes|\Psi_U\rangle=\frac{1}{2}\sum_{i,j=0}^1|\psi(ij)\rangle \otimes R_{ij}U|\alpha\rangle,
\end{equation}
where the single-qubit gate $R_{ij}$ has the form $R_{ij}=UW_{ij}U^\dag$ with $W_{ij}$ defined in (\ref{W ij}). Then Alice makes
Bell measurements $|\psi(ij)\rangle\langle\psi(ij)|\otimes 1\!\! 1_2$ and informs Bob her measurement results labeled by $(i,j)$. Finally,
Bob performs the unitary correction operator $R_{ij}^\dag \in C_{k-1}$ (\ref{C_k}) to attain $U |\alpha\rangle$. See Figure \ref{fig_tele_single_gate}
for the quantum circuit of fault-tolerantly performing the single-qubit gate $U$ using the teleportation protocol.
As the single-qubit gate $U$ is the Hadamard gate $H$ (\ref{Hadamard gate}), the single-qubit gate $R_{ij}$ is given by
$R_{ij}(H)=W_{ji}^T$, which is a tensor product of Pauli gates. And as the single-qubit gate $U$ is the $\pi/8$ gate (\ref{pi/8}),
the single-qubit gate $R_{ij}$ is given by
\begin{equation}
\label{RTij}
R_{ij}(T)=(\frac {X- \sqrt{-1} Y} {\sqrt 2})^i\, Z^j,
\end{equation}
which can be easily fault-tolerantly performed, because it is a Clifford gate.
Note that the difficulty of performing a single-qubit gate $U\in C_k$ becomes how to
fault-tolerantly prepare the state $|\Psi_U\rangle$ and perform the single-qubit gate $R_{ij}^\dag \in C_{k-1}$.
\subsubsection{Fault-tolerant construction of two-qubit Clifford gates}
\label{subsection: two-qubit}
\begin{figure}
\begin{center}
\includegraphics[width=8cm]{tele_two_gate.eps}
\end{center}
\caption{\label{fig_tele_two_gate} Quantum circuit for the fault-tolerant construction of a two-qubit Clifford gate $CU$,
as a diagrammatical representation of the teleportation equation (\ref{tele_two}). Specifically, the single-qubit gates $Q$ and $P$
are calculated with the formula (\ref{Q P}). }
\end{figure}
To perform a two-qubit Clifford gate $CU$ such as the $CNOT$ gate (\ref{CNOT gate}) and the $CZ$ gate (\ref{CZ gate}) on two unknown
single-qubit states $|\alpha\rangle$ and $|\beta\rangle$, we prepare a four-qubit entangled state $|\Psi_{CU}\rangle$,
\begin{equation}
\label{Psi_CU}
|\Psi_{CU}\rangle=(1\!\! 1_2\otimes CU \otimes 1\!\! 1_2)(|\Psi\rangle \otimes |\Psi\rangle),
\end{equation}
with the action of the $CU$ gate, and reformulate the prepared state $|\alpha\rangle\otimes|\Psi_{CU}\rangle\otimes|\beta\rangle$ as
\begin{equation}
\label{tele_two}
\begin{split}
& |\alpha\rangle\otimes|\Psi_{CU}\rangle\otimes|\beta\rangle\\
=&\frac{1}{4}\sum_{i_1,j_1=0}^1\sum_{i_2,j_2=0}^1(1\!\!1_4\otimes
Q\otimes P\otimes 1\!\!1_4) (|\psi(i_1j_1)\rangle\otimes CU|\alpha\beta\rangle\otimes |\psi(i_2j_2)\rangle)
\end{split}
\end{equation}
with $1\!\! 1_4=1\!\! 1_2\otimes 1\!\! 1_2$, which is called the teleportation equation associated with the fault-tolerant
construction of the two-qubit Clifford gate $CU$.
The single-qubit gates $Q$ and $P$ in the teleportation equation (\ref{tele_two}) are calculated by
\begin{equation}
\label{Q P}
Q\otimes P=CU(W_{i_1j_1}\otimes W^T_{i_2j_2})CU^\dag.
\end{equation}
As the $CU$ gate is the $CNOT$ gate, the gates $Q$ and $P$ are expressed as
\begin{equation}
\label{Q P for CNOT}
Q=Z^{j_2}X^{i_1}Z^{j_1},\quad P=Z^{j_2}X^{i_2}X^{i_1}.
\end{equation}
As the $CU$ gate is the $CZ$ gate, the gates $Q$ and $P$ have the form
\begin{equation}
\label{Q P for CZ}
Q=Z^{i_2}X^{i_1}Z^{j_1},\quad P=Z^{j_2}X^{i_2}Z^{i_1}.
\end{equation}
Note that the $Q$ and $P$ gates (\ref{Q P}) may be not single-qubit gates if the $CU$ gate is not a Clifford gate in
accordance with the definition of a Clifford gate \cite{NC2011, Gottesman97}.
Next, we perform the Bell measurements given by
\begin{equation}
\label{six_bell_measurement}
|\psi(i_1 j_1)\rangle\langle\psi(i_1 j_1)|\otimes 1\!\! 1_2\otimes 1\!\! 1_2 \otimes |\psi(i_2 j_2)\rangle\langle\psi(i_2 j_2)|,
\end{equation}
and with the measurement results labeled by $(i_1,j_1)$ and $(i_2,j_2)$, we perform the unitary correction operator, $Q^\dag\otimes P^\dag$, to
obtain the exact action of the Clifford gate $CU$ on the two-qubit state $|\alpha\rangle \otimes |\beta\rangle$, namely, $CU|\alpha\beta\rangle$.
The quantum circuit for the fault-tolerant construction of the two-qubit Clifford gate $CU$ using the teleportation protocol is depicted
in Figure~\ref{fig_tele_two_gate}. Note that the two-qubit Clifford gate $CU$ we study here may be not the controlled-operation two-qubit
gates such as the $CNOT$ gate (\ref{CNOT gate}) and the $CZ$ gate (\ref{CZ gate}).
\subsection{Construction of four-qubit entangled states}
From Subsection~\ref{fault_tolerant_u_g_s}, we already know that, the conceptual point in the fault-tolerant construction of a single-qubit quantum
gate $U$ (or a two-qubit Clifford gate $CU$) using quantum teleportation is the fault-tolerant preparation of the two-qubit entangled state (\ref{Psi_U}) (or
the four-qubit entangled state (\ref{Psi_CU})). In this subsection, we focus on the fault-tolerant preparation of the four-qubit entangled state
$|\Psi_{CU}\rangle$ (\ref{Psi_CU}) using the teleportation protocol, when the $CU$ gate is the $CNOT$ gate (\ref{CNOT gate}) and the $CZ$ gate (\ref{CZ gate})
respectively.
In Gottesman and Chuang's original proposal \cite{GC99} of teleportation-based quantum computation, a four-qubit entangled state $|\Psi_{CU}\rangle$ is
prepared in the following steps. First, we prepare a prior entangled six-qubit state as a tensor product of a three-qubit GHZ state $|\Upsilon\rangle$
\cite{GHZ90},
\begin{equation}
\label{Upsilon}
|\Upsilon\rangle=\frac 1 {\sqrt 2} (|000\rangle+|111\rangle)
\end{equation}
and another three-qubit GHZ state $|\Upsilon\rangle$ with the local action of the Hadamard gate $H$. Second, we perform the Bell measurements on such
the six-qubit entangled state. Third, after classical communication of the Bell measurement outcomes, we perform the unitary correction operators to
obtain the four-qubit entangled state $|\Psi_{CU}\rangle$.
\subsubsection{Construction of the four-qubit entangled state $|\Psi_{CNOT}\rangle$}
\label{construction Psi CNOT}
\begin{figure}
\begin{center}
\includegraphics[width=8cm]{psi_cnot1}
\end{center}
\caption{\label{fig_psi_cnot1} Quantum circuit for the construction of four-qubit entangled state $|\Psi_{CNOT}\rangle$ (\ref{Psi_cnot}), as a
diagrammatical representation of the teleportation equation (\ref{tele_cnot1}). The $|\Upsilon\rangle$ state is the three-quibt GHZ state (\ref{Upsilon})
and the single-qubit gate denoted as $H$ is the Hadamard gate (\ref{Hadamard gate}). The $CNOT$ gate (\ref{CNOT gate}) takes the conventional
configuration in quantum circuit model \cite{NC2011}. }
\end{figure}
The four-qubit entangled state $|\Psi_{CNOT}\rangle_{1256}$ is the target state we want to construct,
\begin{equation}
\label{Psi_cnot}
|\Psi_{CNOT}\rangle_{1256}=(1\!\! 1_2 \otimes CNOT_{25} \otimes 1\!\! 1_2) (|\Psi\rangle_{12} \otimes |\Psi\rangle_{56})
\end{equation}
in which the $CNOT_{ij}$ gate requires the $i$-th qubit as the control qubit and the $j$-th qubit as
the target qubit. We prepare the six-qubit entangled state $H_4H_5H_6|\Upsilon\rangle_{123}|\Upsilon\rangle_{456}$,
which can be reformulated as
\begin{equation}
\label{tele_cnot1}
H_4H_5H_6|\Upsilon\rangle_{123}|\Upsilon\rangle_{456}=\sum_{i,j=0}^1|\psi(ij)\rangle_{34}\,\, Z_2^jX_1^iX_2^i|\Psi_{CNOT}\rangle_{1256},
\end{equation}
then make the joint Bell measurement on both the third and fourth qubits given by
\begin{equation}
\label{Bell_measurement_CNOT}
1\!\! 1_2 \otimes 1\!\! 1_2 \otimes|\psi(ij)\rangle_{34} \langle \psi(ij) | \otimes 1\!\! 1_2 \otimes 1\!\! 1_2.
\end{equation}
With the measurement outcome $(i,j)$, we perform the unitary correction operator
\begin{equation}
X^i\otimes X^iZ^j\otimes 1\!\! 1_2 \otimes 1\!\! 1_2\otimes 1\!\! 1_2 \otimes 1\!\! 1_2,
\end{equation}
on the resultant quantum state to obtain the four-qubit entangled state $|\Psi_{CNOT}\rangle$ (\ref{Psi_cnot}). The quantum
circuit for the construction of the $|\Psi_{CNOT}\rangle$ state is shown in Figure~\ref{fig_psi_cnot1}.
Note that the teleportation equation (\ref{tele_cnot1}) admits an equivalent form
\begin{equation}
\label{tele_cnot2}
H_4H_5H_6|\Upsilon\rangle_{123}|\,\,\Upsilon\rangle_{456}
=\sum_{i,j=0}^1|\psi(ij)\rangle_{34}\,\, X_5^iZ_5^jZ_6^j|\Psi_{CNOT}\rangle_{1256},
\end{equation}
so that we have the second method of constructing the $|\Psi_{CNOT}\rangle$ state (\ref{Psi_cnot}) using the teleportation protocol. These
two methods are found to be equivalent in the low dimensional topological diagrammatical approach, see Subsubsection~\ref{topo_construction_four_CNOT}.
\subsubsection{Construction of four-qubit entangled states $|\Psi_{CNOT}^\uparrow\rangle$ and $|\Psi_{CZ}\rangle$}
In Gottesman and Chuang's original study \cite{GC99}, the fault-tolerant construction of the four-qubit entangled state $|\Psi_{CNOT}^\uparrow\rangle$ (\ref{Psi cnot down})
was presented instead of the state $|\Psi_{CNOT}\rangle$ (\ref{Psi_cnot}) that we are working on. For readers' convenience, the construction of the $|\Psi_{CNOT}^\uparrow\rangle$
state is discussed in Appendix~\ref{construction cnot down}. Besides $|\Psi_{CNOT}\rangle$ and $|\Psi_{CNOT}^\uparrow\rangle$, since the $CZ$ gate (\ref{CZ gate}) has a wide application in measurement-based quantum computation \cite{RB01}, we work out the construction of four-qubit entangled state $|\Psi_{CZ}\rangle$ (\ref{Psi_cz}) in Appendix~\ref{construction Psi CZ}.
\section{The extended Temperley--Lieb diagrammatical approach to teleportation-based quantum computation}
\label{TL diagrammatical for tele based quantum computation}
In this section\footnote{This section is an extended version of the authors' unpublished paper \cite{ZP13} in which topological
features of teleportation-based quantum computation are claimed to be associated with topological features of space-time.
}, we aim at exhibiting topological features of teleportation-based quantum computation in the transparent style. In the
extended Temperley--Lieb diagrammatical approach \cite{Kauffman05, Zhang06}, we study the topological diagrammatical construction
of both universal quantum gate set and four-qubit entangled states in teleportation-based quantum computation \cite{GC99,Nielsen03, Leung04},
whose algebraic counterparts have been presented in Section~\ref{section review teleportation-based quantum computation}.
The extended Temperley--Lieb diagrammatical configuration \cite{Kauffman05, Zhang06} is a kind of the extension of the standard
Temperley--Lieb configuration on which both single-qubit gates and two-qubit gates are allowed to move along the lines under certain
rules. The definition of the Temperley--Lieb algebra and its diagrammatical representation is shown up
in Appendix~\ref{Def TL} and~\ref{Def TL and Bell states}.
\subsection{The extended Temperley--Lieb diagrammatical approach}
A single-qubit state vector $|\varphi\rangle$ is denoted by a vertical line followed by the symbol $\nabla$ on the bottom,
\begin{eqnarray}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(25,16)
\put(2,7){$|\varphi\rangle=$}
\put(25,4){\line(0,1){12}}
\put(23,0){\makebox(4,4){$\nabla$}}
\end{picture}
\end{array}
\end{eqnarray}
and the line with the symbol $\triangle$ on the top represents the covector $\langle \phi|$, so a vertical line with both symbols $\nabla$
and $\triangle$ on the boundary denotes the inner product $\langle \phi|\varphi\rangle$. A vertical line with a solid point represents a
single-qubit gate $U$ acting on the state $|\varphi\rangle$,
\begin{eqnarray}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(35,16)
\put(2,7){$U|\varphi\rangle=$}
\put(30,4){\line(0,1){12}}
\put(30,10){\circle*{2.}}
\put(28,0){\makebox(4,4){$\nabla$}}
\put(32,9){\tiny{$U$}}
\end{picture}
\end{array}
\end{eqnarray}
in which the algebraic expression is read from the right to the left and the diagrammatical representation is read from the bottom to the top.
The diagrammatical configuration of the Bell state $|\psi(ij)\rangle$ (\ref{Bell states}) is the core of the
extended Temperley--Lieb diagrammatical approach, and it is represented by a cup with a solid point
denoting the single-qubit gate $W_{ij}$ (\ref{W ij}),
\begin{eqnarray}
\label{TL diag Bell state}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(45,16)
\put(2,5){$|\psi(ij)\rangle=$}
\put(32,0){\line(0,1){12}}
\put(42,0){\line(0,1){12}}
\put(32,0){\line(1,0){10}}
\put(42,6){\circle*{2.}}
\put(44.,5){\tiny{$W_{ij}$}}
\end{picture}
\end{array}
\end{eqnarray}
so that a cup without a solid point denotes the EPR state $|\Psi\rangle$ (\ref{EPR}).
The adjoint of the Bell state, $\langle \psi(ij)|$, is represented by a cap
\begin{eqnarray}
\label{TL diag Bell state_cap}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(45,16)
\put(2,5){$\langle\psi(ij)|=$}
\put(32,0){\line(0,1){12}}
\put(42,0){\line(0,1){12}}
\put(32,12){\line(1,0){10}}
\put(42,6){\circle*{2.}}
\put(44.,5){\tiny{$W_{ij}^\dag$}}
\end{picture}
\end{array}
\end{eqnarray}
with a solid point denoting the Hermitian conjugation of $W_{ij}$. These diagrammatic states are called
a cup state or a cap state respectively. The projective measurement $|\psi(ij)\rangle\langle\psi(ij) |$
is called the Bell measurement,
\begin{eqnarray}
\label{TL Bell measurement}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(65,30)
\put(62,18){\line(0,1){12}}
\put(52,18){\line(0,1){12}}
\put(52,18){\line(1,0){10}}
\put(62,24){\circle*{2.}}
\put(64.,23){\tiny{$W_{ij}$}}
\put(2,14){$|\psi(ij)\rangle \langle\psi(ij)|=$}
\put(52,0){\line(0,1){12}}
\put(62,0){\line(0,1){12}}
\put(52,12){\line(1,0){10}}
\put(62,6){\circle*{2.}}
\put(64,5){\tiny{$W_{ij}^\dag$}}
\end{picture}
\end{array}
\end{eqnarray}
represented by a top cup state with a bottom cap state. Note that without solid points on the configuration of the Bell measurement,
such the configuration represents the basic generator of the Temperley--Lieb algebra, see
Appendix~\ref{Def TL} and~\ref{Def TL and Bell states}.
The nice property (\ref{property of EPR}) of the EPR state has the corresponding diagrammatical expression
\begin{eqnarray}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(45,16)
\put(2,0){\line(0,1){12}}
\put(12,0){\line(0,1){12}}
\put(2,0){\line(1,0){10}}
\put(12,6){\circle*{2.}}
\put(14.,5){\tiny{$U$}}
\put(23.,6){$=$}
\put(32,5){\tiny{$U^T$}}
\put(40.8,6){\circle*{2.}}
\put(40.8,0){\line(0,1){12}}
\put(50.8,0){\line(0,1){12}}
\put(40.8,0){\line(1,0){10}}
\end{picture}
\end{array}
\label{trans law: 1l}
\end{eqnarray}
with $U$ denoting any single-qubit gate, and the similar representation also for a cap state,
\begin{eqnarray}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(45,16)
\put(2,0){\line(0,1){12}}
\put(12,0){\line(0,1){12}}
\put(2,12){\line(1,0){10}}
\put(12,6){\circle*{2.}}
\put(14.,5){\tiny{$U^\dag$}}
\put(23.,6){$=$}
\put(32,5){\tiny{$U^\ast$}}
\put(40.8,6){\circle*{2.}}
\put(40.8,0){\line(0,1){12}}
\put(50.8,0){\line(0,1){12}}
\put(40.8,12){\line(1,0){10}}
\end{picture}
\end{array}
\label{trans law: 22}
\end{eqnarray}
with the upper index $\ast$ denoting the complex conjugation. In the diagrammatical representations (\ref{trans law: 1l}) and (\ref{trans law: 22}),
a single-qubit gate can flow from the one branch of a cup (or cap) state to its other branch with the transpose conjugation, which is beyond the
conventional utilization of the Temperley--Lieb diagrams but naturally arises in quantum computation. Note that such the operation of moving
single-qubit gates is a crucial technique in the extended Temperley--Lieb diagrammatical approach \cite{Kauffman05, Zhang06} to teleportation-based
quantum circuits.
\subsection{Topological interpretation of quantum teleportation}
In quantum teleportation, Alice and Bob prepare the quantum state $|\alpha\rangle\otimes |\Psi\rangle$, in the extended Temperley--Lieb diagrammatical language,
expressed as
\begin{eqnarray}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(70,16)
\put(10,7){$|\alpha\rangle\otimes |\Psi\rangle=$}
\put(45,0){\makebox(4,4){$\nabla$}}
\put(47,3.7){\line(0,1){10.3}}
\put(57,0){\line(0,1){14}}
\put(67,0){\line(0,1){14}}
\put(57,0){\line(1,0){10}}
\end{picture}
\end{array}
\end{eqnarray}
in which the vertical line with $\nabla$ denotes the unknown quantum state $|\alpha\rangle$ to be transmitted from Alice to Bob. Then, Alice performs
Bell measurements $|\psi(ij)\rangle\langle\psi(ij)|\otimes 1\!\! 1_2$ on the prepared state
$|\alpha\rangle\otimes |\Psi\rangle$, which has the extended Temperley--Lieb diagrammatical configuration
\begin{eqnarray}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(80,45)
\put(10,0){\makebox(4,4){$\nabla$}}
\put(12.,4){\line(0,1){20}}
\put(22,18){\circle*{2.}}
\put(23,17){\tiny{$W_{ij}^\dag$}}
\put(22,0){\line(0,1){24}}
\put(32,0){\line(0,1){42}}
\multiput(8,12)(1,0){30}{\line(1,0){.5}}
\put(22,0){\line(1,0){10}}
\put(12,24){\line(1,0){10}}
\put(12,30){\line(1,0){10}}
\put(12,30){\line(0,1){12}}
\put(22,30){\line(0,1){12}}
\put(22,36){\circle*{2.}}
\put(23,35){\tiny{$W_{ij}$}}
\put(40,21){\makebox(14,10){$=\frac 1 2$}}
\put(62,30){\line(1,0){10}}
\put(62,30){\line(0,1){12}}
\put(72,30){\line(0,1){12}}
\put(72,36){\circle*{2.}}
\put(73,35){\tiny{$W_{ij}$}}
\put(80,0){\makebox(4,4){$\nabla$}}
\put(82,4){\line(0,1){38}}
\put(82,21){\circle*{2.}}
\put(84,20){\tiny{$W_{ij}$}}
\end{picture}
\label{tele: tl}
\end{array}
\end{eqnarray}
with the single-qubit gate $W_{ij}$ defined in (\ref{W ij}). It is obvious that this topological diagram (\ref{tele: tl}) is associated
with both the teleportation equation (\ref{tele_eq}) and the quantum circuit in Figure \ref{fig_tele_circuit}. Note that the
diagram (\ref{tele: tl}) without solid points is a standard diagrammatical representation of the Temperley--Lieb algebra \cite{Kauffman02},
see Appendix~\ref{Def TL} and~\ref{Def TL and Bell states}.
Now let us explain the diagram (\ref{tele: tl}) in detail. On the left hand side of $=$, the diagrammatical part above the dashed line denotes
the Bell measurement, and the part under the dashed line denotes the state preparation. On the right hand side of $=$, the normalization
factor $\frac 1 2$ is contributed by the vanishing cup state and cap state according to the rules of the extend Temperley--Lieb diagrammatical
approach \cite{Kauffman05, Zhang06}. The cup state with the action of $W_{ij}$ denotes the post-measurement state which is usually neglected in
the following study for simplicity. The $W_{ij}$ gate acting on the unknown qubit $|\alpha\rangle$ is due to the operation of
moving $W_{ij}^\dag$ from the one branch of the cap state to the other branch and applying the transposition conjugation, $(W_{ij}^{\dag})^T=W_{ij}$.
Note that both classical communication and unitary correction are not shown in the
diagram (\ref{tele: tl}). In this sense, hence, the quantum information flow sending an
unknown qubit from Alice to Bob in quantum teleportation can be recognized as a result
of topological operation in the extended Temperley--Lieb diagrammatical
approach \cite{Kauffman05, Zhang06}.
The topology in the diagrammatical teleportation (\ref{tele: tl}) may be not
that obvious. Let us consider the topological configuration of the chained teleportation \cite{Childs03}
that Alice sends an unknown qubit $|\alpha\rangle$ to Bob with a sequence of standard teleportation protocols, expressed as
\begin{eqnarray}
\label{chain teleportation}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(70,25)
\put(10,0){\makebox(4,4){$\nabla$}}
\put(12,4){\line(0,1){20}}
\put(22,0){\line(0,1){24}}
\put(32,0){\line(0,1){24}}
\multiput(8,12)(1,0){48}{\line(1,0){.5}}
\put(22,0){\line(1,0){10}}
\put(12,24){\line(1,0){10}}
\put(32,24){\line(1,0){10}}
\put(42,0){\line(0,1){24}}
\put(42,0){\line(1,0){10}}
\put(52,0){\line(0,1){24}}
\put(57,7){\makebox(14,10){$=\frac 1 4$}}
\put(72,0){\makebox(4,4){$\nabla$}}
\put(74,4){\line(0,1){20}}
\end{picture}
\end{array}
\end{eqnarray}
in which the post-measurement states are neglected and only the EPR state measurements $|\Psi\rangle \langle \Psi|$ are considered. The normalization factor $\frac 1 4$
is calculated from two vanishing cup states and two vanishing cap states. Without unitary corrections, Bob obtains the exact quantum state $|\alpha\rangle$. Hence {\em the
teleportation topology in the extended Temperley--Lieb diagrammatical approach is defined as a topological operation which straightens the configuration consisting of cap
states and cup states}. In addition, if we perform the Bell measurements $|\psi(ij)\rangle\langle\psi(ij)|$ instead of the EPR state measurements in the chained
teleportation, first of all, we have to move single-qubit gates along the path formed by top cap states with bottom cup states until boundary points of this path under
the guidance of the properties (\ref{trans law: 1l}) and (\ref{trans law: 22}), and then apply the topological operation by straightening the relevant configuration.
\subsection{Topological construction of universal quantum gate set}
In the authors' knowledge, a topological diagrammatical construction of universal quantum gate set \cite{NC2011, Barenco95b} using quantum teleportation \cite{GC99}
has not been done yet in the published paper, so we study such the realization in the extended Temperley-Lieb diagrammatical approach \cite{Kauffman05,Zhang06}.
The topological diagram for the fault-tolerant construction of a single-qubit $U$ using quantum teleportation has the form
\begin{eqnarray}
\setlength{\unitlength}{0.7mm}
\begin{array}{c}
\begin{picture}(90,25)
\put(10,0){\makebox(4,4){$\nabla$}}
\put(12.,3.7){\line(0,1){20.3}}
\put(22,18){\circle*{1.5}}
\put(24,17){\tiny{$W_{ij}^\dag$}}
\put(22,0){\line(0,1){24}}
\put(32,0){\line(0,1){24}}
\put(32,6){\circle*{1.5}}
\put(34,5){\tiny{$U$}}
\multiput(8,12)(1,0){28}{\line(1,0){.5}}
\put(22,0){\line(1,0){10}}
\put(12,24){\line(1,0){10}}
\put(36,7){\makebox(14,10){$=\frac 1 2$}}
\put(52,0){\makebox(4,4){$\nabla$}}
\put(54.,3.7){\line(0,1){20.3}}
\put(54,12){\circle*{1.5}}
\put(56,6){\makebox(10,10){\tiny{$U W_{ij}$}}}
\put(66,7){\makebox(14,10){$=\frac 1 2$}}
\put(82,0){\makebox(4,4){$\nabla$}}
\put(84,3.7){\line(0,1){20.3}}
\put(84,18){\circle*{1.5}}
\put(86,13){\makebox(5,10){\tiny{$R_{ij}$}}}
\put(84,9){\circle*{1.5}}
\put(85,4){\makebox(5,10){\tiny{$U$}}}
\end{picture}
\label{gate: one}
\end{array}
\end{eqnarray}
and it is directly associated with the teleportation equation (\ref{tele_single}) and the quantum circuit model in Figure~\ref{fig_tele_single_gate}, see
Subsubsection~\ref{subsection: single-qubit}. Below the dashed line, Alice and Bob prepare the quantum state $|\alpha\rangle\otimes|\Psi_U\rangle$, and above
the dashed line, Alice performs the Bell measurements $|\psi(ij)\rangle\langle\psi(ij)|\otimes 1\!\! 1_2$. After moving the single-qubit gate $W_{ij}^\dag$
from Alice to Bob, we straighten the configuration of the top cap with the bottom cup and obtain the normalization factor $\frac 1 2$. Obviously, Bob has to
perform the local unitary correction operator $R^\dag_{ij}$ to obtain the action of the single-qubit gate $U$ on the unknown state $|\alpha\rangle$, namely,
$U|\alpha\rangle$, as is not shown up in the topological diagram~(\ref{gate: one}).
The topological diagram for the fault-tolerant construction of a two-qubit Clifford gate $CU$ using quantum teleportation is expressed as
\begin{eqnarray}
\setlength{\unitlength}{0.8mm}
\begin{array}{c}
\begin{picture}(150,48)
\put(14,9){\makebox(4,4){$\nabla$}}
\put(11.5,0){\makebox(9,8){$\tiny{|\alpha\rangle}$}}
\put(16,12.5){\line(0,1){32.5}}
\put(26,36){\circle*{1.5}}
\put(18,33){\makebox(6,6){\tiny{$W_{i_1j_1}^\dag$}}}
\put(26,9){\line(0,1){36}}
\multiput(12,27)(1,0){58}{\line(1,0){.5}}
\put(36,9){\line(0,1){7}}
\put(36,20){\line(0,1){25}}
\put(34,16){\line(0,1){4}}
\put(48,16){\line(0,1){4}}
\put(34,16){\line(1,0){14}}
\put(34,20){\line(1,0){14}}
\put(39,16){\makebox(4,4){\tiny{$CU$}}}
\put(26,9){\line(1,0){10}}
\put(16,45){\line(1,0){10}}
\put(46,9){\line(0,1){7}}
\put(46,20){\line(0,1){25}}
\put(46,9){\line(1,0){10}}
\put(56,9){\line(0,1){36}}
\put(56,45){\line(1,0){10}}
\put(66,12.5){\line(0,1){32.5}}
\put(66,36){\circle*{1.5}}
\put(59,34){\makebox(4,4){\tiny{$W_{i_2j_2}^\dag$}}}
\put(64,9){\makebox(4,4){$\nabla$}}
\put(61.5,0){\makebox(9,8){$\tiny{|\beta\rangle}$}}
\put(70,22){\makebox(14,10){$=~\frac 1 4$}}
\put(86,9){\makebox(4,4){$\nabla$}}
\put(82.5,0){\makebox(9,8){$\tiny{|\alpha\rangle}$}}
\put(88,29){\line(0,1){16}}
\put(88,12.5){\line(0,1){12.5}}
\put(96,9){\makebox(4,4){$\nabla$}}
\put(95.5,0){\makebox(9,8){$\tiny{|\beta\rangle}$}}
\put(86,25){\line(0,1){4}}
\put(100,25){\line(0,1){4}}
\put(86,25){\line(1,0){14}}
\put(86,29){\line(1,0){14}}
\put(91,25){\makebox(4,4){\tiny{$CU$}}}
\put(98,29){\line(0,1){16}}
\put(98,12.5){\line(0,1){12.5}}
\put(88,18){\circle*{1.5}}
\put(80,15){\makebox(6,6){\tiny{$W_{i_1j_1}$}}}
\put(98,18){\circle*{1.5}}
\put(102,16){\makebox(4,4){\tiny{$W_{i_2j_2}^T$}}}
\put(106,22){\makebox(14,10){$=~\frac 1 4$}}
\put(124,9){\makebox(4,4){$\nabla$}}
\put(120.5,0){\makebox(9,8){$\tiny{|\alpha\rangle}$}}
\put(126,29){\line(0,1){16}}
\put(126,12.5){\line(0,1){12.5}}
\put(134,9){\makebox(4,4){$\nabla$}}
\put(133.5,0){\makebox(9,8){$\tiny{|\beta\rangle}$}}
\put(124,25){\line(0,1){4}}
\put(138,25){\line(0,1){4}}
\put(124,25){\line(1,0){14}}
\put(124,29){\line(1,0){14}}
\put(129,25){\makebox(4,4){\tiny{$CU$}}}
\put(136,29){\line(0,1){16}}
\put(136,12.5){\line(0,1){12.5}}
\put(126,36){\circle*{1.5}}
\put(121,34){\makebox(4,4){\tiny{$Q$}}}
\put(136,36){\circle*{1.5}}
\put(137,34){\makebox(4,4){\tiny{$P$}}}
\end{picture}
\label{gate: two}
\end{array}
\end{eqnarray}
and it is associated with the teleportation equation (\ref{tele_two}) and the quantum circuit model in Figure~\ref{fig_tele_two_gate}. Below
the dashed line, we prepare the six-qubit quantum state $|\alpha\rangle\otimes|\Psi_{CU}\rangle\otimes|\beta\rangle$, and above the dashed
line, we perform the Bell measurement (\ref{six_bell_measurement}) on such the prepared quantum state. The topological (or straightening)
operation occurs after both moving single-qubit gates $W_{i_1j_1}^\dag$ and $W_{i_2j_2}^\dag$ along the path formed by the top cup states
and bottom cup states and moving the two-qubit gate $CU$ along two vertical lines. The vanishing cap and cup states contribute the
normalization factor $\frac 1 4$. Explicitly, the single-qubit gates $Q$ and $P$ are calculated by the formula (\ref{Q P}).
As a remark, the extended Temperley--Lieb diagrammatical approach to the fault-tolerant construction of quantum gates in teleportation-based
quantum computation appears more intuitive and simpler than other original approaches \cite{GC99,Nielsen03,Leung04}. On the topological
representations, such as (\ref{tele: tl}), (\ref{gate: one}) and (\ref{gate: two}), one is allowed to transport an unknown
quantum state by topological operations as well as move single-qubit or two-qubit gates along relevant configurations.
\subsection{Topological construction of four-qubit entangled states}
We combine the quantum circuit models \cite{NC2011} of the Bell states and the GHZ states with the extended Temperley--Lieb diagrammatical approach
to construct the topological diagrammatical representations of four-qubit entangled states including the $|\Psi_{CNOT}\rangle$ state (\ref{Psi_cnot}),
the $|\Psi_{CNOT}^\uparrow\rangle$ state (\ref{Psi cnot down}) and the $|\Psi_{CZ}\rangle$ state (\ref{Psi_cz}).
\subsubsection{Diagrammatical representations of the Bell state $|\Psi\rangle$ and GHZ state $|\Upsilon\rangle$ }
The EPR state $|\Psi\rangle$ (\ref{EPR}) can be generated by the quantum circuit model in terms of the Hadamard gate $H$ and the $CNOT$ gate,
\begin{equation}
|\Psi\rangle = CNOT_{12} (H\otimes 1\!\! 1_2) |0\rangle \otimes |0\rangle
\end{equation}
with the diagrammatic representation
\begin{eqnarray}
\label{EPR_diag1}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(60,22)
\put(8,4){\makebox(28,10){$|\Psi\rangle=$}}
\put(32,0){\makebox(4,4){$\nabla$}}
\put(34,4){\line(0,1){20}}
\put(34,8){\circle*{2.}}
\put(34,16){\circle*{2.}}
\put(37,6){\makebox(4,4){$H$}}
\put(44,0){\makebox(4,4){$\nabla$}}
\put(46,4){\line(0,1){20}}
\put(34,16){\line(1,0){14}}
\put(46,16){\circle{4}}
\end{picture}
\end{array}
\end{eqnarray}
where the vertical line with $\nabla$ denotes the state $|0\rangle$. Such the configuration of the EPR state $|\Psi\rangle$ is
equivalent to the cup configuration (\ref{TL diag Bell state}) of the $|\Psi\rangle$ state in
the extended Temperley--Lieb diagrammatical approach. The EPR state $|\Psi\rangle$ has the other equivalent quantum circuit model
\begin{equation}
|\Psi\rangle=CNOT_{21}(1\!\!1_2\otimes H)|00\rangle
\end{equation}
with the diagrammatical representation
\begin{eqnarray}
\label{EPR_diag2}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(60,22)
\put(8,4){\makebox(28,10){$|\Psi\rangle=$}}
\put(32,0){\makebox(4,4){$\nabla$}}
\put(44,0){\makebox(4,4){$\nabla$}}
\put(39,6){\makebox(4,4){$H$}}
\put(34,4){\line(0,1){20}}
\put(46,4){\line(0,1){20}}
\put(32,16){\line(1,0){14}}
\put(46,8){\circle*{2.}}
\put(46,16){\circle*{2.}}
\put(34,16){\circle{4}}
\end{picture}
\end{array}
\end{eqnarray}
which is associated with the configuration (\ref{EPR_diag1}) via the mirror symmetry.
The three-qubit GHZ state $|\Upsilon\rangle$ (\ref{Upsilon}) can be generated by the quantum circuit model in terms of
the EPR state $|\Psi\rangle$ and the $CNOT$ gate, expressed as
\begin{equation}
|\Upsilon\rangle
=(CNOT_{21}\otimes 1\!\!1_2)|0\rangle\otimes |\Psi\rangle,
\end{equation}
with the diagrammatical representation
\begin{eqnarray}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(36,18)
\put(0,4){\makebox(16,10){$|\Upsilon\rangle=$}}
\put(18,0){\makebox(4,4){$\nabla$}}
\put(30,0){\line(0,1){18}}
\put(40,0){\line(0,1){18}}
\put(20,4){\line(0,1){14}}
\put(30,0){\line(1,0){10}}
\put(18,12){\line(1,0){12}}
\put(20,12){\circle{4}}
\put(30,12){\circle*{2.}}
\end{picture}
\label{GHZ 1}
\end{array}
\end{eqnarray}
where the cup configuration (\ref{TL diag Bell state}) of the EPR state $|\Psi\rangle$ is exploited.
The $|\Upsilon\rangle$ state also allows the other equivalent quantum circuit model
\begin{equation}
|\Upsilon\rangle=(1\!\! 1_2 \otimes CNOT_{23}) |\Psi\rangle \otimes |0\rangle,
\end{equation}
with the diagrammatical representation
\begin{eqnarray}
\label{GHZ_diag2}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(36,18)
\put(2,4){\makebox(16,10){$|\Upsilon\rangle=$}}
\put(20,0){\line(0,1){18}}
\put(20,0){\line(1,0){10}}
\put(30,0){\line(0,1){18}}
\put(38,0){\makebox(4,4){$\nabla$}}
\put(40,4){\line(0,1){14}}
\put(30,12){\line(1,0){12}}
\put(40,12){\circle{4}}
\put(30,12){\circle*{2.}}
\end{picture}
\label{GHZ 2}
\end{array}
\end{eqnarray}
which can be obtained from the configuration (\ref{GHZ 1}) via the mirror symmetry.
Furthermore, the GHZ state $|\Upsilon\rangle$ with the local action of the Hadamard gates such as $(H\otimes H\otimes H)|\Upsilon\rangle$ has
the diagrammatical representation,
\begin{eqnarray}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(74,25)
\put(4,16){\makebox(4,4){\tiny{$H$}}}
\put(16,16){\makebox(4,4){\tiny{$H$}}}
\put(28,16){\makebox(4,4){\tiny{$H$}}}
\put(0,0){\makebox(4,4){$\nabla$}}
\put(14,0){\line(1,0){12}}
\put(0,10){\line(1,0){14}}
\put(2,4){\line(0,1){20}}
\put(14,0){\line(0,1){24}}
\put(26,0){\line(0,1){24}}
\put(2,18){\circle*{2.}}
\put(14,18){\circle*{2.}}
\put(26,18){\circle*{2.}}
\put(14,10){\circle*{2.}}
\put(2,10){\circle{4}}
\put(34,10){\makebox(4,4){$=$}}
\put(42,0){\makebox(4,4){$\nabla$}}
\put(46,8){\makebox(4,4){\tiny{$H$}}}
\put(56,0){\line(0,1){24}}
\put(68,0){\line(0,1){24}}
\put(44,4){\line(0,1){20}}
\put(56,0){\line(1,0){12}}
\put(44,18){\line(1,0){14}}
\put(44,18){\circle*{2.}}
\put(44,10){\circle*{2.}}
\put(56,18){\circle{4}}
\end{picture}
\label{GHZ 3}
\end{array}
\end{eqnarray}
where the configuration (\ref{GHZ 1}) of the GHZ state $|\Upsilon\rangle$ is used and the formula
\begin{equation}
(H\otimes H) CNOT_{21} (H\otimes H) = CNOT_{12}
\end{equation} is applied. Under the mirror symmetry, the $(H\otimes H\otimes H)|\Upsilon\rangle$ state
has the other equivalent configuration
\begin{eqnarray}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(74,25)
\put(2,0){\line(0,1){24}}
\put(2,18){\circle*{2.}}
\put(4,16){\makebox(4,4){\tiny{$H$}}}
\put(2,0){\line(1,0){12}}
\put(14,0){\line(0,1){24}}
\put(14,18){\circle*{2.}}
\put(16,16){\makebox(4,4){\tiny{$H$}}}
\put(24,0){\makebox(4,4){$\nabla$}}
\put(26,4){\line(0,1){20}}
\put(26,18){\circle*{2.}}
\put(28,16){\makebox(4,4){\tiny{$H$}}}
\put(14,10){\line(1,0){14}}
\put(26,10){\circle{4}}
\put(14,10){\circle*{2.}}
\put(34,10){\makebox(4,4){$=$}}
\put(44,0){\line(0,1){24}}
\put(44,0){\line(1,0){12}}
\put(56,0){\line(0,1){24}}
\put(66,0){\makebox(4,4){$\nabla$}}
\put(68,4){\line(0,1){20}}
\put(54,18){\line(1,0){14}}
\put(68,18){\circle*{2.}}
\put(56,18){\circle{4}}
\put(68,10){\circle*{2.}}
\put(70,8){\makebox(4,4){\tiny{$H$}}}
\end{picture}
\label{GHZ 4}
\end{array}
\end{eqnarray}
in which the configuration (\ref{GHZ 2}) of the GHZ state $|\Upsilon\rangle$ is exploited.
\subsubsection{Topological construction of the four-qubit entangled state $|\Psi_{CNOT}\rangle$ }
\label{topo_construction_four_CNOT}
The four-qubit entangled state $|\Psi_{CNOT}\rangle$ (\ref{Psi_cnot}) has the extended Temperley--Lieb diagrammatical configuration, expressed as
a $CNOT$ gate connecting two cup configurations,
\begin{eqnarray}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(55,14)
\put(-4,2){\makebox(16,10){$|\Psi_{CNOT}\rangle=$}}
\put(22,0){\line(0,1){14}}
\put(32,0){\line(0,1){14}}
\put(22,0){\line(1,0){10}}
\put(32,7){\line(1,0){12}}
\put(32,7){\circle*{2.}}
\put(42,0){\line(0,1){14}}
\put(52,0){\line(0,1){14}}
\put(42,0){\line(1,0){10}}
\put(42,7){\circle{4.}}
\end{picture}
\label{Psi_cnot_diag}
\end{array}
\end{eqnarray}
where the second qubit is the control qubit and the fifth qubit is the target qubit and both the third and fourth qubits are not explicitly shown up.
Following the strategy of constructing $|\Psi_{CNOT}\rangle$ (\ref{Psi_cnot}) in Subsubsection~\ref{construction Psi CNOT},
we perform the Bell measurements (\ref{Bell_measurement_CNOT}) on both the third and fourth qubits of the six-qubit state $|\Upsilon\rangle \otimes (H\otimes H \otimes H) |\Upsilon\rangle$. Using the diagrammatical representation (\ref{GHZ 1}) of the GHZ state $|\Upsilon\rangle$ and the diagrammatical representation (\ref{GHZ 3}) of the
state $(H\otimes H \otimes H)|\Upsilon\rangle$ and the diagrammatical representation (\ref{TL Bell measurement}) of the Bell measurements (\ref{Bell_measurement_CNOT}),
we have the extended Temperley--Lieb diagrammatical configuration for the construction of the $|\Psi_{CNOT}\rangle$ state (\ref{Psi_cnot}),
\begin{equation}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(154,50)
\put(8,0){\makebox(4,4){$\nabla$}}
\put(48,8){\makebox(4,4){\tiny{$H$}}}
\put(44,0){\makebox(4,4){$\nabla$}}
\put(38,33){\makebox(6,6){\tiny{$W_{ij}^\dag$}}}
\put(75,20){\makebox(10,8){$= \frac 1 2$}}
\multiput(8,24)(1,0){65}{\line(1,0){.5}}
\put(10,4){\line(0,1){44}}
\put(22,0){\line(0,1){48}}
\put(34,0){\line(0,1){48}}
\put(46,4){\line(0,1){44}}
\put(58,0){\line(0,1){48}}
\put(70,0){\line(0,1){48}}
\put(8,18){\line(1,0){14}}
\put(22,0){\line(1,0){12}}
\put(34,48){\line(1,0){12}}
\put(46,18){\line(1,0){14}}
\put(58,0){\line(1,0){12}}
\put(46,18){\circle*{2.}}
\put(46,10){\circle*{2.}}
\put(22,18){\circle*{2.}}
\put(46,36){\circle*{2.}}
\put(10,18){\circle{4}}
\put(58,18){\circle{4}}
\put(88,0){\makebox(4,4){$\nabla$}}
\put(104,8){\makebox(4,4){\tiny{$H$}}}
\put(100,0){\makebox(4,4){$\nabla$}}
\put(105,41){\makebox(4,4){\tiny{$Z^j$}}}
\put(93,35){\makebox(4,4){\tiny{$X^i$}}}
\put(105,35){\makebox(4,4){\tiny{$X^i$}}}
\put(90,4){\line(0,1){44}}
\put(102,4){\line(0,1){44}}
\put(138,0){\line(0,1){48}}
\put(150,0){\line(0,1){48}}
\put(138,0){\line(1,0){12}}
\put(102,18){\line(1,0){38}}
\put(88,30){\line(1,0){14}}
\put(138,18){\circle{4}}
\put(90,30){\circle{4}}
\put(102,18){\circle*{2.}}
\put(102,10){\circle*{2.}}
\put(102,30){\circle*{2.}}
\put(102,36){\circle*{2.}}
\put(102,42){\circle*{2.}}
\put(90,36){\circle*{2.}}
\end{picture}
\label{fig_psi_cnot_up10}
\end{array}
\end{equation}
which is associated with the teleportation equation~(\ref{tele_cnot1}) and the quantum circuit model in Figure~\ref{fig_psi_cnot1}, and in which below
the dashed line is the prepared six-qubit state and above the dashed line is the Bell measurements.
Let us perform a series of diagrammatical operations on the diagram~(\ref{fig_psi_cnot_up10}). First, move the single-qubit gate $W_{ij}^\dag$ from
the fourth qubit to the second qubit along the given path. Second, continue to move such the $W_{ij}^\dag$ gate across the $CNOT_{21}$ gate with
the formula
\begin{equation}
CNOT_{21}(1\!\!1_2\otimes W^\dag_{ij})CNOT_{21}=Z_2^jX_1^iX_2^i,
\end{equation}
to obtain the single-qubit gates $Z_2^jX_1^iX_2^i$ acting on the first and second qubits. Third, straighten the configuration of the bottom cup with
the top cap between the second qubit and the fourth qubit to derive the normalization factor $\frac 1 2$ and meanwhile to obtain the $CNOT_{25}$ gate
and the Hadamard gate $H_2$ respectively from the original $CNOT_{45}$ gate and the Hadamard gate $H_4$
Obviously, the $CNOT_{21}$ gate commutes with the $CNOT_{25}$ gate, and applying such the fact on the diagram~(\ref{fig_psi_cnot_up10}) brings about
the topological diagrammatical representation
\begin{equation}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(154,50)
\put(8,0){\makebox(4,4){$\nabla$}}
\put(20,0){\makebox(4,4){$\nabla$}}
\put(24,8){\makebox(4,4){\tiny{$H$}}}
\put(25,41){\makebox(4,4){\tiny{$Z^j$}}}
\put(13,35){\makebox(4,4){\tiny{$X^i$}}}
\put(25,35){\makebox(4,4){\tiny{$X^i$}}}
\put(10,4){\line(0,1){44}}
\put(22,4){\line(0,1){44}}
\put(58,0){\line(0,1){48}}
\put(70,0){\line(0,1){48}}
\put(58,0){\line(1,0){12}}
\put(22,30){\line(1,0){38}}
\put(8,18){\line(1,0){14}}
\put(22,10){\circle*{2.}}
\put(22,18){\circle*{2.}}
\put(22,30){\circle*{2.}}
\put(22,36){\circle*{2.}}
\put(22,42){\circle*{2.}}
\put(10,36){\circle*{2.}}
\put(58,30){\circle{4}}
\put(10,18){\circle{4}}
\put(76,22){\makebox(4,4){$=$}}
\put(101,41){\makebox(4,4){\tiny{$Z^j$}}}
\put(89,35){\makebox(4,4){\tiny{$X^i$}}}
\put(101,35){\makebox(4,4){\tiny{$X^i$}}}
\put(86,0){\line(0,1){48}}
\put(98,0){\line(0,1){48}}
\put(134,0){\line(0,1){48}}
\put(146,0){\line(0,1){48}}
\put(86,0){\line(1,0){12}}
\put(98,30){\line(1,0){38}}
\put(134,0){\line(1,0){12}}
\put(98,30){\circle*{2.}}
\put(98,36){\circle*{2.}}
\put(98,42){\circle*{2.}}
\put(86,36){\circle*{2.}}
\put(134,30){\circle{4}}
\end{picture}
\label{fig_psi_cnot_up11}
\end{array}
\end{equation}
in which the diagrammatic representation (\ref{EPR_diag2}) of the EPR state $|\Psi\rangle$ is used. With both classical communication and unitary correction, therefore,
the extended Temperley--Lieb diagrammatical representation (\ref{Psi_cnot_diag}) of the four-qubit entangled state $|\Psi_{CNOT}\rangle$ can be exactly prepared
in the diagrammatical approach.
Now let us derive the teleportation equation (\ref{tele_cnot2}) in the topological diagrammatical approach. On the diagram~(\ref{fig_psi_cnot_up10}),
we replace the diagrammatical representation (\ref{GHZ 1}) of the GHZ state $|\Upsilon\rangle$ with its another diagrammatical representation (\ref{GHZ 2})
and replace the diagrammatical representation (\ref{GHZ 3}) of the $H_1H_2H_3|\Upsilon\rangle$ state with its another diagrammatical representation (\ref{GHZ 4})
so that we have the other equivalent topological diagrammatical representation,
\begin{equation}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(154,50)
\put(30,0){\makebox(4,4){$\nabla$}}
\put(66,0){\makebox(4,4){$\nabla$}}
\put(70,8){\makebox(4,4){\tiny{$H$}}}
\put(36,33){\makebox(6,6){\tiny{$W_{ij}^\dag$}}}
\multiput(6,24)(1,0){65}{\line(1,0){.5}}
\put(8,0){\line(0,1){48}}
\put(20,0){\line(0,1){48}}
\put(32,4){\line(0,1){44}}
\put(44,0){\line(0,1){48}}
\put(56,0){\line(0,1){48}}
\put(68,4){\line(0,1){44}}
\put(8,0){\line(1,0){12}}
\put(20,18){\line(1,0){14}}
\put(32,48){\line(1,0){12}}
\put(44,0){\line(1,0){12}}
\put(54,18){\line(1,0){14}}
\put(68,18){\circle*{2.}}
\put(68,10){\circle*{2.}}
\put(20,18){\circle*{2.}}
\put(44,36){\circle*{2.}}
\put(32,18){\circle{4}}
\put(56,18){\circle{4}}
\put(73,20){\makebox(10,8){$= \frac 1 2$}}
\put(139,35){\makebox(4,4){\tiny{$Z^j$}}}
\put(139,41){\makebox(4,4){\tiny{$X^i$}}}
\put(151,35){\makebox(4,4){\tiny{$Z^j$}}}
\put(88,0){\line(0,1){48}}
\put(100,0){\line(0,1){48}}
\put(136,0){\line(0,1){48}}
\put(148,0){\line(0,1){48}}
\put(88,0){\line(1,0){12}}
\put(100,30){\line(1,0){38}}
\put(136,0){\line(1,0){12}}
\put(136,30){\circle{4}}
\put(100,30){\circle*{2.}}
\put(136,36){\circle*{2.}}
\put(136,42){\circle*{2.}}
\put(148,36){\circle*{2.}}
\end{picture}
\label{fig_psi_cnot_up20}
\end{array}
\end{equation}
in which we firstly move the single-qubit gate $W_{ij}^\dag$ from the fourth qubit to the fifth qubit and then across $CNOT_{65}$ gate with the formula
\begin{equation}
CNOT_{65}(W_{ij}\otimes 1\!\!1_2)CNOT_{65}=X_5^iZ_5^jZ_6^j,
\end{equation}
to generate the single-qubit gate $X_5^i Z_5^j Z_6^j$, and secondly apply the commutative relation of the $CNOT_{25}$ gate and $CNOT_{65}$ and
the diagrammatic representation (\ref{EPR_diag2}) of the EPR state $|\Psi\rangle$. As a result, the diagram~(\ref{fig_psi_cnot_up20}) presents a kind of
proof for the teleportation equation (\ref{tele_cnot2}).
\subsubsection{Topological constructions of four-qubit entangled states $|\Psi_{CNOT}^\uparrow\rangle$ and $|\Psi_{CZ}\rangle$ }
The topological constructions of the four-qubit entangled states $|\Psi_{CNOT}^\uparrow\rangle$ (\ref{Psi cnot down})
and $|\Psi_{CZ}\rangle$ (\ref{Psi_cz}) are shown up in Appendix~\ref{construction cnot down} and Appendix~\ref{construction Psi CZ} respectively.
\section{The Yang--Baxter gate and its extended Temperley--Lieb diagrammatical representation}
\label{section the YBG and TL}
In this section, we consider two special types of the Yang--Baxter gates derived in Appendix A as well as present their associated extended
Temperley--Lieb diagrammatical configurations, and then apply the special type II Yang--Baxter gates to teleportation-based quantum computation
in Section~\ref{YBG approach to tele based QC}. Note that a brief study on the algebraic properties of the special type I Yang--Baxter gates
is made in Appendix B. In the first place, the extended Temperley--Lieb diagrammatical representation of a two-qubit quantum gate will be discussed.
\subsection{The extended Temperley-Lieb diagrammatical representation of a two-qubit gate}
The conventional Temperley--Lieb diagram consists of the configuration of a pair of cup and cap \cite{Kauffman02}, see Appendix~\ref{Def TL} for
the diagrammatical representation of the Temperley--Lieb algebra. As we have introduced in Section~\ref{TL diagrammatical for tele based quantum computation}, the extended Temperley--Lieb diagram permits the action of single-qubit gates, namely, it includes the following typical configuration,
\begin{equation}
\label{extended TL generator}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(65,30)
\put(62,18){\line(0,1){12}}
\put(52,18){\line(0,1){12}}
\put(52,18){\line(1,0){10}}
\put(62,24){\circle*{2.}}
\put(64,23){\tiny{$M$}}
\put(-3,14){$|\Psi_M\rangle \langle\Psi_N|=$}
\put(52,0){\line(0,1){12}}
\put(62,0){\line(0,1){12}}
\put(52,12){\line(1,0){10}}
\put(62,6){\circle*{2.}}
\put(64,5){\tiny{$N^\dag$}}
\end{picture}
\end{array}
\end{equation}
which represents the EPR state projector $|\Psi\rangle\langle\Psi|$ with the local action of single-qubit gates $M$ and $N^\dag$. Obviously, when $M=N=1\!\! 1_2$, such
the configuration (\ref{extended TL generator}) becomes the standard Temperley--Lieb configuration in Appendix~\ref{Def TL}.
In quantum information and computation \cite{NC2011}, the two-qubit Hilbert space $\mathcal{H}_2\otimes \mathcal{H}_2$ has the two types of the orthonormal bases:
the one is denoted by the product basis $|ij\rangle$, the other is denoted by the Bell basis $|\psi(ij)\rangle$, and the unitary basis transformation matrix $T$
between these two bases satisfies $T|ij\rangle=|\psi(ij)\rangle$ and has the form,
\begin{equation}
\label{basis_transformation}
T=\frac 1 {\sqrt 2}\left(
\begin{array}{cccc}
1 & 1 & 0 & 0 \\
0 & 0 & 1 & 1 \\
0 & 0 & 1 & -1 \\
1 & -1 & 0 & 0 \\
\end{array}
\right).
\end{equation}
In terms of the product base $|ij\rangle\langle kl|$, a two-qubit quantum gate $G$ as a $4\times 4$ unitary matrix has an
expression given by
\begin{equation}
G=\sum_{i,j,k,l=0}^1 G_{ij,kl}|ij\rangle\langle kl|,
\end{equation}
and in terms of the Bell base $|\psi(ij)\rangle\langle\psi(kl)|$, it has another form
\begin{equation}
\label{two-qubit gate TL representation}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(95,30)
\put(6,14){$G=\sum_{i,j,k,l=0}^1 \tilde{G}_{ij,kl}$}
\put(82,18){\line(0,1){12}}
\put(72,18){\line(0,1){12}}
\put(72,18){\line(1,0){10}}
\put(82,24){\circle*{2.}}
\put(84,23){\tiny{$X^iZ^j$}}
\put(72,0){\line(0,1){12}}
\put(82,0){\line(0,1){12}}
\put(72,12){\line(1,0){10}}
\put(82,6){\circle*{2.}}
\put(84,5){\tiny{$Z^lX^k$}}
\end{picture}
\end{array}
\end{equation}
in the extended Temperley--Lieb diagrammatical approach. Applying the basis transformation matrix $T$ (\ref{basis_transformation}),
the coefficient matrix $\tilde{G}=(\tilde{G}_{ij,kl})$ can be obtained from
\begin{equation}
\label{trans A}
\tilde{G}=T^\dag G T.
\end{equation}
For example, the two-qubit identity gate $1\!\!1_4$ has an expansion of the Bell basis (\ref{Bell states}),
\begin{equation}
1\!\!1_4=\sum_{i,j=0}^1|\psi(ij)\rangle\langle\psi(ij)|,
\end{equation}
which represents the completeness relation defining the Bell basis, so the two-qubit identity
gate $1\!\!1_4$ has the extended Temperley--Lieb diagrammatic representation,
\begin{equation}
\label{TL_identity}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(170,30)
\put(40,14){$=$}
\put(22,0){\line(0,1){30}}
\put(32,0){\line(0,1){30}}
\put(62,18){\line(0,1){12}}
\put(52,18){\line(0,1){12}}
\put(52,18){\line(1,0){10}}
\put(52,0){\line(0,1){12}}
\put(62,0){\line(0,1){12}}
\put(52,12){\line(1,0){10}}
\put(70,14){$+$}
\put(92,18){\line(0,1){12}}
\put(82,18){\line(0,1){12}}
\put(82,18){\line(1,0){10}}
\put(92,24){\circle*{2.}}
\put(94,23){\tiny{$Z$}}
\put(82,0){\line(0,1){12}}
\put(92,0){\line(0,1){12}}
\put(82,12){\line(1,0){10}}
\put(92,6){\circle*{2.}}
\put(94,5){\tiny{$Z$}}
\put(100,14){$+$}
\put(122,18){\line(0,1){12}}
\put(112,18){\line(0,1){12}}
\put(112,18){\line(1,0){10}}
\put(122,24){\circle*{2.}}
\put(124,23){\tiny{$X$}}
\put(112,0){\line(0,1){12}}
\put(122,0){\line(0,1){12}}
\put(112,12){\line(1,0){10}}
\put(122,6){\circle*{2.}}
\put(124,5){\tiny{$X$}}
\put(130,14){$+$}
\put(152,18){\line(0,1){12}}
\put(142,18){\line(0,1){12}}
\put(142,18){\line(1,0){10}}
\put(152,24){\circle*{2.}}
\put(154,23){\tiny{$XZ$}}
\put(142,0){\line(0,1){12}}
\put(152,0){\line(0,1){12}}
\put(142,12){\line(1,0){10}}
\put(152,6){\circle*{2.}}
\put(154,5){\tiny{$ZX$}}
\end{picture}
\end{array},
\end{equation}
where, in the conventional topological viewpoint \cite{Kauffman02}, the two-qubit identity gate $1\!\!1_4$ is depicted as two parallel vertical lines.
For another example, let us consider the extended Temperley--Lieb diagrammatical configuration of the $CZ$ gate (\ref{CZ gate}), which is
\begin{equation}
\label{Extended_TL_CZ}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(150,30)
\put(7,14){$CZ=$}
\put(42,18){\line(0,1){12}}
\put(32,18){\line(0,1){12}}
\put(32,18){\line(1,0){10}}
\put(32,0){\line(0,1){12}}
\put(42,0){\line(0,1){12}}
\put(32,12){\line(1,0){10}}
\put(42,6){\circle*{2.}}
\put(44,5){\tiny{$Z$}}
\put(50,14){$+$}
\put(72,18){\line(0,1){12}}
\put(62,18){\line(0,1){12}}
\put(62,18){\line(1,0){10}}
\put(72,24){\circle*{2.}}
\put(74,23){\tiny{$Z$}}
\put(62,0){\line(0,1){12}}
\put(72,0){\line(0,1){12}}
\put(62,12){\line(1,0){10}}
\put(80,14){$+$}
\put(102,18){\line(0,1){12}}
\put(92,18){\line(0,1){12}}
\put(92,18){\line(1,0){10}}
\put(102,24){\circle*{2.}}
\put(104,23){\tiny{$X$}}
\put(92,0){\line(0,1){12}}
\put(102,0){\line(0,1){12}}
\put(92,12){\line(1,0){10}}
\put(102,6){\circle*{2.}}
\put(104,5){\tiny{$X$}}
\put(110,14){$+$}
\put(132,18){\line(0,1){12}}
\put(122,18){\line(0,1){12}}
\put(122,18){\line(1,0){10}}
\put(132,24){\circle*{2.}}
\put(134,23){\tiny{$XZ$}}
\put(122,0){\line(0,1){12}}
\put(132,0){\line(0,1){12}}
\put(122,12){\line(1,0){10}}
\put(132,6){\circle*{2.}}
\put(134,5){\tiny{$ZX$}}
\end{picture}
\end{array},
\end{equation}
where the first and second configurations from the left handside are explicitly beyond the standard Temperley--Lieb diagrammatical representation in Appendix~\ref{Def TL}.
Note that for a given two-qubit gate $G$, its extended Tempereley--Lieb diagrammatical representation is not fixed and depends on the choices of single-qubit gates acting on
the cup or cap configurations. For example, the extended Temperley--Lieb diagrammatic representation of the $CZ$ gate (\ref{CZ gate}) can not be (\ref{Extended_TL_CZ}) when
single-qubit gates on the configurations are not the Pauli gates.
\subsection{Special type I Yang--Baxter gate and its extended Temperley-Lieb diagrammatical representation}
\label{type_I_YBE_TL}
The special type I Yang--Baxter gate is a typical example for the Yang--Baxter gate via the Temperley--Lieb algebra, and it is derived in Appendix~\ref{type_I_YBE}.
The main reason that we study it in this paper is that it is directly associated with the Bell states (\ref{Bell states}), and we denote the special type I Yang--Baxter gate
by $R(ij)$, which is formulated as
\begin{equation}
\label{R ij}
R(i j)=1\!\!1_4-2|\psi(i j)\rangle\langle \psi(i j)|.
\end{equation}
As an example, for $i=j=1$, we have the Yang--Baxter gate $P \equiv R(11)$, given by
\begin{equation}
\label{permutation gate}
P=\left(
\begin{array}{cccc}
1 & 0 & 0 & 0 \\
0 & 0 & 1 & 0 \\
0 & 1 & 0 & 0 \\
0 & 0 & 0 & 1 \\
\end{array}
\right)
\end{equation}
which is the Permutation gate (or the Swap gate) $P$ in quantum computation \cite{NC2011}. Note that all of the special type I Yang--Baxter gates $R(ij)$
are permutation-like quantum gates satisfying $R(ij)R(ij)=1\!\!1_4$, and a further research has been done in Appendix~\ref{teleportation_swapping_operator}.
The special type I Yang--Baxter gate $R(ij)$ in the extended Temperley--Lieb diagrammatical approach is shown as
\begin{equation}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(120,30)
\put(0,14){$R(ij)=\sum_{k,l=0}^1(1-2\delta_{i,k}\delta_{j,l})$}
\put(102,18){\line(0,1){12}}
\put(92,18){\line(0,1){12}}
\put(92,18){\line(1,0){10}}
\put(102,24){\circle*{2.}}
\put(104,23){\tiny{$X^kZ^l$}}
\put(92,0){\line(0,1){12}}
\put(102,0){\line(0,1){12}}
\put(92,12){\line(1,0){10}}
\put(102,6){\circle*{2.}}
\put(104,5){\tiny{$Z^lX^k$}}
\end{picture}
\end{array}
\end{equation}
where the symbol $\delta$ denotes the Kronecker delta function. For example, the Yang--Baxter gate $P=R(11)$ has the extended Temperley--Lieb
diagrammatical representation,
\begin{equation}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(140,30)
\put(0,14){$P=$}
\put(32,18){\line(0,1){12}}
\put(22,18){\line(0,1){12}}
\put(22,18){\line(1,0){10}}
\put(22,0){\line(0,1){12}}
\put(32,0){\line(0,1){12}}
\put(22,12){\line(1,0){10}}
\put(40,14){$+$}
\put(62,18){\line(0,1){12}}
\put(52,18){\line(0,1){12}}
\put(52,18){\line(1,0){10}}
\put(62,24){\circle*{2.}}
\put(64,23){\tiny{$X$}}
\put(52,0){\line(0,1){12}}
\put(62,0){\line(0,1){12}}
\put(52,12){\line(1,0){10}}
\put(62,6){\circle*{2.}}
\put(64,5){\tiny{$X$}}
\put(70,14){$+$}
\put(92,18){\line(0,1){12}}
\put(82,18){\line(0,1){12}}
\put(82,18){\line(1,0){10}}
\put(92,24){\circle*{2.}}
\put(94,23){\tiny{$Z$}}
\put(82,0){\line(0,1){12}}
\put(92,0){\line(0,1){12}}
\put(82,12){\line(1,0){10}}
\put(92,6){\circle*{2.}}
\put(94,5){\tiny{$Z$}}
\put(100,14){$-$}
\put(122,18){\line(0,1){12}}
\put(112,18){\line(0,1){12}}
\put(112,18){\line(1,0){10}}
\put(122,24){\circle*{2.}}
\put(124,23){\tiny{$XZ$}}
\put(112,0){\line(0,1){12}}
\put(122,0){\line(0,1){12}}
\put(112,12){\line(1,0){10}}
\put(122,6){\circle*{2.}}
\put(124,5){\tiny{$ZX$}}
\end{picture}
\end{array}.
\end{equation}
Note that the special type I Yang--Baxter gate $R(ij)$ will not be applied to our study on teleportation-based quantum computation in
Section~\ref{YBG approach to tele based QC}, see Appendix~\ref{teleportation_swapping_operator} for a relevant interpretation. We
present them here because they have a simplest realization in the extended Temperley--Lieb diagrammatical approach.
\subsection{Special type II Yang--Baxter gate and its extended Temperley-Lieb diagrammatical representation}
\label{type_II_YBE TL}
The special type II Yang--Baxter gates $B(\epsilon,\eta)$, or equivalently denoted as $B_{\epsilon,\eta}$,
with $\epsilon=\pm1$ and $\eta=\pm1$, have the form
\begin{equation}
\label{B Bell transform}
B(\epsilon,\eta)=\frac 1 {\sqrt 2}\left(
\begin{array}{cccc}
1 & 0 & 0 & \eta \\
0 & 1 & \epsilon & 0 \\
0 & -\epsilon & 1 & 0 \\
-\eta & 0 & 0 & 1 \\
\end{array}
\right),
\end{equation}
which are derived in Appendix~\ref{type_II_YBE} as examples for the Yang--Baxter gates generated by the Temperley--Lieb algebra.
These four Yang--Baxter gates $B(\epsilon,\eta)$ are related to one another by the relations
\begin{equation}
\label{Four B relation}
B(\epsilon,\eta)=B^\dag(-\epsilon,-\eta),\quad B(\epsilon,\eta)=P B(-\epsilon,\eta)P,
\end{equation}
where $P$ is the Permutation gate (\ref{permutation gate}), and the inverse of such the Yang--Baxter gate $B(\epsilon,\eta)$ is
related to itself with the aid of the Pauli $Z$ gate,
expressed as
\begin{equation}
\label{B and B inverse relation}
B(\epsilon,\eta)=Z_1B^\dag(\epsilon,\eta)Z_1=Z_2B^\dag(\epsilon,\eta) Z_2.
\end{equation}
The special type II Yang--Baxter gates $B(\epsilon,\eta)$ are to be applied to our study on the reformulation of the teleportaiton-based quantum
computation in Section~\ref{YBG approach to tele based QC}, because they can generate the Bell states (\ref{Bell states}) from the product basis,
see \cite{Dye03, KL04,ZKG04}. For example, when the parameters $\epsilon=1$ and $\eta=1$, we denote the Yang--Baxter gate $B$\footnote{The Yang--Baxter
gate $B$ in this paper is denoted as the Yang--Baxter gate $B'$ in the authors' another paper \cite{ZZ14}.} as an
abbreviation of the notation $B(1,1)$,
\begin{equation}
\label{YBG B}
B=\frac{1}{\sqrt 2}\left(
\begin{array}{cccc}
1 & 0 & 0 & 1 \\
0 & 1 & 1 & 0 \\
0 & -1 & 1 & 0 \\
-1 & 0 & 0 & 1 \\
\end{array}
\right)
\end{equation}
which gives rise to all four Bell states from the product basis in the way
\begin{equation}
\label{example B Bell transform}
B\left(
\begin{array}{c}
|00\rangle \\
|01\rangle \\
|10\rangle \\
|11\rangle \\
\end{array}
\right)=\left(
\begin{array}{c}
|\psi(01)\rangle \\
|\psi(11)\rangle \\
|\psi(10)\rangle \\
|\psi(00)\rangle \\
\end{array}
\right).
\end{equation}
We can directly read the extended Temperley--Lieb configuration of the Yang--Baxter gate $B(\epsilon,\eta)$, shown as
\begin{equation}
\label{TL B Bell transform}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(130,40)
\put(-31,-2){\footnotesize{$B(\epsilon,\eta)$}}
\put(-31,29){\line(0,1){6}}
\put(-31,11){\line(0,-1){6}}
\put(-13,29){\line(0,1){6}}
\put(-13,11){\line(0,-1){6}}
\put(-34,8){\line(0,1){24}}
\put(-10,8){\line(0,1){24}}
\put(-34,8){\line(1,0){24}}
\put(-34,32){\line(1,0){24}}
\put(-31,29){\line(1,-1){18}}
\put(-31,11){\line(1,1){7.3}}
\put(-13,29){\line(-1,-1){7.3}}
\put(-3,19){$=\frac 1 {\sqrt 2}($}
\put(32,5){\line(0,1){30}}
\put(22,5){\line(0,1){30}}
\put(38,19){$+\eta$}
\put(62,23){\line(0,1){12}}
\put(52,23){\line(0,1){12}}
\put(52,23){\line(1,0){10}}
\put(62,29){\circle*{2.}}
\put(64,28){\tiny{$Z$}}
\put(52,5){\line(0,1){12}}
\put(62,5){\line(0,1){12}}
\put(52,17){\line(1,0){10}}
\put(68,19){$-\eta$}
\put(92,23){\line(0,1){12}}
\put(82,23){\line(0,1){12}}
\put(82,23){\line(1,0){10}}
\put(82,5){\line(0,1){12}}
\put(92,5){\line(0,1){12}}
\put(82,17){\line(1,0){10}}
\put(92,11){\circle*{2.}}
\put(94,10){\tiny{$Z$}}
\put(98,19){$-\epsilon$}
\put(122,23){\line(0,1){12}}
\put(112,23){\line(0,1){12}}
\put(112,23){\line(1,0){10}}
\put(122,29){\circle*{2.}}
\put(124,28){\tiny{$X$}}
\put(112,5){\line(0,1){12}}
\put(122,5){\line(0,1){12}}
\put(112,17){\line(1,0){10}}
\put(122,11){\circle*{2.}}
\put(124,10){\tiny{$ZX$}}
\put(128,19){$+\epsilon$}
\put(152,23){\line(0,1){12}}
\put(142,23){\line(0,1){12}}
\put(142,23){\line(1,0){10}}
\put(152,29){\circle*{2.}}
\put(154,28){\tiny{$XZ$}}
\put(142,5){\line(0,1){12}}
\put(152,5){\line(0,1){12}}
\put(142,17){\line(1,0){10}}
\put(152,11){\circle*{2.}}
\put(154,10){\tiny{$X$}}
\put(164,19){$)$}
\end{picture}
\end{array}
\end{equation}
where the over-crossing diagram represents the braiding feature \cite{Kauffman02} of the Yang--Baxter gate $B(\epsilon,\eta)$ (\ref{B Bell transform}) and the box
around it marks the feature that it can be regarded as a two-qubit quantum gate, and two parallel vertical lines represent the two-qubit identity gate (\ref{TL_identity}).
To maintain the diagrammatical representations consistent, the braiding configuration (the over-crossing configuration) has the same acting direction as the other
extended Temperley--Lieb diagrammatic representations, namely, it is read from the bottom to the top.
\section{The Yang--Baxter gate approach to teleportation-based quantum computation}
\label{YBG approach to tele based QC}
In this section, we apply the special type II Yang--Baxter gates $B(\epsilon,\eta)$ (\ref{B Bell transform}), derived in Appendix A
and presented in Section~\ref{section the YBG and TL} with their associated extended Temperley--Lieb diagrammatical configurations (\ref{TL B Bell transform}),
to teleportation-based quantum computation. First of all, such the special type II Yang--Baxter gates $B(\epsilon,\eta)$ are found to be a type of examples
for the Bell transform defined in \cite{ZZ14}, where teleportation-based quantum computation using the Bell transform has been explored in detail. Note that
the Yang--Baxter gate approach to teleportation-based quantum computation admits an interpretation in the extended Temperley--Lieb diagrammatical approach,
see Section~\ref{relationship_YBE_TL}.
\subsection{The Yang--Baxter gate $B(\epsilon,\eta)$ as the Bell transform and Clifford gate}
We recognize the special type II Yang--Baxter gate $B(\epsilon,\eta)$ (\ref{B Bell transform}) as the Bell transform \cite{ZZ14}, and especially
verify them as Clifford gates \cite{NC2011, Gottesman97} in three equivalent approaches.
A type of the Bell transform \cite{ZZ14} is defined as a two-qubit quantum gate capable of generating the four Bell states (\ref{Bell states})
from the product states modulo global phase factors. For examples, the special type II Yang--Baxter gates $B(\epsilon,\eta)$,
denoted by $B_{\epsilon,\eta}$ in this section, are the Bell transform because they produce the Bell states in the way,
\begin{eqnarray}
&&B_{-1,1}|ij\rangle = (-1)^{i\cdot(j+1)}|\psi(i+j,j+1)\rangle,\\
&&B_{1,-1}|ij\rangle = (-1)^{i\cdot j}|\psi(i+j,j)\rangle,\\
\label{B Bell transform index}
&&B_{1,1}|ij\rangle = |\psi(i+j,i+1)\rangle,\\
\label{B inverse Bell transform index}
&&B_{-1,-1}|ij\rangle = (-1)^i|\psi(i+j,i)\rangle,
\end{eqnarray}
in which the addition is the binary addition and the multiplication is the logical AND operation. Obviously, the global phase factors of such the Bell states
generated by the $B_{\epsilon,\eta}$ gates are $\pm1$, so the special type II Yang--Baxter gates $B_{\epsilon,\eta}$ are Clifford gates \cite{NC2011, Gottesman97},
because the global phase factors associated with Clifford gates are allowed to be only $\pm 1$, $\pm i$ \cite{ZZ14}.
According to discussed in Subsubsection~\ref{clifford_gate_fault_tolerant}, the fact that the special type II Yang--Baxter gates $B_{\epsilon,\eta}$ are Clifford gates
is crucial in the fault-tolerant construction of universal quantum gate set using quantum teleportation in the following subsections.
We will verify the $B_{\epsilon,\eta}$ gates as Clifford gates in another two ways. A Clifford gate can be expressed as a tensor product of elementary Clifford gates,
namely the Hadamard gate $H$ (\ref{Hadamard gate}), the phase gate $S$ (\ref{phase gate}) and the $CNOT$ gate (\ref{CNOT gate}), so we reformulate the $B_{-1,1}$ gate as
\begin{equation}
B_{-1,1}=CNOT_{21}(1\!\!1_2\otimes ZH) CNOT_{21},
\end{equation}
with $Z=S^2$. With the algebraic relations (\ref{Four B relation}) among the $B_{\epsilon,\eta}$ gates, for examples, the $B_{1,-1}$ gate is the inverse of the $B_{-1,1}$ gate and the $B_{1,1}$ gate is the conjugation of the $B_{-1,1}$ gate under the Permutation gate, we can decompose all the $B_{\epsilon,\eta}$ gates as
tensor products of elementary Clifford gates.
\begin{table}
\begin{center}
\footnotesize
\begin{tabular}{c|c|c}
\hline\hline
Operation & Input & Output \\ \hline
& $X_1$ & $X_1$ \\
& $X_2$ & $X_1Z_2$ \\
\raisebox{1.8ex}[0pt]{$B_{-1,1}$} & $Z_1$ & $-Y_1Y_2$ \\
& $Z_2$ & $-X_1X_2$ \\ \hline
& $X_1$ & $X_1$ \\
& $X_2$ & $-X_1Z_2$ \\
\raisebox{1.8ex}[0pt]{$B_{1,-1}$} & $Z_1$ & $Y_1Y_2$ \\
& $Z_2$ & $X_1X_2$ \\ \hline
& $X_1$ & $Z_1X_2$ \\
& $X_2$ & $X_2$ \\
\raisebox{1.8ex}[0pt]{$B_{1,1}$} & $Z_1$ & $-X_1X_2$ \\
& $Z_2$ & $-Y_1Y_2$ \\ \hline
& $X_1$ & $-Z_1X_2$ \\
& $X_2$ & $X_2$ \\
\raisebox{1.8ex}[0pt]{$B_{-1,-1}$} & $Z_1$ & $X_1X_2$ \\
& $Z_2$ & $Y_1Y_2$ \\ \hline\hline
\end{tabular}
\caption{\label{clifford_YBG} Transformation properties of elements of the Pauli group ${\mathcal P}_2$ under conjugation
by the Yang--Baxter gates $B_{\epsilon,\eta}$ (\ref{B Bell transform}) where the variables $\epsilon$ and $\eta$ are
relabeled as subscripts. }
\end{center}
\end{table}
Besides such the decompositions, the Clifford gates $B_{\epsilon,\eta}$ are verified to preserve
the properties of Pauli group $\mathcal P_2$ generated by $X_1, X_2$ and $Z_1, Z_2$ under conjugation
by the Yang--Baxter gates $B_{\epsilon,\eta}$ (\ref{B Bell transform}), see Table~\ref{clifford_YBG}.
Note that the results in Table~\ref{clifford_YBG} will be exploited in the study of
the Yang--Baxter gate approach to teleportation-based quantum computation.
\subsection{Quantum teleportation circuit using the Yang--Baxter gate}
\begin{table}
\begin{center}
\footnotesize
\begin{tabular}{c|c|c|c|c|c}
\hline\hline
$\epsilon$ & $\eta$ & $p_{\epsilon,\eta}$ & $p'_{\epsilon,\eta}$ & $a_{\epsilon,\eta}$ & $b_{\epsilon,\eta}$\\ \hline
$-1$ & $1$ & $j\cdot(j+k+l)$ & $i\cdot j+(k+l)\cdot (j+l+1)$ & & \\ \cline{1-4}
$1$ & $-1$ & $k\cdot l+(k+l)\cdot(j+l+1)$ & $k\cdot l+(j+1)\cdot(i+k+l)$ & \raisebox{1.5ex}[0pt]{$j+l+1$} & \raisebox{1.6ex}[0pt]{$i+j+k+l$} \\ \hline
$1$ & $1$ & $i\cdot j+(k+l)\cdot(i+k+1)$ & $i\cdot (i+k+l)$ & & \\ \cline{1-4}
$-1$ & $-1$ & $k\cdot l+(i+1)\cdot(j+k+l)$ & $l+i\cdot(k+l)$ & \raisebox{1.5ex}[0pt]{$i+k+1$} & \raisebox{1.6ex}[0pt]{$i+j+k+l$} \\ \hline\hline
\end{tabular}
\caption{\label{index p a b} The indices $p_{\epsilon,\eta}$, $p'_{\epsilon,\eta}$, $a_{\epsilon,\eta}$ and $b_{\epsilon,\eta}$ for both the local unitary gate
$W_{\epsilon,\eta}$ (\ref{W epsilon}) in the teleportation equation (\ref{tele eqa B1}) and the local unitary gate $W'_{\epsilon,\eta}$ (\ref{W epsilon prime})
in the teleportation equation (\ref{tele eqa B2}). The multiplication $\cdot$ is a logical AND operation and the addition $+$ is a binary addition. }
\end{center}
\end{table}
The essential ingredient in quantum teleportation circuit model is the teleportation operator introduced in \cite{ZZ14} which is the tensor product of
the identity operator, the Bell transform and its inverse. Here, the special type II Yang--Baxter gates $B_{\epsilon,\eta}$ are the Bell transform,
so we make use of a similar type of the teleportation operators
\begin{equation}
\label{tele operator}
(B_{\epsilon,\eta}\otimes 1\!\!1_2)(1\!\!1_2\otimes B_{\epsilon,\eta})\quad \text{or} \quad (1\!\!1_2\otimes B_{\epsilon,\eta})(B_{\epsilon,\eta}\otimes 1\!\!1_2)
\end{equation}
to act on the product state $|\alpha\rangle|kl\rangle$ or $|kl\rangle|\alpha\rangle$ respectively. They give rise to the type of the teleportation equation
\begin{equation}
\label{tele eqa B1}
(B_{\epsilon,\eta}\otimes 1\!\!1_2)(1\!\!1_2\otimes B_{\epsilon,\eta})(|\alpha\rangle\otimes |kl\rangle)=\frac{1}{2}\sum_{i,j=0}^1|ij\rangle W_{\epsilon,\eta}|\alpha\rangle,
\end{equation}
which is associated with the teleportation equation (\ref{tele_eq}) with the local single-qubit gate
\begin{equation}
\label{W epsilon}
W_{\epsilon,\eta}=(-1)^{p_{\epsilon,\eta}}Z^{a_{\epsilon,\eta}}X^{b_{\epsilon,\eta}},
\end{equation}
to represent the protocol of transmitting an unknown qubit $|\alpha\rangle$ from Alice to Bob,
and the other type of the teleportation equation
\begin{equation}
\label{tele eqa B2}
(1\!\!1_2 \otimes B_{\epsilon,\eta})(B_{\epsilon,\eta}\otimes 1\!\!1_2)(|kl\rangle\otimes|\alpha\rangle )=\frac{1}{2}\sum_{i,j=0}^1
W'_{\epsilon,\eta}|\alpha\rangle|ij\rangle,
\end{equation}
which is related to the teleportation equation (\ref{tele_eq_transpose}) with the local single-qubit gate
\begin{equation}
\label{W epsilon prime}
W'_{\epsilon,\eta}=(-1)^{p'_{\epsilon,\eta}}Z^{a_{\epsilon,\eta}}X^{b_{\epsilon,\eta}},
\end{equation}
to represent the protocol of transmitting an unknown qubit $|\alpha\rangle$ from Bob to Alice, where the explicit expressions
for the indices $p_{\epsilon, \eta}$, $p'_{\epsilon, \eta}$, $a_{\epsilon,\eta}$ and $b_{\epsilon,\eta}$ in both single-qubit gates $W_{\epsilon,\eta}$ (\ref{W epsilon}) and $W'_{\epsilon,\eta}$
(\ref{W epsilon prime}) are shown up in Table~\ref{index p a b}. Note that the local single-qubit gates $W_{\epsilon,\eta}$ (\ref{W epsilon}) and $W'_{\epsilon,\eta}$
(\ref{W epsilon prime}) are almost the same except the global phase factors.
Based on the above teleportation equations (\ref{tele eqa B1}) or (\ref{tele eqa B2}) using the special type II Yang--Baxter gates $B_{\epsilon,\eta}$, we continue to
perform single-qubit measurements, classical communication and local unitary corrections to complete the entire quantum teleportation protocol. For example, the
quantum teleportation circuit model associated with the teleportation equation (\ref{tele eqa B1}) is drawn in Figure~\ref{fig_tele_YBG}, where the braiding
configuration (\ref{TL B Bell transform}) of the Yang--Baxter gate $B_{\epsilon,\eta}$ has been applied. About the teleportation circuit model
in Figure~\ref{fig_tele_YBG}, the special type II Yang--Baxter gate $B(\epsilon,\eta)$ acting on the product state $|kl\rangle$ gives the prepared Bell states,
and the same Yang--Baxter gate $B(\epsilon,\eta)$ followed by two single-qubit measurements works as the Bell measurements.
\begin{figure}
\begin{center}
\includegraphics[width=8cm]{tele_YBG}
\end{center}
\caption{\label{fig_tele_YBG} Quantum circuit of teleportation corresponding to the teleportation equation (\ref{tele eqa B1}).
The Yang--Baxter gate $B(\epsilon,\eta)$ (\ref{B Bell transform}) is represented by a braiding configuration inside a two-qubit gate box.
In accordance with the conventional rules of
the quantum circuit diagram in Figure~\ref{fig_tele_circuit}, the braiding configuration (\ref{TL B Bell transform}) of the Yang--Baxter gate $B(\epsilon,\eta)$
has been adjusted in the left-to-right direction from the original down-to-up direction. The left lower braiding on the product state $|kl\rangle$ is used to generate
the prepared Bell states, and the right higher braiding followed with single-qubit state measurements works as the Bell measurements. The present quantum circuit model is essentially equivalent with the one in Figure~\ref{fig_tele_circuit}, while the superficial differences between them are due to the fact that we are using various
of presentations of both Bell states and Bell measurements. }
\end{figure}
Note that the inverse of the Yang--Baxter gate $B(\epsilon,\eta)$, denoted by $B^\dag(\epsilon,\eta)$, is related to itself with the algebraic relation (\ref{B and B inverse relation}). In other words, the inverse of the Yang--Baxter gate $B(\epsilon,\eta)$ is still the Bell transform \cite{ZZ14}. Hence in terms of the inverse of the Yang--Baxter
gate $B(\epsilon,\eta)$, the teleportation operators (\ref{tele operator}) can be reformulated as the same type of the teleportation operators used in \cite{ZZ14} where the
inverse of the Bell transform \cite{ZZ14} with single-qubit measurements represents the Bell measurements.
\begin{figure}
\begin{center}
\includegraphics[width=8cm]{single_tele_YGB}
\end{center}
\caption{\label{fig_single_tele_YGB} Quantum circuit for the fault-tolerant construction of a single-qubit gate $U$ in the Yang--Baxter gate approach to teleportation-based quantum computation, associated with the teleportation equation (\ref{tele one qubit gate B}), and it is equivalent with the quantum
circuit in Figure~\ref{fig_tele_single_gate}. }
\end{figure}
\subsection{Teleportation-based quantum computation using the Yang--Baxter gate}
Under the guidance of the fault-tolerant construction of single-qubit gates in Subsubsection~\ref{subsection: single-qubit}, we prepare the three-qubit quantum state in the
the Yang--Baxter gate approach, expressed as
\begin{equation}
(1\!\!1_2\otimes 1\!\!1_2\otimes U)(1\!\!1_2\otimes B_{\epsilon,\eta})(|\alpha\rangle\otimes|kl\rangle)
\end{equation}
with the single-qubit gate $U$ acting on the Bell states, which are generated by the special type II Yang--Baxter gate $B(\epsilon,\eta)$ on the product basis $|kl\rangle$,
and then apply the Yang--Baxter gate $B(\epsilon,\eta)$ on the first and second qubits to derive the teleportation equation
\begin{equation}
\label{tele one qubit gate B}
(B_{\epsilon,\eta}\otimes 1\!\!1_2)(1\!\!1_2\otimes 1\!\!1_2\otimes U)(1\!\!1_2 \otimes B_{\epsilon,\eta})(|\alpha\rangle\otimes |kl\rangle)=\frac{1}{2}\sum_{i,j=0}^1|ij\rangle R_{\epsilon,\eta}U|\alpha\rangle,
\end{equation}
which is associated with the teleportation equation (\ref{tele_single}), with $R_{\epsilon,\eta}=UW_{\epsilon,\eta}U^\dag$, the single-qubit gate $W_{\epsilon,\eta}$ defined in
(\ref{W epsilon}).
For example, when the single-qubit gate $U$ is the Hadamard gate $H$ (\ref{Hadamard gate}) and $\pi/8$ gate $T$ (\ref{pi/8}), the corresponding single-qubit gates $R_{\epsilon,\eta}$ have the form respectively,
\begin{equation}
\label{local correction R for H and T}
R_{\epsilon,\eta}(H)=(-1)^{p_{\epsilon,\eta}}X^{a_{\epsilon,\eta}}Z^{b_{\epsilon,\eta}},\quad R_{\epsilon,\eta}(T)=(-1)^{p_{\epsilon,\eta}}Z^{a_{\epsilon,\eta}}\left(\frac{X-\sqrt{-1}Y}{\sqrt{2}}\right)^{b_{\epsilon,\eta}},
\end{equation}
where the indices $p_{\epsilon,\eta}$, $a_{\epsilon,\eta}$ and $b_{\epsilon,\eta}$ are those in Table~\ref{index p a b}. We can compare these results with their counterparts
obtained in Subsubsection~\ref{subsection: single-qubit}, such as (\ref{RTij}). Note that the $R_{\epsilon,\eta}(H)$ gates are the Pauli gates with the global phase factors,
whereas the $R_{\epsilon,\eta}(T)$ gates are the Clifford gates.
Combining the teleportation equation (\ref{tele one qubit gate B}) with both two single-qubit measurements and the local unitary correction operator $R^{\dag}_{\epsilon,\eta}$,
the quantum circuit model for the fault-tolerant construction of the single-qubit gate $U$ in the Yang--Baxter gate approach is presented in Figure~\ref{fig_single_tele_YGB},
which can be compared with the quantum circuit model in Figure~\ref{fig_tele_single_gate}.
In accordance with the fault-tolerant construction of a two-qubit Clifford gate in Subsubsection \ref{subsection: two-qubit}, we perform the fault-tolerant
construction of the special type II Yang--Baxter gate $B(\epsilon,\eta)$ (\ref{B Bell transform}) in the following steps. Firstly, we prepare the four-qubit
entangled state in the Yang--Baxter gate approach as
\begin{equation}
(1\!\!1_2\otimes B_{\epsilon,\eta}\otimes 1\!\!1_2)(B_{\epsilon,\eta}\otimes B_{\epsilon,\eta})(|k_1l_1\rangle\otimes |k_2l_2\rangle).
\end{equation}
Secondly, we perform two Bell measurements respectively via two Yang--Baxter gates $B(\epsilon,\eta)$ followed by product-basis measurements,
so that we have the teleportation equation
\begin{eqnarray}
\label{tele two qubit gate B}
&&(B_{\epsilon,\eta}\otimes B_{\epsilon,\eta}\otimes B_{\epsilon,\eta})(1\!\!1_2\otimes B_{\epsilon,\eta}\otimes B_{\epsilon,\eta} \otimes1\!\!1_2)(|\alpha\rangle\otimes |k_1l_1\rangle\otimes |k_2l_2\rangle\otimes |\beta\rangle) \nonumber\\
&=& \frac{1}{4}\sum_{i_1,j_1=0}^1\sum_{i_2,j_2=0}^1(1\!\!1_4\otimes Q_{\epsilon,\eta}\otimes P_{\epsilon,\eta}\otimes 1\!\!1_4)(|i_1j_1\rangle\otimes B_{\epsilon,\eta}|\alpha\beta\rangle\otimes|i_2j_2\rangle),
\end{eqnarray}
which is related to the teleportation equation (\ref{tele_two}) and has the quantum circuit in Figure~\ref{fig_two_tele_YGB}. The single-qubit gates $Q_{\epsilon,\eta}$ and $P_{\epsilon,\eta}$ are calculated by the formula
\begin{equation}
Q_{\epsilon,\eta}\otimes P_{\epsilon,\eta}=B_{\epsilon,\eta}(W_{\epsilon,\eta}\otimes W'_{\epsilon,\eta})B_{-\epsilon,-\eta},
\end{equation}
in which $W_{\epsilon,\eta}$ is defined in (\ref{tele eqa B1}) and depends on the indices $i_1$, $j_1$, and $W'_{\epsilon,\eta}$ is
defined in (\ref{tele eqa B2}) and depends on the indices $i_2$, $j_2$, and the explicit expressions of $Q_{\epsilon,\eta}$ and
$P_{\epsilon,\eta}$ are listed in Table~\ref{Q and P}.
\begin{figure}
\begin{center}
\includegraphics[width=8cm]{two_tele_YGB}
\end{center}
\caption{\label{fig_two_tele_YGB}
Quantum circuit for the fault-tolerant construction of the two-qubit gate $B_{\epsilon,\eta}$ (\ref{B Bell transform}) in the Yang--Baxter gate approach to
teleportation-based quantum computation, associated with the teleportation equation (\ref{tele two qubit gate B}), and it is equivalent with
the quantum circuit in Figure~\ref{fig_tele_two_gate}. Explicit expressions for single-qubit gates $Q_{\epsilon,\eta}^\dag$ and $P_{\epsilon,\eta}^\dag$
are shown up in Table~\ref{Q and P}.}
\end{figure}
\begin{table}
\begin{center}
\footnotesize
\begin{tabular}{c|c|c|c}
\hline\hline
$\epsilon$ & $\eta$ & $Q_{\epsilon,\eta}$ & $P_{\epsilon,\eta}$ \\ \hline
$-1$ & $1$ & $(-1)^{p_1+a_1}Y^{a_1}X^{b_1+b_2+a_2}$ & $(-1)^{p'_2+a_2}Y^{a_1}X^{a_2}Z^{b_2}$ \\ \hline
$1$ & $-1$ & $(-1)^{p_1}Y^{a_1}X^{b_1+b_2+a_2}$ & $(-1)^{p'_2+b_2}Y^{a_1}X^{a_2}Z^{b_2}$ \\ \hline
$1$ & $1$ & $(-1)^{p_1+a_1}X^{a_1}Z^{b_1}Y^{a_2}$ & $(-1)^{p'_2+a_2}X^{a_1+b_1}Y^{a_2}X^{b_2}$ \\ \hline
$-1$ & $-1$ & $(-1)^{p_1+b_1}X^{a_1}Z^{b_1}Y^{a_2}$ & $(-1)^{p'_2}X^{a_1+b_1}Y^{a_2}X^{b_2}$ \\ \hline\hline
\end{tabular}
\caption{\label{Q and P} The $Q_{\epsilon,\eta}$ and $P_{\epsilon,\eta}$ gates in the teleportation equation (\ref{tele two qubit gate B}).
For simplicity, the subscripts $\epsilon$ and $\eta$ of $p$, $p'$, $a$ and $b$, are omitted in the present table without confusion.
The parameters $p_1$, $p_1'$, $a_1$ and $b_1$ depend on the parameters $i_1$, $j_1$, $k_1$ and $l_1$, and are calculated by the formulas
in Table~\ref{index p a b}, while the parameters $p_2$, $p_2'$, $a_2$ and $b_2$ depend on the parameters $i_2$, $j_2$, $k_2$ and $l_2$,
and are calculated by the formulas in Table~\ref{index p a b}. }
\end{center}
\end{table}
\subsection{Special example: $\epsilon=1,\eta=1$ and $|kl\rangle=|11\rangle$}
For readers' convenience, we make a brief summary on the results in the Yang--Baxter gate approach to the teleportation-based quantum computation
using the Yang--Baxter gate $B$ (\ref{YBG B}), which is the $B_{\epsilon,\eta}$ gate (\ref{B Bell transform}) of $\epsilon=1,\eta=1$.
To further simplify the calculation results without losing any significant physical meaning, we choose the product state $|kl\rangle=|11\rangle$ in
the state preparation. Note that the Yang--Baxter gate $B$ (\ref{YBG B}) acting on the state $|11\rangle$ gives rise to the EPR state (\ref{EPR}).
First, the teleportation equation associated with the teleportation operator $(B\otimes 1\!\!1_2)(1\!\!1_2 \otimes B)$ on the product state
$|\alpha\rangle\otimes |11\rangle$ has the form
\begin{equation}
\label{example tele eqa B1}
(B\otimes 1\!\!1_2)(1\!\!1_2 \otimes B)(|\alpha\rangle\otimes |11\rangle) = \frac 1 2 \sum_{i,j=0}^1 |ij\rangle W_B|\alpha\rangle,
\end{equation}
with the single-qubit gate $W_B$ given by
\begin{equation}
\label{W_B}
W_B=(-1)^{i\cdot j}Z^iX^{i+j},
\end{equation}
which is a special case of the teleportation equation (\ref{tele eqa B1}). Similarly, the teleportation equation, as a special example
of (\ref{tele eqa B2}), is obtained to be
\begin{equation}
\label{example tele eqa B1_transpose}
(1\!\!1_2 \otimes B)(B\otimes 1\!\!1_2)(|11\rangle\otimes|\alpha\rangle ) = \frac{1}{2}\sum_{i,j=0}^1 W^T_B|\alpha\rangle|ij\rangle,
\end{equation}
where $W^T_B$ is the matrix transpose of $W_B$ (\ref{W_B}).
Second, the teleportation equation associated with the fault-tolerant construction of the single-qubit gate $U$ has the form
\begin{equation}
\label{example single qubit gate via B}
(B\otimes 1\!\!1_2)(1\!\!1_2\otimes 1\!\!1_2\otimes U)(1\!\!1_2 \otimes B)(|\alpha\rangle\otimes |11\rangle)
=\frac{1}{2}\sum_{i,j=0}^1|ij\rangle R_B\, U|\alpha\rangle,
\end{equation}
with $R_B=UW_BU^\dag$, the $W_B$ gate defined in (\ref{W_B}), which is a special example of the teleportation equation (\ref{tele one qubit gate B}).
As the single-qubit gate $U$ is the Hadamard gate $H$ (\ref{Hadamard gate}) and the $\pi/8$ gate $T$ (\ref{pi/8}) respectively, the local
unitary gate $R_B$ has the explicit forms
\begin{equation}
R_B(H)=(-1)^{i\cdot j}X^iZ^{i+j},\quad R_B(T)=(-1)^{i\cdot j}Z^i\left(\frac{X-\sqrt{-1}Y}{\sqrt{2}}\right)^{i+j}.
\end{equation}
Third, the teleportation equation associated with the fault-tolerant construction of the Yang--Baxter gate $B$, as a special case of (\ref{tele two qubit gate B}),
has the form
\begin{eqnarray}
\label{example two qubit gate via B}
&&(B\otimes B\otimes B)(1\!\!1_2\otimes B\otimes B \otimes1\!\!1_2)(|\alpha\rangle\otimes |11\rangle\otimes |11\rangle\otimes |\beta\rangle) \nonumber\\
&=& \frac{1}{4}\sum_{i_1,j_1=0}^1\sum_{i_2,j_2=0}^1(1\!\!1_4\otimes Q_B\otimes P_B\otimes 1\!\!1_4)(|i_1j_1\rangle\otimes B|\alpha\beta\rangle\otimes|i_2j_2\rangle),
\end{eqnarray}
with the single-qubit gates $Q_B$ and $P_B$ calculated by
\begin{equation}
\label{Q P for B}
Q_B\otimes P_B=B(W_B\otimes W_B^T)B^\dag,
\end{equation}
where $W_B$ depends on the parameters $i_1$, $j_1$ and $W_B^T$ depends on the parameters $i_2$, $j_2$. The explicit formalisms of $Q_B$ and $P_B$ are
expressed as
\begin{equation}
Q_B=(-1)^{i_1\cdot(j_1+1)}X^{i_1}Z^{i_1+j_1}Y^{i_2},\quad P_B=X^{j_1}Y^{i_2}X^{i_2+j_2}
\end{equation}
which can be derived with Table~\ref{Q and P} by picking up the third row of $\epsilon=1,\eta=1$, and setting $k_1=l_1=k_2=l_2=1$.
Note that the key reason that the results in this subsection are presented is that they are to be explained in the extended Temperley--Lieb
diagrammatical approach in Section~\ref{relationship_YBE_TL}.
\section{Relationship between the extended Temperley--Lieb diagrammatical approach and the Yang--Baxter gate approach}
\label{relationship_YBE_TL}
So far, we study the two approaches to teleportation-based quantum computation: the one is the extended Temperley--Lieb diagrammatical approach
in Section~\ref{TL diagrammatical for tele based quantum computation}, and the other is the Yang--Baxter gate approach
in Section~\ref{YBG approach to tele based QC}. In this section, we are in a position to consider the relationship between these two approaches.
With the extended Temperley--Lieb configuration (\ref{TL B Bell transform}) of the special type II Yang--Baxter
gate $B(\epsilon,\eta)$ (\ref{B Bell transform}) in Section~\ref{section the YBG and TL}, we are able to recast all the algebraic results
of the Yang--Baxter gate approach in Section~\ref{YBG approach to tele based QC} into the topological diagrammatical configurations, which are
found to be those in the extended Temperley--Lieb diagrammatical approach in Section~\ref{TL diagrammatical for tele based quantum computation}.
In the following discussion, we concentrate on the Yang--Baxter gate $B$ (\ref{YBG B}) as an example for the special type II Yang--Baxter
gate $B(\epsilon,\eta)$ with $\epsilon=\eta=1$.
\subsection{The product basis and the Bell basis}
The Yang--Baxter gate approach to teleportation-based quantum computation in Section~\ref{YBG approach to tele based QC} is based on
the observation that the Bell states (\ref{Bell states}) can be replaced by the Yang--Baxter gate acting on the product states and
the Bell measurements can be substituted by the Yang--Baxter gate followed with product-basis measurements. Hence, the first thing
that we are going to do is to describe the product basis and the Bell basis in the extended Temperley--Lieb diagrammatical approach.
With the definition of the Bell basis (\ref{Bell states}), it is easy to formulate the two-qubit product basis in terms of the Bell basis,
so the two-qubit product states have the extended Temperley--Lieb diagrammatical configurations respectively expressed as
\begin{equation}
\label{product state in TL configuration 1}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(70,12)
\put(-10,2){\line(0,1){10}}
\put(0,2){\line(0,1){10}}
\put(-2.1,0){\tiny{$\nabla$}}
\put(-12.1,0){\tiny{$\nabla$}}
\put(8,4){$=\frac 1 {\sqrt2}($}
\put(42,0){\line(0,1){12}}
\put(32,0){\line(0,1){12}}
\put(32,0){\line(1,0){10}}
\put(48,4){$+$}
\put(69,0){\line(0,1){12}}
\put(59,0){\line(0,1){12}}
\put(59,0){\line(1,0){10}}
\put(69,6){\circle*{2.}}
\put(71,5){\tiny{$Z$}}
\put(75,4){$);$}
\end{picture}
\end{array}
\end{equation}
\begin{equation}
\label{product state in TL configuration 2}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(70,12)
\put(-10,2){\line(0,1){10}}
\put(0,2){\line(0,1){10}}
\put(-2.1,0){\tiny{$\nabla$}}
\put(-12.1,0){\tiny{$\nabla$}}
\put(0,7){\circle*{2.}}
\put(2,6){\tiny{$X$}}
\put(8,4){$=\frac 1 {\sqrt2}($}
\put(42,0){\line(0,1){12}}
\put(32,0){\line(0,1){12}}
\put(32,0){\line(1,0){10}}
\put(42,6){\circle*{2.}}
\put(44,5){\tiny{$X$}}
\put(48,4){$+$}
\put(69,0){\line(0,1){12}}
\put(59,0){\line(0,1){12}}
\put(59,0){\line(1,0){10}}
\put(69,6){\circle*{2.}}
\put(71,5){\tiny{$XZ$}}
\put(79,4){$);$}
\end{picture}
\end{array}
\end{equation}
\begin{equation}
\label{product state in TL configuration 3}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(70,12)
\put(-10,2){\line(0,1){10}}
\put(0,2){\line(0,1){10}}
\put(-2.1,0){\tiny{$\nabla$}}
\put(-12.1,0){\tiny{$\nabla$}}
\put(-10,7){\circle*{2.}}
\put(-8,6){\tiny{$X$}}
\put(8,4){$=\frac 1 {\sqrt2}($}
\put(42,0){\line(0,1){12}}
\put(32,0){\line(0,1){12}}
\put(32,0){\line(1,0){10}}
\put(42,6){\circle*{2.}}
\put(44,5){\tiny{$X$}}
\put(48,4){$-$}
\put(69,0){\line(0,1){12}}
\put(59,0){\line(0,1){12}}
\put(59,0){\line(1,0){10}}
\put(69,6){\circle*{2.}}
\put(71,5){\tiny{$XZ$}}
\put(79,4){$);$}
\end{picture}
\end{array}
\end{equation}
\begin{equation}
\label{product state in TL configuration 4}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(70,12)
\put(-10,2){\line(0,1){10}}
\put(0,2){\line(0,1){10}}
\put(-2.1,0){\tiny{$\nabla$}}
\put(-12.1,0){\tiny{$\nabla$}}
\put(-10,7){\circle*{2.}}
\put(-8,6){\tiny{$X$}}
\put(0,7){\circle*{2.}}
\put(2,6){\tiny{$X$}}
\put(8,4){$=\frac 1 {\sqrt2}($}
\put(42,0){\line(0,1){12}}
\put(32,0){\line(0,1){12}}
\put(32,0){\line(1,0){10}}
\put(48,4){$-$}
\put(69,0){\line(0,1){12}}
\put(59,0){\line(0,1){12}}
\put(59,0){\line(1,0){10}}
\put(69,6){\circle*{2.}}
\put(71,5){\tiny{$Z$}}
\put(75,4){$);$}
\end{picture}
\end{array}
\end{equation}
where the vertical line with $\nabla$ stands for the state $|0\rangle$ and naturally the one with the Pauli $X$ gate stands for the state $|1\rangle$.
And the adjoint of the above algebraic relations (\ref{product state in TL configuration 1})-(\ref{product state in TL configuration 4}) have the
extended Temperley--Lieb diagrammatical configurations respectively as
\begin{equation}
\label{conjugation product state in TL configuration 1}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(70,12)
\put(-10,0){\line(0,1){10}}
\put(0,0){\line(0,1){10}}
\put(-2.1,10){\tiny{$\triangle$}}
\put(-12.1,10){\tiny{$\triangle$}}
\put(8,4){$=\frac 1 {\sqrt2}($}
\put(42,0){\line(0,1){12}}
\put(32,0){\line(0,1){12}}
\put(32,12){\line(1,0){10}}
\put(48,4){$+$}
\put(69,0){\line(0,1){12}}
\put(59,0){\line(0,1){12}}
\put(59,12){\line(1,0){10}}
\put(69,6){\circle*{2.}}
\put(71,5){\tiny{$Z$}}
\put(75,4){$);$}
\end{picture}
\end{array}
\end{equation}
\begin{equation}
\label{conjugation product state in TL configuration 2}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(70,12)
\put(-10,0){\line(0,1){10}}
\put(0,0){\line(0,1){10}}
\put(-2.1,10){\tiny{$\triangle$}}
\put(-12.1,10){\tiny{$\triangle$}}
\put(0,5){\circle*{2.}}
\put(2,4){\tiny{$X$}}
\put(8,4){$=\frac 1 {\sqrt2}($}
\put(42,0){\line(0,1){12}}
\put(32,0){\line(0,1){12}}
\put(32,12){\line(1,0){10}}
\put(42,6){\circle*{2.}}
\put(44,5){\tiny{$X$}}
\put(48,4){$+$}
\put(69,0){\line(0,1){12}}
\put(59,0){\line(0,1){12}}
\put(59,12){\line(1,0){10}}
\put(69,6){\circle*{2.}}
\put(71,5){\tiny{$ZX$}}
\put(79,4){$);$}
\end{picture}
\end{array}
\end{equation}
\begin{equation}
\label{conjugation product state in TL configuration 3}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(70,12)
\put(-10,0){\line(0,1){10}}
\put(0,0){\line(0,1){10}}
\put(-2.1,10){\tiny{$\triangle$}}
\put(-12.1,10){\tiny{$\triangle$}}
\put(-10,5){\circle*{2.}}
\put(-8,4){\tiny{$X$}}
\put(8,4){$=\frac 1 {\sqrt2}($}
\put(42,0){\line(0,1){12}}
\put(32,0){\line(0,1){12}}
\put(32,12){\line(1,0){10}}
\put(42,6){\circle*{2.}}
\put(44,5){\tiny{$X$}}
\put(48,4){$-$}
\put(69,0){\line(0,1){12}}
\put(59,0){\line(0,1){12}}
\put(59,12){\line(1,0){10}}
\put(69,6){\circle*{2.}}
\put(71,5){\tiny{$ZX$}}
\put(79,4){$);$}
\end{picture}
\end{array}
\end{equation}
\begin{equation}
\label{conjugation product state in TL configuration 4}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(70,12)
\put(-10,0){\line(0,1){10}}
\put(0,0){\line(0,1){10}}
\put(-2.1,10){\tiny{$\triangle$}}
\put(-12.1,10){\tiny{$\triangle$}}
\put(-10,5){\circle*{2.}}
\put(-8,4){\tiny{$X$}}
\put(0,5){\circle*{2.}}
\put(2,4){\tiny{$X$}}
\put(8,4){$=\frac 1 {\sqrt2}($}
\put(42,0){\line(0,1){12}}
\put(32,0){\line(0,1){12}}
\put(32,12){\line(1,0){10}}
\put(48,4){$-$}
\put(69,0){\line(0,1){12}}
\put(59,0){\line(0,1){12}}
\put(59,12){\line(1,0){10}}
\put(69,6){\circle*{2.}}
\put(71,5){\tiny{$Z$}}
\put(75,4){$).$}
\end{picture}
\end{array}
\end{equation}
Now we study the action of the Yang--Baxter gate $B$ (\ref{YBG B}) on the product state $|ij\rangle$ in the extended Temperley--Lieb diagrammatical
approach. The extended Temperley--Lieb configuration of the Yang--Baxter gate $B$ (\ref{YBG B}), has the form
\begin{equation}
\label{example TL B Bell transform}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(130,40)
\put(-25,-2){\footnotesize{$B$}}
\put(-31,29){\line(0,1){6}}
\put(-31,11){\line(0,-1){6}}
\put(-13,29){\line(0,1){6}}
\put(-13,11){\line(0,-1){6}}
\put(-34,8){\line(0,1){24}}
\put(-10,8){\line(0,1){24}}
\put(-34,8){\line(1,0){24}}
\put(-34,32){\line(1,0){24}}
\put(-31,29){\line(1,-1){18}}
\put(-31,11){\line(1,1){7.3}}
\put(-13,29){\line(-1,-1){7.3}}
\put(-3,19){$=\frac 1 {\sqrt 2}($}
\put(32,5){\line(0,1){30}}
\put(22,5){\line(0,1){30}}
\put(40,19){$+$}
\put(62,23){\line(0,1){12}}
\put(52,23){\line(0,1){12}}
\put(52,23){\line(1,0){10}}
\put(62,29){\circle*{2.}}
\put(64,28){\tiny{$Z$}}
\put(52,5){\line(0,1){12}}
\put(62,5){\line(0,1){12}}
\put(52,17){\line(1,0){10}}
\put(70,19){$-$}
\put(92,23){\line(0,1){12}}
\put(82,23){\line(0,1){12}}
\put(82,23){\line(1,0){10}}
\put(82,5){\line(0,1){12}}
\put(92,5){\line(0,1){12}}
\put(82,17){\line(1,0){10}}
\put(92,11){\circle*{2.}}
\put(94,10){\tiny{$Z$}}
\put(100,19){$-$}
\put(122,23){\line(0,1){12}}
\put(112,23){\line(0,1){12}}
\put(112,23){\line(1,0){10}}
\put(122,29){\circle*{2.}}
\put(124,28){\tiny{$X$}}
\put(112,5){\line(0,1){12}}
\put(122,5){\line(0,1){12}}
\put(112,17){\line(1,0){10}}
\put(122,11){\circle*{2.}}
\put(124,10){\tiny{$ZX$}}
\put(130,19){$+$}
\put(152,23){\line(0,1){12}}
\put(142,23){\line(0,1){12}}
\put(142,23){\line(1,0){10}}
\put(152,29){\circle*{2.}}
\put(154,28){\tiny{$XZ$}}
\put(142,5){\line(0,1){12}}
\put(152,5){\line(0,1){12}}
\put(142,17){\line(1,0){10}}
\put(152,11){\circle*{2.}}
\put(154,10){\tiny{$X$}}
\put(164,19){$)$}
\end{picture}
\end{array}
\end{equation}
which is directly obtained from (\ref{TL B Bell transform}) by setting $\epsilon=1,\eta=1$. With the extended Temperley--Lieb diagrammatical
rules \cite{Zhang06} that assign a normalization factor 1 to a loop configuration and assign a normalized trace of single-qubit gates to a loop
with the action of associated single-qubit gates, for example, we apply the Yang--Baxter gate $B$ on the product state $|11\rangle$, namely
calculate $B|11\rangle$ in the diagrammatical approach,
\begin{equation}
\label{B 11 state}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(160,15)
\put(-6,4){$B|11\rangle=\frac 1 2($}
\put(42,0){\line(0,1){12}}
\put(32,0){\line(0,1){12}}
\put(32,0){\line(1,0){10}}
\put(48,4){$-$}
\put(69,0){\line(0,1){12}}
\put(59,0){\line(0,1){12}}
\put(59,0){\line(1,0){10}}
\put(69,6){\circle*{2.}}
\put(71,5){\tiny{$Z$}}
\put(76,4){$+$}
\put(96,0){\line(0,1){12}}
\put(86,0){\line(0,1){12}}
\put(86,0){\line(1,0){10}}
\put(96,6){\circle*{2.}}
\put(98,5){\tiny{$Z$}}
\put(103,4){$+$}
\put(123,0){\line(0,1){12}}
\put(113,0){\line(0,1){12}}
\put(113,0){\line(1,0){10}}
\put(128,4){$)=$}
\put(151,0){\line(0,1){12}}
\put(141,0){\line(0,1){12}}
\put(141,0){\line(1,0){10}}
\end{picture}
\end{array}
\end{equation}
which is the cup representation of the EPR state $|\Psi\rangle$, see (\ref{TL diag Bell state}). In the same manner, performing the Yang--Baxter gate $B$
on the other product states, we attain the extended Temperley--Lieb configurations of the other Bell states, summarized in
\begin{equation}
\label{B ij state}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(170,15)
\put(0,4){$B(|00\rangle,|01\rangle,|10\rangle,|11\rangle)=($}
\put(95,0){\line(0,1){12}}
\put(85,0){\line(0,1){12}}
\put(85,0){\line(1,0){10}}
\put(95,6){\circle*{2.}}
\put(97,5){\tiny{$Z$}}
\put(100,1){$,$}
\put(116,0){\line(0,1){12}}
\put(106,0){\line(0,1){12}}
\put(106,0){\line(1,0){10}}
\put(116,6){\circle*{2.}}
\put(117,5){\tiny{$XZ$}}
\put(122,1){$,$}
\put(138,0){\line(0,1){12}}
\put(128,0){\line(0,1){12}}
\put(128,0){\line(1,0){10}}
\put(138,6){\circle*{2.}}
\put(140,5){\tiny{$X$}}
\put(144,1){$,$}
\put(160,0){\line(0,1){12}}
\put(150,0){\line(0,1){12}}
\put(150,0){\line(1,0){10}}
\put(165,4){$)$}
\end{picture}
\end{array}
\end{equation}
which allows a more concise expression as
\begin{equation}
\label{B kl state index}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(50,15)
\put(0,4){$B|kl\rangle=$}
\put(38,0){\line(0,1){12}}
\put(28,0){\line(0,1){12}}
\put(28,0){\line(1,0){10}}
\put(38,6){\circle*{2.}}
\put(40,5){\tiny{$V_{kl}$}}
\end{picture}
\end{array}
\end{equation}
with the single-qubit gate $V_{kl}=X^{k+l}Z^{k+1}$. Note that the diagrammatical representation (\ref{B kl state index}) is associated
with the algebraic relation~(\ref{B Bell transform index}).
The Yang--Baxter gate $B$ followed with the product-state measurements has the algebraic form $|ij\rangle\langle ij|B$. For simplicity,
we neglect the post-measurement state $|ij\rangle$, and only calculate $\langle ij|B$ in the extended Temperley--Lieb diagrammatical
approach. For example, the quantum state $\langle00|B$ is calculated in the way
\begin{equation}
\label{11 B state}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(160,15)
\put(-6,4){$\langle00|B=\frac 1 2($}
\put(42,0){\line(0,1){12}}
\put(32,0){\line(0,1){12}}
\put(32,12){\line(1,0){10}}
\put(48,4){$+$}
\put(69,0){\line(0,1){12}}
\put(59,0){\line(0,1){12}}
\put(59,12){\line(1,0){10}}
\put(69,6){\circle*{2.}}
\put(71,5){\tiny{$Z$}}
\put(76,4){$-$}
\put(96,0){\line(0,1){12}}
\put(86,0){\line(0,1){12}}
\put(86,12){\line(1,0){10}}
\put(96,6){\circle*{2.}}
\put(98,5){\tiny{$Z$}}
\put(103,4){$+$}
\put(123,0){\line(0,1){12}}
\put(113,0){\line(0,1){12}}
\put(113,12){\line(1,0){10}}
\put(128,4){$)=$}
\put(151,0){\line(0,1){12}}
\put(141,0){\line(0,1){12}}
\put(141,12){\line(1,0){10}}
\end{picture}
\end{array}
\end{equation}
which is the cap representation of the EPR state $|\Psi\rangle$, see (\ref{TL diag Bell state_cap}). After calculating the
other cases, it turns out that the state $\langle ij|B$ has the representation,
\begin{equation}
\label{ij B state index}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(50,15)
\put(0,4){$\langle ij|B=$}
\put(38,0){\line(0,1){12}}
\put(28,0){\line(0,1){12}}
\put(28,12){\line(1,0){10}}
\put(38,6){\circle*{2.}}
\put(40,5){\tiny{$U_{ij}$}}
\end{picture}
\end{array}
\end{equation}
with the single-qubit gate $U_{ij}=(-1)^iZ^iX^{i+j}$. Note that the diagrammatical representation (\ref{ij B state index}) is associated with
the adjoint of the algebraic relation~(\ref{B inverse Bell transform index}) in which the Bell transform $B_{-1,-1}$ is the inverse of the Bell
transform $B$.
\subsection{Teleportation-based quantum computation}
Let us derive the extended Temperley--Lieb configurations in Section~\ref{TL diagrammatical for tele based quantum computation} respectively from
the Yang--Baxter gate approach to teleportation-based quantum computation in Section~\ref{YBG approach to tele based QC}.
The topological representation of quantum teleportation, such as (\ref{tele: tl}), can be regarded as the diagrammatical representation of
a matrix element of the teleportation equation~(\ref{example tele eqa B1}) using the Yang--Baxter gate $B$. With the cup state (\ref{B 11 state}) generated
by $B|11\rangle$ and the cap state (\ref{ij B state index}) generated by $\langle ij|B$, the teleportation operator $(B\otimes 1\!\!1_2)(1\!\!1\otimes B)$
has the matrix element as
\begin{equation}
\label{example teleportation operator as teleportation}
\setlength{\unitlength}{0.7mm}
\begin{array}{c}
\begin{picture}(140,25)
\put(-8,11){$(\langle ij|\otimes 1\!\! 1_2)(B\otimes 1\!\!1_2)(1\!\!1_2\otimes B)|\alpha\rangle\otimes|11\rangle=$}
\put(96,0){\line(0,1){24}}
\put(86,3.5){\line(0,1){20.5}}
\put(86,24){\line(1,0){10}}
\put(84,0){\makebox(4,4){$\nabla$}}
\put(96,0){\line(1,0){10}}
\put(106,0){\line(0,1){24}}
\put(96,18){\circle*{1.5}}
\put(97.5,17){\tiny{$U_{ij}$}}
\multiput(81,12)(1,0){30}{\line(1,0){.5}}
\put(115,11){$=\frac 1 2$}
\put(130,3.5){\line(0,1){20.5}}
\put(128,0){\makebox(4,4){$\nabla$}}
\put(130,13){\circle*{1.5}}
\put(132.5,12){\tiny{$U_{ij}^T$}}
\end{picture}
\end{array}
\end{equation}
in which the diagrammatical part below the dashed line denotes the prepared state $|\alpha\rangle\otimes|\Psi\rangle$ and the part above the dashed line
denotes the Bell measurement labeled by the single-qubit gate $U_{ij}$ in (\ref{ij B state index}). The transpose of $U_{ij}$ is
the local unitary gate $W_B$ (\ref{W_B}), namely, $U_{ij}^T=W_B$. Note that the post-measurement state $B^\dag|ij\rangle$ is neglected for simplicity. As
another example, the topological configuration (\ref{chain teleportation}) of the chained teleportation is obtained in the Yang--Baxter
gate approach,
\begin{eqnarray}
\label{chain teleportation B}
\setlength{\unitlength}{0.6mm}
\begin{array}{c}
\begin{picture}(170,25)
\put(-22,11){($_{1234}\langle0000|\otimes 1\!\! 1_2)B_{12}B_{34}B_{23}B_{45}|\alpha\rangle_1\otimes|1111\rangle_{2345}=$}
\put(110,0){\makebox(4,4){$\nabla$}}
\put(112,4){\line(0,1){20}}
\put(122,0){\line(0,1){24}}
\put(132,0){\line(0,1){24}}
\multiput(108,12)(1,0){48}{\line(1,0){.5}}
\put(122,0){\line(1,0){10}}
\put(112,24){\line(1,0){10}}
\put(132,24){\line(1,0){10}}
\put(142,0){\line(0,1){24}}
\put(142,0){\line(1,0){10}}
\put(152,0){\line(0,1){24}}
\put(158,7){\makebox(14,10){$=\frac 1 4$}}
\put(172,0){\makebox(4,4){$\nabla$}}
\put(174,4){\line(0,1){20}}
\end{picture}
\end{array}
\end{eqnarray}
where $B|11\rangle$ denotes the cup state (\ref{B 11 state}) and $\langle00|B$ denotes the cap state (\ref{11 B state}) and the symbol $B_{ij}$ means
that the Yang--Baxter gate $B$ is acting on both the $i$-th and $j$-th qubits.
The topological configuration (\ref{gate: one}) for the fault-tolerant construction of the single-qubit gate $U$ in teleportation-based
quantum computation is associated with the matrix element of the teleportation equation~(\ref{example single qubit gate via B}),
\begin{equation}
\label{example single qubit gate as teleportation}
\setlength{\unitlength}{0.7mm}
\begin{array}{c}
\begin{picture}(145,25)
\put(-20,11){$(\langle ij|\otimes 1\!\! 1_2)(B\otimes 1\!\!1_2)(1\!\!1_2\otimes1\!\!1_2\otimes U)(1\!\!1\otimes B)|\alpha\rangle\otimes|11\rangle=$}
\put(116,0){\line(0,1){24}}
\put(106,3.5){\line(0,1){20.5}}
\put(106,24){\line(1,0){10}}
\put(104,0){\makebox(4,4){$\nabla$}}
\put(116,0){\line(1,0){10}}
\put(126,0){\line(0,1){24}}
\put(116,18){\circle*{1.5}}
\put(117.5,17){\tiny{$W_B^T$}}
\put(126,6){\circle*{1.5}}
\put(127.5,5){\tiny{$U$}}
\multiput(101,12)(1,0){30}{\line(1,0){.5}}
\put(134,11){$=\frac 1 2$}
\put(149,3.5){\line(0,1){20.5}}
\put(147,0){\makebox(4,4){$\nabla$}}
\put(149,18){\circle*{1.5}}
\put(151.5,17){\tiny{$R_B$}}
\put(149,8){\circle*{1.5}}
\put(151.5,7){\tiny{$U$}}
\end{picture}
\end{array}
\end{equation}
with $R_B=UW_BU^\dag$, where the single-qubit gate $U_{ij}$ in (\ref{ij B state index}) is denoted as the $W_B^T$ gate to avoid a possible notational confusion. Furthermore,
the algebraic counterpart of the topological configuration (\ref{gate: two}) of the fault-tolerant construction of the Yang--Baxter gate $B$ which is a two-qubit Clifford
gate, is the matrix element of the teleportation equation (\ref{example two qubit gate via B}), expressed as
\begin{eqnarray}
\label{example two qubit gate as teleportation}
\setlength{\unitlength}{0.7mm}
\begin{array}{c}
\begin{picture}(150,60)
\put(80,50){\makebox(9,8){$(\tiny{\langle i_1j_1|\otimes 1\!\!1_2\otimes 1\!\!1_2\otimes \langle i_2j_2|)(B\otimes B\otimes B)(1\!\!1_2\otimes B\otimes B \otimes1\!\!1_2)(|\alpha\rangle\otimes |11\rangle\otimes |11\rangle\otimes |\beta\rangle)}$}}
\put(0,22){\makebox(14,10){$=$}}
\put(14,9){\makebox(4,4){$\nabla$}}
\put(16,12.5){\line(0,1){32.5}}
\put(26,36){\circle*{1.5}}
\put(18,33){\makebox(6,6){\tiny{$W^T_B$}}}
\put(26,9){\line(0,1){36}}
\multiput(12,27)(1,0){58}{\line(1,0){.5}}
\put(36,9){\line(0,1){7}}
\put(36,20){\line(0,1){25}}
\put(34,16){\line(0,1){4}}
\put(48,16){\line(0,1){4}}
\put(34,16){\line(1,0){14}}
\put(34,20){\line(1,0){14}}
\put(39,16){\makebox(4,4){\tiny{$B$}}}
\put(26,9){\line(1,0){10}}
\put(16,45){\line(1,0){10}}
\put(46,9){\line(0,1){7}}
\put(46,20){\line(0,1){25}}
\put(46,9){\line(1,0){10}}
\put(56,9){\line(0,1){36}}
\put(56,45){\line(1,0){10}}
\put(66,12.5){\line(0,1){32.5}}
\put(66,36){\circle*{1.5}}
\put(59,34){\makebox(4,4){\tiny{$W^T_B$}}}
\put(64,9){\makebox(4,4){$\nabla$}}
\put(70,22){\makebox(14,10){$=~\frac 1 4$}}
\put(86,9){\makebox(4,4){$\nabla$}}
\put(88,29){\line(0,1){16}}
\put(88,12.5){\line(0,1){12.5}}
\put(96,9){\makebox(4,4){$\nabla$}}
\put(86,25){\line(0,1){4}}
\put(100,25){\line(0,1){4}}
\put(86,25){\line(1,0){14}}
\put(86,29){\line(1,0){14}}
\put(91,25){\makebox(4,4){\tiny{$B$}}}
\put(98,29){\line(0,1){16}}
\put(98,12.5){\line(0,1){12.5}}
\put(88,18){\circle*{1.5}}
\put(80,15){\makebox(6,6){\tiny{$W_B$}}}
\put(98,18){\circle*{1.5}}
\put(101,16){\makebox(4,4){\tiny{$W_B^T$}}}
\put(106,22){\makebox(14,10){$=~\frac 1 4$}}
\put(124,9){\makebox(4,4){$\nabla$}}
\put(126,29){\line(0,1){16}}
\put(126,12.5){\line(0,1){12.5}}
\put(134,9){\makebox(4,4){$\nabla$}}
\put(124,25){\line(0,1){4}}
\put(138,25){\line(0,1){4}}
\put(124,25){\line(1,0){14}}
\put(124,29){\line(1,0){14}}
\put(129,25){\makebox(4,4){\tiny{$B$}}}
\put(136,29){\line(0,1){16}}
\put(136,12.5){\line(0,1){12.5}}
\put(126,36){\circle*{1.5}}
\put(119,34){\makebox(4,4){\tiny{$Q_B$}}}
\put(136,36){\circle*{1.5}}
\put(139,34){\makebox(4,4){\tiny{$P_B$}}}
\put(133.5,0){\makebox(9,8){\footnotesize{$|\beta\rangle$}}}
\put(120.5,0){\makebox(9,8){\footnotesize{$|\alpha\rangle$}}}
\put(95.5,0){\makebox(9,8){\footnotesize{$|\beta\rangle$}}}
\put(82.5,0){\makebox(9,8){\footnotesize{$|\alpha\rangle$}}}
\put(61.5,0){\makebox(9,8){\footnotesize{$|\beta\rangle$}}}
\put(11.5,0){\makebox(9,8){\footnotesize{$|\alpha\rangle$}}}
\end{picture}
\label{gate: two2}
\end{array}
\end{eqnarray}
where the single-qubit gate $W_B^T$ acting on the second qubit depends on the indices $i_1$, $j_1$ and
the single-qubit gate $W_B^T$ on the sixth qubit depends on the indices $i_2$, $j_2$ and the single-qubit
gates $Q_B$ and $P_B$ are derived by the formula~(\ref{Q P for B}).
\subsection{The teleportation operator and the teleportation equation}
\label{The teleportation operator and the teleportation equation}
As shown up in Section~\ref{YBG approach to tele based QC}, the teleportation operator (\ref{tele operator}) plays the key role in the Yang--Baxter gate approach
to teleportation-based quantum computation, and it is accompanied with both the product state preparation and the product state measurement to perform the teleportation protocol. In this subsection, we study the diagrammatical representation of the teleportation operator $(B\otimes 1\!\!1_2)(1\!\!1_2\otimes B)$ by combining the configuration of the product basis with the extended Temperley--Lieb diagrammatical approach, and from such the diagrammatical representation, we can easily derive the teleportation equation of the type~(\ref{example tele eqa B1}).
With the extended Temperley--Lieb configuration (\ref{TL_identity}) of the two-qubit identity gate, the Yang--Baxter gate $B$ (\ref{YBG B})
has the configuration
\begin{equation}
\label{example TL B Bell transform expansion}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(20,40)
\put(-115,-2){\footnotesize{$B$}}
\put(-121,29){\line(0,1){6}}
\put(-121,11){\line(0,-1){6}}
\put(-103,29){\line(0,1){6}}
\put(-103,11){\line(0,-1){6}}
\put(-124,8){\line(0,1){24}}
\put(-100,8){\line(0,1){24}}
\put(-124,8){\line(1,0){24}}
\put(-124,32){\line(1,0){24}}
\put(-121,29){\line(1,-1){18}}
\put(-121,11){\line(1,1){7.3}}
\put(-103,29){\line(-1,-1){7.3}}
\put(-93,19){$=\frac 1 {\sqrt 2}($}
\put(-58,23){\line(0,1){12}}
\put(-68,23){\line(0,1){12}}
\put(-68,23){\line(1,0){10}}
\put(-68,5){\line(0,1){12}}
\put(-58,5){\line(0,1){12}}
\put(-68,17){\line(1,0){10}}
\put(-53,19){$+$}
\put(-32,23){\line(0,1){12}}
\put(-42,23){\line(0,1){12}}
\put(-42,23){\line(1,0){10}}
\put(-32,29){\circle*{2.}}
\put(-30,28){\tiny{$Z$}}
\put(-42,5){\line(0,1){12}}
\put(-32,5){\line(0,1){12}}
\put(-42,17){\line(1,0){10}}
\put(-32,11){\circle*{2.}}
\put(-30,10){\tiny{$Z$}}
\put(-26,19){$+$}
\put(-6,23){\line(0,1){12}}
\put(-16,23){\line(0,1){12}}
\put(-16,23){\line(1,0){10}}
\put(-6,29){\circle*{2.}}
\put(-4,28){\tiny{$X$}}
\put(-16,5){\line(0,1){12}}
\put(-6,5){\line(0,1){12}}
\put(-16,17){\line(1,0){10}}
\put(-6,11){\circle*{2.}}
\put(-4,10){\tiny{$X$}}
\put(0,19){$+$}
\put(20,23){\line(0,1){12}}
\put(10,23){\line(0,1){12}}
\put(10,23){\line(1,0){10}}
\put(20,29){\circle*{2.}}
\put(22,28){\tiny{$XZ$}}
\put(10,5){\line(0,1){12}}
\put(20,5){\line(0,1){12}}
\put(10,17){\line(1,0){10}}
\put(20,11){\circle*{2.}}
\put(22,10){\tiny{$ZX$}}
\put(24,19){$+$}
\put(44,23){\line(0,1){12}}
\put(34,23){\line(0,1){12}}
\put(34,23){\line(1,0){10}}
\put(44,29){\circle*{2.}}
\put(46,28){\tiny{$Z$}}
\put(34,5){\line(0,1){12}}
\put(44,5){\line(0,1){12}}
\put(34,17){\line(1,0){10}}
\put(48,19){$-$}
\put(68,23){\line(0,1){12}}
\put(58,23){\line(0,1){12}}
\put(58,23){\line(1,0){10}}
\put(58,5){\line(0,1){12}}
\put(68,5){\line(0,1){12}}
\put(58,17){\line(1,0){10}}
\put(68,11){\circle*{2.}}
\put(70,10){\tiny{$Z$}}
\put(72,19){$+$}
\put(92,23){\line(0,1){12}}
\put(82,23){\line(0,1){12}}
\put(82,23){\line(1,0){10}}
\put(92,29){\circle*{2.}}
\put(94,28){\tiny{$X$}}
\put(82,5){\line(0,1){12}}
\put(92,5){\line(0,1){12}}
\put(82,17){\line(1,0){10}}
\put(92,11){\circle*{2.}}
\put(94,10){\tiny{$ZX$}}
\put(96,19){$-$}
\put(116,23){\line(0,1){12}}
\put(106,23){\line(0,1){12}}
\put(106,23){\line(1,0){10}}
\put(116,29){\circle*{2.}}
\put(118,28){\tiny{$XZ$}}
\put(106,5){\line(0,1){12}}
\put(116,5){\line(0,1){12}}
\put(106,17){\line(1,0){10}}
\put(116,11){\circle*{2.}}
\put(118,10){\tiny{$X$}}
\put(124,19){$)$}
\end{picture}
\end{array}
\end{equation}
which is equivalent to a special case of the extended Temperley--Lieb configuration (\ref{TL B general}). Applying the algebraic relations (\ref{conjugation product state in TL configuration 1})-(\ref{conjugation product state in TL configuration 4}) on the configuration~(\ref{example TL B Bell transform expansion}),
the Yang--Baxter gate $B$ (\ref{YBG B}) has its compact diagrammatical representation
\begin{equation}
\label{diagram B Bell transform}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(210,40)
\put(15,-2){\footnotesize{$B$}}
\put(9,29){\line(0,1){6}}
\put(9,11){\line(0,-1){6}}
\put(27,29){\line(0,1){6}}
\put(27,11){\line(0,-1){6}}
\put(6,8){\line(0,1){24}}
\put(30,8){\line(0,1){24}}
\put(6,8){\line(1,0){24}}
\put(6,32){\line(1,0){24}}
\put(9,29){\line(1,-1){18}}
\put(9,11){\line(1,1){7.3}}
\put(27,29){\line(-1,-1){7.3}}
\put(37,19){$=$}
\put(62,23){\line(0,1){12}}
\put(52,23){\line(0,1){12}}
\put(52,23){\line(1,0){10}}
\put(62,29){\circle*{2.}}
\put(64,28){\tiny{$Z$}}
\put(52,5){\line(0,1){10}}
\put(62,5){\line(0,1){10}}
\put(49.9,15){\tiny{$\triangle$}}
\put(59.9,15){\tiny{$\triangle$}}
\put(70,19){$+$}
\put(92,23){\line(0,1){12}}
\put(82,23){\line(0,1){12}}
\put(82,23){\line(1,0){10}}
\put(92,29){\circle*{2.}}
\put(94,28){\tiny{$XZ$}}
\put(82,5){\line(0,1){10}}
\put(92,5){\line(0,1){10}}
\put(92,11){\circle*{2.}}
\put(94,10){\tiny{$X$}}
\put(79.9,15){\tiny{$\triangle$}}
\put(89.9,15){\tiny{$\triangle$}}
\put(100,19){$+$}
\put(122,23){\line(0,1){12}}
\put(112,23){\line(0,1){12}}
\put(112,23){\line(1,0){10}}
\put(122,29){\circle*{2.}}
\put(124,28){\tiny{$X$}}
\put(112,5){\line(0,1){10}}
\put(122,5){\line(0,1){10}}
\put(112,11){\circle*{2.}}
\put(114,10){\tiny{$X$}}
\put(109.9,15){\tiny{$\triangle$}}
\put(119.9,15){\tiny{$\triangle$}}
\put(130,19){$+$}
\put(152,23){\line(0,1){12}}
\put(142,23){\line(0,1){12}}
\put(142,23){\line(1,0){10}}
\put(142,5){\line(0,1){10}}
\put(152,5){\line(0,1){10}}
\put(142,11){\circle*{2.}}
\put(144,10){\tiny{$X$}}
\put(152,11){\circle*{2.}}
\put(154,10){\tiny{$X$}}
\put(139.9,15){\tiny{$\triangle$}}
\put(149.9,15){\tiny{$\triangle$}}
\put(160,19){$=\sum_{k,l=0}^1$}
\put(204,23){\line(0,1){12}}
\put(194,23){\line(0,1){12}}
\put(194,23){\line(1,0){10}}
\put(204,29){\circle*{2.}}
\put(206,28){\tiny{$V_{kl}$}}
\put(194,5){\line(0,1){10}}
\put(204,5){\line(0,1){10}}
\put(204,11){\circle*{2.}}
\put(205,10){\tiny{$X^l$}}
\put(194,11){\circle*{2.}}
\put(195,10){\tiny{$X^k$}}
\put(191.9,15){\tiny{$\triangle$}}
\put(201.9,15){\tiny{$\triangle$}}
\end{picture}
\end{array}
\end{equation}
where the single-qubit gate $V_{kl}$ is defined in~(\ref{B kl state index}), and the associated algebraic formulation of such the
configuration~(\ref{diagram B Bell transform}) is expressed as
\begin{equation}
\label{algebraic B Bell transform}
B=\sum_{k,l=0}^1 |\Psi_{V_{kl}}\rangle\langle kl|=\sum_{k,l=0}^1|\psi(k+l,k+1)\rangle\langle kl|,
\end{equation}
with $|\Psi_{V_{kl}}\rangle$ defined in (\ref{Psi_U}), which is obviously the defining relation~(\ref{B Bell transform index})
of the Bell transform $B$.
On the other hand, applying the algebraic relations (\ref{product state in TL configuration 1})-(\ref{product state in TL configuration 4}) on
the configuration (\ref{example TL B Bell transform expansion}) of the Yang--Baxter gate $B$ gives rise to its another compact configuration
\begin{equation}
\label{diagram B inverse Bell transform}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(210,40)
\put(15,-2){\footnotesize{$B$}}
\put(9,29){\line(0,1){6}}
\put(9,11){\line(0,-1){6}}
\put(27,29){\line(0,1){6}}
\put(27,11){\line(0,-1){6}}
\put(6,8){\line(0,1){24}}
\put(30,8){\line(0,1){24}}
\put(6,8){\line(1,0){24}}
\put(6,32){\line(1,0){24}}
\put(9,29){\line(1,-1){18}}
\put(9,11){\line(1,1){7.3}}
\put(27,29){\line(-1,-1){7.3}}
\put(37,19){$=$}
\put(62,25){\line(0,1){10}}
\put(52,25){\line(0,1){10}}
\put(49.9,23){\tiny{$\nabla$}}
\put(59.9,23){\tiny{$\nabla$}}
\put(52,5){\line(0,1){12}}
\put(62,5){\line(0,1){12}}
\put(52,17){\line(1,0){10}}
\put(70,19){$+$}
\put(92,25){\line(0,1){10}}
\put(82,25){\line(0,1){10}}
\put(92,29){\circle*{2.}}
\put(94,28){\tiny{$X$}}
\put(79.9,23){\tiny{$\nabla$}}
\put(89.9,23){\tiny{$\nabla$}}
\put(82,5){\line(0,1){12}}
\put(92,5){\line(0,1){12}}
\put(82,17){\line(1,0){10}}
\put(92,11){\circle*{2.}}
\put(94,10){\tiny{$X$}}
\put(100,19){$-$}
\put(122,25){\line(0,1){10}}
\put(112,25){\line(0,1){10}}
\put(112,29){\circle*{2.}}
\put(114,28){\tiny{$X$}}
\put(109.9,23){\tiny{$\nabla$}}
\put(119.9,23){\tiny{$\nabla$}}
\put(112,5){\line(0,1){12}}
\put(122,5){\line(0,1){12}}
\put(112,17){\line(1,0){10}}
\put(122,11){\circle*{2.}}
\put(124,10){\tiny{$ZX$}}
\put(130,19){$-$}
\put(152,25){\line(0,1){10}}
\put(142,25){\line(0,1){10}}
\put(152,29){\circle*{2.}}
\put(154,28){\tiny{$X$}}
\put(142,29){\circle*{2.}}
\put(144,28){\tiny{$X$}}
\put(139.9,23){\tiny{$\nabla$}}
\put(149.9,23){\tiny{$\nabla$}}
\put(142,5){\line(0,1){12}}
\put(152,5){\line(0,1){12}}
\put(142,17){\line(1,0){10}}
\put(152,11){\circle*{2.}}
\put(154,10){\tiny{$Z$}}
\put(160,19){$=\sum_{i,j=0}^1$}
\put(204,25){\line(0,1){10}}
\put(194,25){\line(0,1){10}}
\put(204,29){\circle*{2.}}
\put(205,28){\tiny{$X^j$}}
\put(194,29){\circle*{2.}}
\put(195,28){\tiny{$X^i$}}
\put(192.2,23){\tiny{$\nabla$}}
\put(202.2,23){\tiny{$\nabla$}}
\put(194,5){\line(0,1){12}}
\put(204,5){\line(0,1){12}}
\put(194,17){\line(1,0){10}}
\put(203.9,11){\circle*{2.}}
\put(205.9,10){\tiny{$U_{ij}$}}
\end{picture}
\end{array}
\end{equation}
where the single-qubit gate $U_{ij}$ is defined in (\ref{ij B state index}), and the associated algebraic expression is
\begin{equation}
\label{algebraic B inverse Bell transform}
B=\sum_{i,j=0}^1|ij\rangle\langle\Psi_{U_{ij}}|=\sum_{i,j=0}^1(-1)^j|ij\rangle\langle\psi(i+j,i)|,
\end{equation}
with $|\Psi_{U_{ij}}\rangle$ defined in (\ref{Psi_U}), from which the Yang--Baxter gate $B$ can be also viewed as the inverse of
the Bell transform from the Bell basis to the product basis. Note that in \cite{ZZ14} the inverse of the Bell transform with
the product basis measurement is regarded as the Bell measurement, hence the Yang--Baxter gate $B$ acted by the product state
can be viewed as the Bell measurement.
In the algebraic approach, the teleportation operator $(B\otimes 1\!\!1_2)(1\!\!1_2\otimes B)$ has the form
\begin{equation}
(B\otimes 1\!\!1_2)(1\!\!1_2\otimes B)=\sum_{i,j,k,l=0}^1(|ij\rangle\langle\Psi_{U_{ij}}|\otimes 1\!\!1_2)(1\!\!1_2\otimes
|\Psi_{V_{kl}}\rangle\langle kl| )
\end{equation}
where the left Yang--Baxter gate $B$ represents the inverse of the Bell transform (\ref{algebraic B inverse Bell transform}) with the configuration (\ref{diagram B inverse Bell transform}) and the right one denotes the Bell transform (\ref{algebraic B Bell transform}) with the configuration (\ref{diagram B Bell transform}).
Therefore, such the teleportation operator is calculated in the diagrammatical approach,
\begin{equation}
\label{TL for teleportation operator in Bell transform}
\setlength{\unitlength}{0.5mm}
\begin{array}{c}
\begin{picture}(30,65)
\put(-66,25){\line(0,1){6}}
\put(-66,7){\line(0,-1){6}}
\put(-48,25){\line(0,1){6}}
\put(-48,7){\line(0,-1){6}}
\put(-69,4){\line(0,1){24}}
\put(-45,4){\line(0,1){24}}
\put(-69,4){\line(1,0){24}}
\put(-69,28){\line(1,0){24}}
\put(-66,25){\line(1,-1){18}}
\put(-66,7){\line(1,1){7.3}}
\put(-48,25){\line(-1,-1){7.3}}
\put(-84,55){\line(0,1){6}}
\put(-84,37){\line(0,-1){6}}
\put(-66,55){\line(0,1){6}}
\put(-66,37){\line(0,-1){6}}
\put(-87,34){\line(0,1){24}}
\put(-63,34){\line(0,1){24}}
\put(-87,34){\line(1,0){24}}
\put(-87,58){\line(1,0){24}}
\put(-84,55){\line(1,-1){18}}
\put(-84,37){\line(1,1){7.3}}
\put(-66,55){\line(-1,-1){7.3}}
\put(-48,31){\line(0,1){30}}
\put(-84,31){\line(0,-1){30}}
\put(-40,29){$=$}
\put(-30,29){$\sum_{i,j,k,l=0}^1$}
\put(14.5,48){\line(0,1){12}}
\put(2,48){\line(0,1){12}}
\put(14.5,54){\circle*{2.}}
\put(15.5,53){\tiny{$X^j$}}
\put(2,54){\circle*{2.}}
\put(3,53){\tiny{$X^i$}}
\put(-.2,45){{\scriptsize$\nabla$}}
\put(12.3,45){\scriptsize{$\nabla$}}
\put(14.5,30){\circle*{2.}}
\put(16,29){\tiny{$U_{ij}$}}
\put(14.5,0){\line(0,1){12}}
\put(27,0){\line(0,1){12}}
\put(27,6){\circle*{2.}}
\put(28,5){\tiny{$X^l$}}
\put(14.5,6){\circle*{2.}}
\put(15.5,5){\tiny{$X^k$}}
\put(12.3,12){\tiny{$\triangle$}}
\put(24.8,12){\tiny{$\triangle$}}
\put(2,0){\line(0,1){37.5}}
\put(2,37.5){\line(1,0){12.5}}
\put(14.5,22.5){\line(0,1){15}}
\put(27,22.5){\line(0,1){37.5}}
\put(14.5,22.5){\line(1,0){12.5}}
\put(27,30){\circle*{2.}}
\put(28.5,29){\tiny{$V_{kl}$}}
\put(40,29){$=\sum_{i,j,k,l=0}^1$}
\put(94.5,48){\line(0,1){12}}
\put(82,48){\line(0,1){12}}
\put(94.5,54){\circle*{2.}}
\put(95.5,53){\tiny{$X^j$}}
\put(82,54){\circle*{2.}}
\put(83,53){\tiny{$X^i$}}
\put(79.8,45){{\scriptsize$\nabla$}}
\put(92.3,45){\scriptsize{$\nabla$}}
\put(94.5,0){\line(0,1){12}}
\put(107,0){\line(0,1){12}}
\put(107,6){\circle*{2.}}
\put(108,5){\tiny{$X^l$}}
\put(94.5,6){\circle*{2.}}
\put(95.5,5){\tiny{$X^k$}}
\put(92.3,12){\tiny{$\triangle$}}
\put(104.8,12){\tiny{$\triangle$}}
\put(82,0){\line(0,1){37.5}}
\put(82,37.5){\line(1,0){12.5}}
\put(94.5,22.5){\line(0,1){15}}
\put(107,22.5){\line(0,1){37.5}}
\put(94.5,22.5){\line(1,0){12.5}}
\put(107,54){\circle*{2.}}
\put(108.5,53){\tiny{$W_{1,1}$}}
\end{picture}
\end{array}
\end{equation}
where the single-qubit gate $W_{1,1}=V_{kl}U^T_{ij}$ expressed as
\begin{equation}
\label{W 1 1}
W_{1,1} = (-1)^{i\cdot j+(k+l)\cdot(i+k+1)}Z^{i+k+1}X^{i+j+k+l},
\end{equation}
can be also calculated from the formula~(\ref{W epsilon}) by setting $\epsilon=\eta=1$. See Table~\ref{Table W 1 1} for the explicit
expressions of the single-qubit gate $W_{1,1}$.
\begin{table}
\begin{center}
\footnotesize
\begin{tabular}{c|c|c|c|c}
\hline\hline
& $i=0$, $j=0$ & $i=0$, $j=1$ & $i=1$, $j=0$ & $i=1$, $j=1$ \\ \hline
$k=0$, $l=0$ & $Z$ & $-XZ$ & $X$ & $-1\!\!1_2$ \\ \hline
$k=0$, $l=1$ & $XZ$ & $-Z$ & $1\!\!1_2$ & $-X$ \\ \hline
$k=1$, $l=0$ & $X$ & $1\!\!1_2$ & $-Z$ & $XZ$ \\ \hline
$k=1$, $l=1$ & $1\!\!1_2$ & $X$ & $-XZ$ & $-Z$ \\ \hline\hline
\end{tabular}
\caption{\label{Table W 1 1} The single-qubit gate $W_{1,1}$ (\ref{W 1 1}) in both the algebraic expansion (\ref{algebraic tl tele operator product state sum}) of
the teleportation operator $(B\otimes 1\!\!1_2)(1\!\!1_2\otimes B)$ and the teleportation equation (\ref{example tele eqa B}).}
\end{center}
\end{table}
Furthermore, the algebraic counterpart of the diagram~(\ref{TL for teleportation operator in Bell transform}) can be rewritten as
\begin{equation}
\label{algebraic tl tele operator product state sum}
(B\otimes 1\!\!1_2)(1\!\!1_2\otimes B)=\sum_{i,j,k,l=0}^1(|ij\rangle, W_{1,1}, \langle kl|),
\end{equation}
with the algebraic notation $(|ij\rangle, W_{1,1}, \langle kl|)$ of the associated diagrammatical term, which gives rise to
the teleportation equation
\begin{equation}
\label{example tele eqa B}
(B\otimes 1\!\!1_2)(1\!\!1_2\otimes B)|\alpha\rangle\otimes |kl\rangle=\frac 1 2\sum_{i,j=0}^1|ij\rangle W_{1,1}|\alpha\rangle,
\end{equation}
with the teleportation equation~(\ref{example tele eqa B1}) as its special example of $k=l=1$. In this sense, each one of the total 16 teleportation
configurations in the diagram~(\ref{TL for teleportation operator in Bell transform}) can be extracted by applying the product basis measurement on
the teleportation operator $(B\otimes 1\!\!1_2)(1\!\!1_2\otimes B)$. In addition, Appendix E presents another equivalent method of deriving the
algebraic structure (\ref{algebraic tl tele operator product state sum}) of the teleportation operator $(B\otimes 1\!\!1_2)(1\!\!1_2\otimes B)$.
\begin{figure}
\begin{center}
\includegraphics[width=13cm]{tl_tele_operator_Bell_measurement}
\end{center}
\caption{\label{fig_tl_tele_operator_Bell_measurement} The extended Temperley--Lieb configuration of the teleportation operator $(B\otimes 1\!\!1_2)(1\!\!1_2\otimes B)$
from the viewpoint of Bell measurements. The result is derived by directly applying the extended Temperley--Lieb configuration~(\ref{example TL B Bell transform expansion})
of the Yang--Baxter gate $B$ (\ref{YBG B}). The vertical braiding configuration can be regarded as the one obtained by clockwise rotating the horizontal braiding
configuration in the quantum circuit in Figure~\ref{fig_tele_YBG} by 90 degrees. }
\end{figure}
\subsection{The teleportation operator using Bell measurements }
The extended Temperley--Lieb diagrammatical representation of the teleportation operator $(B\otimes 1\!\!1_2)(1\!\!1_2\otimes B)$
can be derived in a rather straightforward diagrammatical approach using the extended Temperley--Lieb configuration~(\ref{example TL B Bell transform expansion})
of the Yang--Baxter gate $B$, and the result is shown up in Figure~\ref{fig_tl_tele_operator_Bell_measurement}, which can be explicitly regarded as a kind of
linear combination of 16 typical quantum teleportation processes using Bell measurements.
Although each one of these 16 diagrammatical teleportation terms in Figure~\ref{fig_tl_tele_operator_Bell_measurement} can be
interpreted in the viewpoint of quantum teleportation, a linear combination of these diagrammatical terms can not be usually viewed as quantum teleportation.
The reason is that the teleportation operator with the product basis measurement (instead of Bell measurements) gives rise to quantum teleportation,
as discussed in Subsection~\ref{The teleportation operator and the teleportation equation}. Note that the simplified version of
Figure~\ref{fig_tl_tele_operator_Bell_measurement} is shown up in Figure~\ref{fig_tl_tele_operator}.
Appendix~\ref{algebraic_study_fig_tl_tele_operator} presents an algebraic method of deriving the extended Temperley--Lieb configuration
in Figure~\ref{fig_tl_tele_operator}, and it is obvious that the topological configuration in Figure~\ref{fig_tl_tele_operator_Bell_measurement} can
be easily obtained from the configuration in Figure~\ref{fig_tl_tele_operator}.
It is worthwhile pointing out that such the extended Temperley--Lieb configurations in both the diagram~(\ref{TL for teleportation operator in Bell transform})
and Figure~\ref{fig_tl_tele_operator} come naturally from our topological and algebraic reformulations
of teleportation-based quantum computation, whereas they are indeed unexpected if we consider the standard low dimensional
topology \cite{Kauffman02}. Interested readers are invited to refer to Appendix A.
\section{Concluding remarks}
\label{Concluding remarks}
In this paper, we reformulate teleportation-based quantum computation \cite{GC99,Nielsen03,Leung04} in both the extended Temperley--Lieb diagrammatical
approach \cite{Kauffman05,Zhang06} and the Yang--Baxter gate approach \cite{Dye03, KL04, ZKG04}. Such the two approaches can be respectively regarded as
the topological aspect and algebraic aspect of a unified approach. On the other hand, through our research, the Yang--Baxter gate
configuration (the braiding configuration) admits an equivalent description of a set of the extended Temperley--Lieb diagrammatical configurations,
so we finally propose the extended Temperley--Lieb diagrammatical approach as an interesting topic for physicists in quantum information and computation.
Our results show that the fact that quantum entanglement (or quantum non-locality) admits a kind of interpretation of topological
entanglement (topological non-locality) takes the responsibility for topological
features in teleportation-based quantum computation. Such topological features regard teleportation-based quantum circuit models as the two-dimensional
topological deformations of the extended Temperley--Lieb diagrammatical configurations, and they greatly simplify the algebraic analysis in
teleportation-based quantum computation.
About further research, since teleportation-based quantum computation is an example for measurement-based quantum computation which includes
the one-way quantum computation \cite{RB01, BFN08} as another example, so we expect that the one-way quantum computation \cite{RB01} can be also
understood from both the extended Temperley--Lieb diagrammatical approach and the Yang--Baxter gate approach. Furthermore, if we consider the categorical
description \cite{Coecke04, AC04, Abramsky09} of quantum teleportation, it is no doubt to obtain new insights on both our research results in this paper
and categorical quantum information and computation. In addition, further research problems in mathematical physics are collected in
Appendix~\ref{YBE_further_research}.
\begin{figure}
\begin{center}
\includegraphics[width=13cm]{tl_tele_operator}
\end{center}
\caption{\label{fig_tl_tele_operator} The extended Temperley--Lieb configuration of the teleportation operator $(B\otimes 1\!\!1_2)(1\!\!1_2\otimes B)$ as
a simplified version of the configuration in Figure~\ref{fig_tl_tele_operator_Bell_measurement}. }
\end{figure}
\section*{Acknowledgements}
Yong Zhang is supported by the starting grant--273732 of Wuhan University.
|
\section{\label{sec:level1}Introduction}
Graphene -- a two-dimensional allotrope of carbon, has attracted considerable attention in recent years,
largely due to its remarkable electronic properties stemming from the massless-Dirac-fermion nature of
its low-energy quasiparticle states.~\cite{2009RvMP...81..109C}
Most of the electronic properties of graphene that have been studied
experimentally can be well described by non-interacting single-particle theory.
However, electron-electron interactions in graphene are expected to be strong.
In undoped clean graphene, the density of states
at the Fermi level vanishes and therefore the Coulomb potential
is not screened.~\cite{2007PhT....60h..35G,RevModPhys.84.1067}
Recent experiments have shown that unscreened Coulomb interactions lead to reshaping of
the ideal conical energy dispersion expected in graphene.~\cite{2011NatPh...7..701E} More precisely, the Fermi
velocity near the Dirac point acquires a logarithmic correction as a result
of interactions.
From a theoretical viewpoint, it can be shown that this logarithmic enhancement arises from
the non-local exchange interaction,
already at the level of the first-order Hartree-Fock perturbation theory.~\cite{katsnelson2012graphene}
Hence, it is the long-range nature of the electron-electron interactions in graphene
that is responsible for the logarithmic correction, the most striking interaction
effect observed so far in this material in the absence of external magnetic fields.
As a result, theoretical work has mostly focused on investigations of long-range interactions in graphene,
using a variety of techniques ranging from mean field~\cite{PhysRevB.84.085446} to renormalization group
approaches.~\cite{Gonzalez1994595, PhysRevB.59.R2474}
It should be noted, however, that there are several important conditions that need to be
satisfied in order to observe significant long-range interaction effects
experimentally. It is necessary to
be able to probe a wide range of carrier concentrations and
to tune the Fermi level sufficiently close the Dirac point,
where the renormalization of the Fermi velocity is expected to be dramatic due to
the vanishing density of states. Moreover, spurious screening effects, e.g.
dielectric screening from the substrate, should be avoided.
This makes undoped high-quality suspended graphene
an ideal platform for studying the effects
of long-range electron-electron interactions.~\cite{2011NatPh...7..701E}
On the contrary, in the case of graphene on a substrate
or in the presence of disorder and impurities these
effects are less relevant.
In particular, doping introduces a finite density of states at the Dirac
point of graphene and the long-range part of
the Coulomb potential is screened. In this case,
short-range interactions become crucial. If these interactions
are fairly strong, they can lead to interesting effects
on the electronic structure, especially
on the impurity states in the vicinity of the Fermi level.
In fact, estimates of the on-site Hubbard $U$ parameter
in carbon-based molecules~\cite{Parr1950,PhysRevLett.56.1509} suggest that
the short-range Coulomb interactions
among $\pi$-electron in graphene can be indeed quite large, i.e.
of the order of $10$~eV. A similar value is obtained
by accurate \textit{ab initio} calculations.~\cite{PhysRevLett.106.236805}
In this paper, we study the effects of short-range electron-electron interactions
on the electronic structure of doped graphene. We should note
that the importance of impurity effects in graphene has been addressed in many theoretical
studies.~\cite{PhysRevB.73.241402,PhysRevB.86.045448,PhysRevB.80.214201,PhysRevLett.92.256805,PhysRevB.73.125414,
PhysRevB.84.245446,PhysRevB.85.035404} A number of
interesting features have been revealed, including the opening of the
gap upon doping~\cite{PhysRevB.80.214201,PhysRevB.86.045448} and the appearance of impurity (acceptor or donor) states in the vicinity of
the Fermi level.~\cite{PhysRevB.73.241402,PhysRevB.84.245446} There is also a great interest in
addressing these properties experimentally~\cite{Zhao19082011,doi:10.1021/nn1002425,PhysRevB.85.161408}
since doping graphene with impurities is one way to further explore and tune its
electronic, magnetic and transport properties.
However, the interplay of short-range electron-electron interactions
and impurity potentials in graphene has not yet been fully explored.
We use a single-band ($\pi$ orbital) tight-binding (TB) model to describe the electronic structure
of graphene. A supercell method is employed to
study the effects of finite doping. A substitutional impurity
is introduced in the TB Hamiltonian as a local modification of the on-site
potential at the impurity site. Here we focus on attractive impurity
potentials, mimicking nitrogen impurity atoms which are typical
dopants in graphene.~\cite{Zhao19082011} The short-range interactions are described by means of the
Hubbard model in the mean-field approximation, which is the simplest way of
treating the many-body interacting problem.
Interaction terms are introduced at each site in the TB Hamiltonian.
We use a numerical self-consistent scheme to
account for a redistribution of electronic charge
around the impurity caused by interactions. As the output of our numerical
calculations, we
obtain the band structure and the local density of states (LDOS) around the impurity site.
Furthermore, we calculate scanning tunneling microscopy (STM) images
by integrating the LDOS over a small energy window above the Fermi level.
By using this approach, we show that
short-range interactions introduce several remarkable
features in the electronic structure of doped graphene.
Importantly, they enhance the resonant character of states
localized in real space around the impurity, which are induced in the vicinity of the Dirac point. The complex
interplay between short-range interactions and impurity potential is also responsible for non-trivial gaps at
high-symmetry crossing points in the band structure of graphene, in particular at the Dirac point.
The paper is organized as follows. In Sec.~\ref{sec:level2} we introduce our TB model and
describe how the impurity potential and short-range interactions are incorporated in the Hamiltonian.
We also provide some details of the self-consistent supercell calculations.
Our findings are described in Sec.~\ref{sec:level3}. In particular,
in Sec.~\ref{bands} we focus on the effects of interactions on the band structure of graphene
and on some issues related to the supercell geometry. In Sec.~\ref{ldos} we discuss the changes in
the resonant character of the LDOS around the impurity for varying impurity potential strength and interaction
strength. The comparison between the simulated STM topographies
for non-interacting and interacting cases is provided. Finally, we draw some conclusions.
\section{\label{sec:level2}Methodology}
The second-quantized Hamiltonian for interacting electrons on a honeycomb lattice in the presence
of impurity can be written as
\begin{equation}\label{eq:HamiltonianForTheSystem}
H={\sum_{i{{\sigma}}}{\varepsilon}_{i}c^{\dagger}_{i{\sigma}}c_{i{\sigma}}}+{\sum_{\left\langle
{i,j}\right\rangle{\sigma}}}{t_{ij} c^{\dagger}_{i{\sigma}}c_{j{\sigma}}}+U_{\mathrm{im}}{c^{\dagger}_{0{\sigma}}c_{0{\sigma}}}+U{\sum_{i}}{n_{i{\uparrow}}n_{i{\downarrow}}}.
\end{equation}
Here $c^{\dagger}_{i{\sigma}}$ and $c_{i{\sigma}}$ are the creation and annihilation operators
for electron on site $i$ and with spin $\sigma$;
${\varepsilon}_{i}$ and $t_{ij}$ are on-site energies and hopping parameters, respectively. Only
hopping between nearest neighbors on the honeycomb lattice is included. We
assume that the TB parameters are uniform, except for the
on-site energy at the impurity site, and we use the values
obtained by fitting the TB band structures to density functional theory calculations,
namely $\varepsilon_{i}=0$ and $t_{ij}=-2.97$~eV.~\cite{PhysRevB.66.035412}
The third term in Eq.~(\ref{eq:HamiltonianForTheSystem}) represents the local
impurity potential, with $U_{\mathrm{im}}$ being the impurity potential strength
($U_{\mathrm{im}}<0$ for attractive impurity). In our calculations we
use $U_{\mathrm{im}}=-10$~eV and $U_{\mathrm{im}}=-20$~eV in order
to obtain visible trends for the impurity states in the vicinity of the Fermi level.
The last term describes
the on-site interaction between two electrons with opposite spins on site $i$ (including the impurity site),
with $U$ ($U\ge 0$) being the Hubbard $U$ parameter, which expresses the strength of the
intra-atomic Coulomb repulsion. Here $n_{i{\sigma}}$ is the number operator,
defined as $n_{i{\sigma}}=c^{\dagger}_{i{\sigma}}c_{i{\sigma}}$. We consider
$U=0$, or non-interacting case, and $U=9.3$~eV,
which is the value obtained for graphene using the constrained Random Phase Approximation method.~\cite{PhysRevLett.106.236805}
In order to extract the trends in the electronic structure with
increasing the interaction strength we also use a larger value of $U=20$~eV.
In the mean-field approximation, the two-body interaction term in Eq.~(\ref{eq:HamiltonianForTheSystem}) becomes
\begin{equation}\label{eq:HamiltonianInMeanField}
U{\sum_{i}}{n_{i{\uparrow}}n_{i{\downarrow}}}=
U{\sum_{i}}{\left({\left\langle {n_{i{\downarrow}}}\right\rangle}c^{\dagger}_{i{\uparrow}}c_{i{\uparrow}}
+{\left\langle {n_{i{\uparrow}}}\right\rangle}c^{\dagger}_{i{\downarrow}}c_{i{\downarrow}}\right)},
\end{equation}
where $\left\langle {n_{i{\sigma}}}\right\rangle$ is the average electron occupation number, or density, for
spin-up ($\sigma=\uparrow$) and spin-down ($\sigma=\downarrow$) electrons. Here we consider
a non spin-polarized case so that ${\left\langle {n_{i{\uparrow}}}\right\rangle}={\left\langle {n_{i{\downarrow}}}\right\rangle}$.
In pristine graphene with the Fermi level exactly at the Dirac point, the average electron occupation number is a constant
equal to $1/2$. Adding a mean-field field on-site potential does
not break the translational invariance of the crystal
and the average occupation number remains constant. In fact, in orthogonal basis such a potential merely introduces
a rigid shift of the energy bands (note that in non-orthogonal basis
the interplay between the overlap integrals and the on-site interactions
can lead to renormalization of the Fermi velocity~\cite{unpbl}).
However, the presence of both mean-field on-site interactions and
impurity potential can lead to non-trivial effects in the electronic
structure. In this case,
the potential at each site
depends on the average occupation number ${\left\langle {n_{i{\sigma}}}\right\rangle}$, which is not
necessarily the same on all sites. As a result, when a carbon atom is replaced by an impurity,
there will be a redistribution of electronic charge in the system. In order to capture this effect,
we need to perform self-consistent calculations for the Hamiltonian in Eq.~(\ref{eq:HamiltonianForTheSystem})
and (\ref{eq:HamiltonianInMeanField}).
At each step of the self-consistent cycle,
the average occupation number for site $i$ is calculated as
\begin{equation}
\left\langle {n_{i{\sigma}}}\right\rangle={\frac{1}{N}}{\sum_{k}^{occ}}{\left|{b^{k}_{i{\sigma}}}\right|^{2}},
\end{equation}
where $N$ is the number of $k$-points in the Brillouin zone and $b^{k}_{i{\sigma}}$ are
the coefficients in the expansions of the wavefunctions of the Hamiltonian in terms of
the localized atomic orbitals $\left| i\sigma \right\rangle$. These
are obtained by diagonalization of the Hamiltonian at each $k$-point.
The sum
runs over all occupied states up to the Fermi level. Note that all calculations
are done at half-filling.
As initial values we use the occupation numbers calculated for a non-interacting problem,
i.e. for a supercell of graphene with impurity ($U_{\mathrm{im}}\ne 0$) and with $U=0$.
The criterion of self-consistency is
\begin{equation}
{\sum_{i{\sigma}}}\left|{{ \left\langle n_{i{\sigma}}\right\rangle }^{s}-{\left\langle n_{i{\sigma}}\right\rangle}^{s-1}}\right|<{\eta},
\end{equation}
where $s$ is the index of the self-consistent cycle and ${\eta}$ is a small parameter (we choose ${\eta}=10^{-7}$).
We use a linear mixing scheme, in which the input density
$\left\langle n_{i\sigma} \right\rangle^{s+1}_{\mathrm{in}}$ at step $s+1$
is calculated as a linear combination of outputs $\left\langle n_{i\sigma} \right\rangle^{s}_{\mathrm{out}}$ and $\left\langle n_{i\sigma} \right\rangle^{s-1}_{\mathrm{out}}$
from two previous steps
\begin{equation}
\left\langle n_{i\sigma} \right\rangle^{s+1}_{\mathrm{in}}=\left({1-{\lambda}}\right)\left\langle n_{i\sigma} \right\rangle^{s-1}_{\mathrm{out}}+
{\lambda}\left\langle n_{i\sigma} \right\rangle^{s}_{\mathrm{out}},
\end{equation}
where ${\lambda}$ is the mixing coefficient; we use ${\lambda}=0.25$, which
allows us to achieve self-consistency in less than $100$ steps.
\begin{figure}
\centering
\includegraphics[width=0.97\linewidth]{fig0.png}
\caption{(Color online) (a) $6\times6$ supercell of graphene with a substitutional impurity. A magenta sphere
represents the impurity atom. Dashed lines mark the unit cell of pristine graphene and
arrows show the primitive lattice vectors. A and B denote carbon atoms in the two equivalent sublattices.
(b) Brillouin zone folding in graphene. Shaded area (1) represents the first Brillouin zone
of a $p\times p$ supercell of graphene. Numbered curves correspond to the first Brillouin zone of the
unit cell for different $p$: $p=3q-1$ (2), $p=3q$ (3) and $p=3q+1$. For $p=3q$ the Dirac points of graphene ($K$, $K^{\prime}$)
are mapped onto $\Gamma$ point of the folded Brillouin zone. In other cases, $K$ and $K^{\prime}$
are mapped onto $K$ and $K^{\prime}$ of the folded Brillouin zone.}
\label{fig:00}
\end{figure}
In order to
model the effect of finite impurity concentration,
we construct a $p\times p$ supercell
by replicating a graphene unit cell $p$ time along each of the two-dimensional lattice vectors
[see Fig.~\ref{fig:00}(a)].
The impurity atom substitutes a carbon atom in the supercell.
In this work, we use two different supercells with $p=6$ and $p=7$.
Atomic concentration of the dopants depends
on the size of the supercell so the concentration is slightly different for the two choices,
namely 1.0$\%$ for a $7\times 7$ and 1.4$\%$ for a $6\times 6$ supercell.
It is known that for $p=3 q$, where $q$ is an integer, the Dirac points of graphene, $K$ and $K^{\prime}$,
are mapped onto the $\Gamma$ point of the Brillouin zone of the supercell.~\cite{0957-4484-24-22-225705,PhysRevB.86.045448,PhysRevB.81.245420}
as illustrated in Fig.~\ref{fig:00}(b).
This does not happen
if $p$ is not divisible by $3$. Therefore,
the $6\times 6$ supercell is special. As we explain in Sec.~\ref{bands}, the effects of impurity potential
and interactions in this case are rather non-trivial.
This is the main reason for considering two different supercell sizes.
\section{\label{sec:level3}Results}
\subsection{\label{bands}Bandstructure}
It is known that the finite amount of doping opens up
an energy gap at the Dirac point of graphene.~\cite{PhysRevB.86.045448,PhysRevLett.92.256805,PhysRevB.84.245446}
Here we address the question of how the details of the bandstructure near the gap are affected by interactions.
We start with a special supercell geometry $p\times p$, with $p$ divisible by $3$ ($p=6$ in our calculations).
Figure~\ref{fig:02}(a)-(b) shows the bandstructure of the $6\times 6$ supercell with impurity potential $U_{\mathrm{im}}=-10$~eV
and $U_{\mathrm{im}}=-20$~eV, respectively, for three values of the interaction strength, $U=0$,
$U=9.3$~eV and $U=20$~eV.
Note that in the bandstructure calculations, different impurity potential and interaction strength
introduce a shift of the energy bands with respect to a reference case, i.e. non-interacting
pristine graphene ($U_{\mathrm{im}}=0$ and $U=0$). In order to examine the features around the gap
for different choices of parameters, we align the position of the doubly degenerate
state (see the discussion below) in all curves in Fig.~\ref{fig:02}(a)-(b) to the value found for $U=0$ for a given impurity potential strength.
\begin{figure}
\centering
\includegraphics[width=0.97\linewidth]{fig1.jpg}
\caption{(Color online) Bandstructure of doped graphene along the high-symmetry lines of the
Brillouin zone for $6\times 6$ (a,b) and $7\times 7$ (c,d) supercell, for
varying impurity potential strength $U_{\mathrm{im}}$ and interaction strength $U$. Left panels
are for $U_{\mathrm{im}}=-10$~eV, right panels for $U_{\mathrm{im}}=-20$~eV. In each panel
three cases are shown: $U=0$ (black), $U=9.3$~eV (red) and $U=20$~eV (blue). Horizontal line
in (a) and (b) is the position of the
doubly degenerate state at $\Gamma$, adjusted to the value found
for $U=0$ (see text for details). Horizontal line in (c) and (d) marks the
conduction band maximum, which has been aligned with the value found for $U=0$.}
\label{fig:02}
\end{figure}
As we mentioned before, in the case of $p=6$ both $K$ and $K^{\prime}$ are mapped
onto ${\Gamma}$ point,~\cite{0957-4484-24-22-225705,PhysRevB.86.045448,PhysRevB.81.245420}
producing four degenerate states at $\Gamma$ in the absence of impurities
and interactions. When one carbon atom in the supercell is substituted
by an impurity atom, a gap opens up between two states at $\Gamma$, however the other two states remain degenerate.
More precisely, for $U=0$ three of the four states at $\Gamma$ are degenerate while one of the states moves away from the Dirac point.
This situation is referred to as the \textit{pseudogap}~\cite{PhysRevB.86.045448} since there
is still a pair of linearly dispersed states crossing at the neutrality point. Hence, effectively there is no band gap
for this special supercell size.
One can clearly see from Fig.~\ref{fig:02}(a)
that the size of the pseudogap decreases with increasing the interaction strength.
On-site interactions cause a redistribution of charge around the impurity. We
find that in the case of the attractive impurity potential $U_{\mathrm{im}}=-10$~eV,
the total average occupation at the impurity site decreases from $\left\langle n_{0} \right\rangle=0.89$
in the non-interacting case to $\left\langle n_{0} \right\rangle=0.82$ for $U=9.3$~eV.
The on-site Coulomb repulsion prevents extra charge from accumulating on the impurity and
therefore the strength of the impurity potential is effectively reduced.
For the impurity potential which is twice stronger, the pseudogap
for the same values of $U$ is noticeably larger [see Fig.~\ref{fig:02}(b)].
In addition to a large pseudogap, for $U\ne 0$ there is also a smaller pseudogap which opens up at $\Gamma$ [see the insets in Fig.~\ref{fig:02}(a)-(b)].
There are now only two generate states at $\Gamma$, while the other two states shift, respectively, above and below the crossing point.
Interestingly, the effect of interactions on the small pseudogap
is opposite to that on the large one, e.g. its value increases with increasing the interaction strength.
A perturbative model described in Appendix~\ref{sec:appendix} suggests that the
smaller gap results from the contribution of the states localized on
sublattice B, if we assume that the impurity is substituted in sublattice A.
Within our model, this effect is solely due to interactions (the contribution of sublattice B to the gap is zero in the absence of interactions~\cite{PhysRevB.86.045448}).
We find the following values for the two gaps at $\Gamma $ from analytical calculations (see Appendix \ref{sec:appendix} for details).
In the non-interacting case and $U_{\mathrm{im}}=-10$~eV, the large pseudogap is -0.56~eV. For $U=9.3$~eV, the large pseudogap decreases
to 0.39~eV. At the same time, a small pseudogap of 0.05~eV opens up. For $U=20$~eV, the large and small pseudogaps
become 0.28~eV and 0.06~eV, respectively.
These values are all in good agreement with the pseudogaps found in Fig.~\ref{fig:02}(a).
Note that for the larger impurity potential, the trends with increasing $U$ are the same, however for a given $U$ both gaps are larger
than in $U_{\mathrm{im}}=-10$~eV case.
Analytical calculations using the perturbative model in this case also agree with numerical results.
Similar features are found for a regular $7\times 7$ supercell [see Fig.~\ref{fig:02}(c) and (d)]. Note that the
conduction band minima at $\Gamma$ have been aligned with the reference $U=0$ case. The main difference from the
$6\times 6$ supercell is that in this case there
is a real band gap at $K$ ($K^\prime$). Our perturbative analysis shows that
there is no contribution from sublattice B to the gap at the Dirac point, if we assumed that the impurity
is substituted in sublattice A. As in the case of the $6\times 6$ supercell, the size of the gap decreases
with increasing the interaction strength. Analytically, for $U_{\mathrm{im}}=-10$~eV we find a gap of 0.20 eV for
$U=0$, 0.14 eV for $U=9.3$~eV and 0.10 eV for $U=20$~eV. These values are in good agreement with
numerical calculations. Somewhat smaller values of the gaps compared to a $6\times 6$ supercell
for the same $U_{\mathrm{im}}$ and $U$ are expected since the atomic concentration of impurities
is smaller.
Bandstructure calculations presented in this section lead to a conclusion that short-range interactions effectively reduce the strength
of the impurity potential, which results in a decrease of the large gap (pseudogap) at the Dirac point.
In oder to see how the character, e.g. the energy and the spatial extent, of the electronic states around the Dirac point
is affected by interactions, we need to look at the LDOS around the impurity.
\subsection{\label{ldos}Local density of states}
Calculations of LDOS at the impurity site reveal several important features.
A substitutional impurity introduces electronic states at energies
comparable to the impurity potential ($|U_\mathrm{im}|\sim 10$~eV), i.e.
far away from the Fermi level. However, there are also states
appearing in the vicinity (within $\sim 1$~eV) of the Fermi level.~\cite{PhysRevB.77.115109,PhysRevB.86.045448,
PhysRevB.73.241402,PhysRevB.75.125425}
These states are the
most relevant for the low-energy electronic properties
of graphene and will be examined in detail.
Figure~\ref{fig:01} shows the double- or multi-peak impurity resonances
close to the Fermi level for the $6\times 6$
and $7\times 7$ supercells, respectively, for different impurity potential and interaction strengths.
The multi-peak structure of the impurity resonances most likely originate
from the long-range interaction, or interference, between the impurity potentials, caused by the periodicity of the supercell
geometry.~\cite{PhysRevB.86.045448}
\begin{figure}
\centering
\includegraphics[width=0.97\linewidth]{fig2.jpg}
\caption{(Color online) LDOS of doped graphene at the impurity site for
$6\times 6$ (a,b) and $7\times 7$ (c,d) supercell, for
varying impurity potential strength $U_{\mathrm{im}}$ and interaction strength $U$. Left panels
are for $U_{\mathrm{im}}=-10$~eV, right panels for $U_{\mathrm{im}}=-20$~eV. In each panel
three cases are shown: $U=0$ (black), $U=9.3$~eV (red) and $U=20$~eV (blue).
Vertical lines mark the position of the Fermi level (see text for details).}
\label{fig:01}
\end{figure}
With increasing the impurity potential strength,
the resonant peaks move closer to low energies.
This is very similar to the case of strong potential impurities on the surface of a topological insulator.~\cite{PhysRevB.85.121103} As shown
in Fig.~\ref{fig:01}(a) and (b) with $U=0$, the double-peak resonance in the $6\times 6$ supercell approaches the Fermi level as
$U_{\mathrm{im}}$ increases. At the same time
its amplitude decreases.
This finding is in
agreement with semi-analytical calculations in
Refs.~\onlinecite{PhysRevB.77.115109,PhysRevB.73.241402,PhysRevB.75.125425}.
The effect of the impurity potential is more
striking in the case of a regular $7\times 7$ supercell
[Fig.~\ref{fig:01}(c) and (d)].
In this case, in addition
to shifting the peaks to lower energies,
a stronger impurity potential $U_{\mathrm{im}}=-20$~eV also
makes the peaks narrower, thus enhancing
their resonant character [see in particular the first peak in
Fig.~\ref{fig:01}(c) and (d) with $U=0$].
In the interacting case, the amplitude of the resonances increases with $U$.
This can be seen in all panels in Fig.~\ref{fig:01} with $U=9.3$~eV and $U=20$~eV,
with the exception of the $6\times 6$ supercell with $U_{\mathrm{im}}=-10$~eV and $U=20$~eV, where
the amplitude of the peak decreases slightly.
In the case of $U_{\mathrm{im}}=-10$~eV, for both supercells
the impurity resonances move further away from the Fermi level
with increasing $U$. This is perfectly consistent with our observations for
the non-interacting case with decreasing impurity potential.
However, in the case of a very large impurity potential $U_{\mathrm{im}}=-20$~eV,
the trend in the position of the resonances is less obvious.
The peaks either do not move appreciably as in the case of $U=20$~eV or
even seem to move slightly towards the low-energy region for $U=9.3$~eV.
Below we elaborate more on these findings.
Increasing $U$ reduces the overall strength of the impurity potential,
which is confirmed by decrease of the energy gap at the Dirac point due to
the presence of impurities (Sec.~\ref{bands}).
However, short-range electron-electron interactions controlled by $U$
do not only change the potential directly at the impurity site but
also affect the on-site potential and the charge density around the impurity (primarily nearest and
next-nearest neighbors of the impurity atom).
Hence, both the amplitude and the spatial extent of the impurity potential is
modified by interactions.
Let us assume that an attractive
impurity can be described by a delta-function potential well.
When interactions are included, the shape of
the impurity potential is smoothed out (it acquires, say, a Gaussian shape).
Therefore, although the strength of the potential
is reduced by a certain amount with increasing $U$,
the potential can become more long-ranged (in a certain parameter space).
This, in turn, will increase the overlap of the potentials from neighboring cells
and enhance the inter-supercell interaction.
This seems to be the situation for $U_{\mathrm{im}}=-20\ \text{eV}$ and
$U=9.3$~eV, for both choices of the supercell.
In this case, the impurity potential decreases slightly due to interactions
($U<U_\mathrm{im}$), leading to a small decrease of the gap at the Dirac point
compared to $U=0$ case [Fig.~\ref{fig:02}(b) and (d)].
At the same time, we find a significant difference between the average occupation numbers
of the nearest and next-nearest neighbors for $U=0$ and $U=9.3$~eV
(this can be also partly seen in the STM images in Fig.~\ref{fig:04}(a)-(b) and (d)-(e), for
states in a small energy window above the Fermi level, where
the neighbors of the impurity site appear brighter in the $U=9.3$~eV case compared to $U=0$).
As a result, the amplitude of impurity resonances in LDOS increases but their position
shifts to lower energies.
In contrast to this, for $U_{\mathrm{im}}=-20\ \text{eV}$ and $U=20$~eV,
the average occupation numbers of the nearest and next-nearest neighbors of the impurity site do not change
appreciably compared to $U=0$ case. The strength of the impurity potential for this value of $U$
is significantly reduced, leading to
a large decrease of the energy gap [Fig.~\ref{fig:02}(b) and (d)]. As a result,
the amplitude of impurity resonances increases further, however their position
remain close to the $U=0$ value.
These features strongly suggest that the position of the impurity resonances is sensitive to the
spatial extent of the impurity potential, namely the resonances move closer to zero energy
when the potential becomes more long-ranged.
To further clarify the
changes in the intensity and the spatial character of the low-energy impurity peaks,
brought about by interactions,
we present the simulated STM topographies in Fig.~\ref{fig:03} and Fig.~\ref{fig:04}
for $U_\mathrm{im}=-10$~eV and $U_\mathrm{im}=-20$~eV, respectively. For this
we plot LDOS for each atom in the supercell,~\footnote{The discrete values of LDOS
at lattice points are smeared out by adding a Gaussian broadening of
${\gamma}=N_{\text{atoms}}/E_{\text{mesh}}$, where $N_{\text{atoms}}$ is
the number of atoms in the supercell and $E_{\text{mesh}}$ is the energy mesh.
} integrated over the energy window $\Delta E$
above the Fermi level (we choose $\Delta E=0.25$~eV). This gives an estimate of the tunneling current
as electrons tunnel out of the occupied states of the STM tip into the unoccupied states of graphene.
The increase of the electronic density of states in
this energy window gives rise to a bright triangular feature around the impurity.
Note that for the $6\times 6$ supercell the impurity is located in sublattice A, while
for the $7\times 7$ supercell it is located in sublattice B. Therefore the bright triangular features
in the two supercells appear rotated by $180^{\circ}$ with respect to each other.
\begin{figure}
\centering
\includegraphics[width=0.97\linewidth]{fig3.jpg}
\caption{(Color online) Simulated STM topographies (LDOS for all atoms in the supercell)
for $6\times 6$ (left panels) and $7\times 7$ (right panels) supercell, for
a fixed impurity potential strength $U_{\mathrm{im}}=-10$~eV and varying
interaction strength: $U=0$ (a,d), $U=9.3$~eV (b,e) and $U=20$~eV (c,f). Arrows
mark the position of the impurity atom. Note the logarithmic color scale.
}
\label{fig:03}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=0.97\linewidth]{fig4.png}
\caption{(Color online) Same as Fig.~\ref{fig:03} but
with impurity potential strength $U_{\mathrm{im}}=-20$~eV.}
\label{fig:04}
\end{figure}
The difference between non-interacting and interacting cases is clearly visible in the STM images.
In all cases considered, the impurity site becomes progressively brighter
compared to its neighbors with increasing $U$.
This means that the electronic states
in the small energy window above the Fermi level become
more localized on the impurity site as a result of interactions.
This is most evident in the case of the $7\times 7$ supercell [Fig.~\ref{fig:03}(d)-(f)].
For $U_{\mathrm{im}}=-10$~eV, the overall
intensity of the images decreases with $U$.
For a stronger $U_{\mathrm{im}}=-20$~eV impurity potential, the trend is similar with the exception
of the intermediate interaction strength $U=9.3$~eV. In this case, the contribution of the nearest
and next-nearest neighbors is even stronger than in the $U=0$ case. This correlates
with the corresponding features in the LDOS discussed above.
%
\section{\label{sec:level5}Conclusions}
We have presented a theoretical study of the effects of
electron-electron interactions on the electronic states of graphene in the presence of
substitutional impurities.
Using a self-consistent TB model with on-site interactions treated at the mean-field level,
we have shown that the size of the gap, which opens up at the Dirac point in graphene upon doping,
and the character of the low-energy electronic states
are modified by interactions. The mechanism
for these effects is provided by the interplay between the impurity potential and the on-site repulsion,
which leads to significant re-arrangement of the electronic charge around the impurity
compared to the non-interacting case.
In particular, we found that the size of the gap
decreases with increasing the interaction strength.
Intuitively, this can be understood as follows. In the case of
an attractive impurity potential, which mimics nitrogen dopants in graphene,
adding the on-site Coulomb repulsion effectively reduces the strength of the potential,
i.e. the depth of the potential well decreases. This is due to the fact the on-site repulsion
prevents extra electronic charge from accumulating on the impurity site.
For a special supercell size $p\times p$, where $p$ is divisible by $3$, both
$K$ and $K^{\prime}$ are mapped onto $\Gamma$ point of the folded Brillouin zone.
Therefore, in the case
of undoped non-interacting graphene, there are four degenerate states at the neutrality point.
It is known that when the impurity potential is included, a gap (pseudogap) opens up between two of these states
while the other pair remains degenerate.~\cite{PhysRevB.86.045448}
We have shown that the size of this pseudogap
is reduced by interactions. Interestingly, in addition to this,
a small gap opens up between the second pair
of states at the $\Gamma$ point, which are otherwise degenerate in the absence of
interactions. We explain these features both qualitatively and quantitatively,
using a perturbative model based on the generalization of the approach developed by Lambin
\textit{et al.}~\cite{PhysRevB.86.045448} to the interacting case.
Furthermore, we have studied the behavior of the impurity-induced electronic states with and without interactions.
There are two groups of states which can be detected in the density of states when a carbon atom in
the supercell is replaced by an impurity atom. First, there are states which emerge far away from
the Fermi energy, with their energies of the same order of magnitude as the impurity potential (
$\approx 10$~eV in our calculations). Second, there are states appearing close to the Fermi energy
and are therefore of particular interest. Although the
way the electron-electron interactions affect the LDOS at low energies
in general depends on the impurity concentration, we found
clear trends in the behavior of the impurity-resonances as both parameters,
i.e. the interaction strength and the impurity potential strength,
are modified.
Regardless of the interactions, the impurity levels
move closer to the low-energy region (e.g. to the original Dirac point)
with increasing the impurity potential strength.
This finding is consistent with previous
calculations for graphene.~\cite{PhysRevB.77.115109,PhysRevB.73.241402,PhysRevB.75.125425}
Similar result was found in another
class of Dirac materials, namely in three-dimensional topological insulators
in the presence of strong potential impurities on the surface.~\cite{PhysRevB.81.233405,PhysRevB.85.121103}
However, our self-consistent calculations for graphene in the presence of both impurities and interactions
reveal novel features.
For a fixed impurity potential,
sort-range interactions
tend to enhance the amplitude of the impurity-resonances in the vicinity
of the Fermi level. The position of the resonances is also affected by
the spatial extent of the effective impurity potential, which is modified by interactions.
As the interaction strength increases, the states become more localized on the impurity atom.
The differences in the spatial distribution of the low-energy impurity states
in the non-interacting and interacting cases are clearly detectable in the simulated STM topographies.
|
\section{Introduction}\label{sec:intro}
In this paper we will address error estimation for shape optimization problems and study an adaptive finite element method based on
an optimal microstructured elastic material composed of sequential laminates. A classical shape optimization problem amounts to
distribute a fixed amount of rigid material throughout a domain in order to optimize a particular cost functional
when confronted with surface and volume loads. However it is known that the problem is ill-posed if only scale invariant cost functionals
are taken into account. As the formation of microstructures comes along with this deficiency most approaches
enforce regularity by an additional regularizing cost functional such as the shape perimeter. Alternatively the problem
can be relaxed to allow for more general shapes characterized by local effective material properties and densities. They stem from
the shape structure on the microscale influencing the effective macroscopic description via homogenization. Optimal microstructures exist but are not unique.
A particular flexible microstructure is constructed via sequential lamination. This allows for a convenient algorithmic treatment as a small set of parameters
characterizing the local microstructure can be computed explicitly from the local stress.\\
Numerical computations suggest that optimal shapes are characterized by regions with substantially different mircostructure and sharp transitions between these regions. Thus an adaptive mesh strategy based on an a posteriori error control is evidently the appropriate strategy.
To this end, the error in the achieved cost has to be controlled.
{\it Review of related work.}
Shape optimization is a classical field in PDE constrained optimization and has been covered extensively in the literature, see e.\,g.\
the textbooks \cite{Be95a,Al02}. For the link to homogenization theory we refer to \cite{Jikov1994,BrDe98,DoCi99,BuDa91}.
Optimal microstructures where
first derived by Hashin in 1962 via the concentric sphere construction for hydrostatic loads \cite{Ha62}. The sequential lamination construction
dates back to the 1980s \cite{Ta85,MuTa85,FrMu86,LuCh86,GiCh87,Av87} and was later used in a practical numerical scheme for topology optimization
in~\cite{AlBoFr97}. In all these cases proofs of optimality rely on the Hashin-Shtrikman bounds on the attainable sets of effective elastic properties \cite{HaSh63}.
Related to the homogenization approach is the so called free material optimization, see e.\,g.~\cite{HaKoLe10}. Here the coefficients of the elasticity
are the degrees of freedom to be optimized. Possibly yielding materials without a physical equivalent, best approximating realizations are searched
in a post processing step, sometimes referred to as inverse homogenization.
Applying the classical homogenization paradigm one asks for the optimal (geometric) microstructures and the effective material properties
such that a macroscopic cost functional is optimized.
In~\cite{BaTo10,BaTo10a} a displacement approach on the microscale is investigated which consists of affine plus periodic functions on
generalized periodic domains. For the numerical optimization of the microscopic geometry a boundary tracking method was used.
The incorporation of topological derivatives allowed for the nucleation of small holes and additional remeshing steps ensured mesh quality.
For specific geometric microstructures the effective stress response of
basic affine strains on the microscopic cell boundary are investigated in \cite{BaTo12} and used to describe the effective macroscopic behavior.
The geometry of the locally periodic microstructures are then optimized.
The same methodology was picked up in \cite{CoGeRu14} while investigating microstructured materials with geometrically simple
perforations. The resulting shapes offered compliance values close to optimal ones generated by twofold sequential lamination. Moreover they
indicate the need to resolve sharp interfaces between regions with different types of local microstructure.
To be able to both capture microscopic effects and macroscopic material behavior while maintaining feasible computational complexity in
PDE simulation tailored numerical methods have to be proposed.
The heterogeneous multi-scale method (HMM) presented in \cite{EEnHu03,EEn03,EEn05,EMiZh05} depicts a very general
paradigm using independent macroscopic and microscopic schemes.
In \cite{Oh05,HeOh09} a posteriori error estimates for elliptic homogenization problems involving a fine scale diffusion term were derived.
Here the heterogeneous multi-scale method is reformulated as a direct finite element discretization of the two-scale homogenized
equation in variational form.
The obtained error indicator provides separate terms for errors in the macroscopic and microscopic discretization.
Concerning the a posteriori control it is often not desirable to measure errors in classical function space norms but the error for a prescribed cost functionals.
From a practical point of view quantities of interest for a real composite work piece like e.\,g.\ stresses in critical sections,
averaged surface tensions or (mollified) pointwise stresses can be considered as cost functionals and corresponding a
posteriori error estimates are derived \cite{PrOd99,OdVe00,OdVe00a,Ve04}. \\
The approach we pick up in this contribution is the dual weighted residual (DWR) method introduced in \cite{BeRa97} and applied to optimal control problems
for instance in \cite{BeKaRa00}. Here residual type error estimates get weighted using duality methods.
In \cite{BeEsTr11} a special marking strategy was employed allowing to show quasi-optimality of the associated adaptive finite element
method while maintaining sound convergence.
Control and state constraints were taken into account in \cite{BeVe09,VeWo08,LeMeVe13}.
Matrix valued $L^1$ optimal control in coefficients is studied in \cite{KoLe13}.
Shape optimization was addressed in \cite{KiVe13} using a tracking type cost functional, a Helmholtz state equation, and a graph representation of
the boundary. For this control function higher regularity could be shown and a priori estimates were derived.
An adaptive finite element method for shape optimization via boundary tracking, remeshing and perimeter penalization was presented in
\cite{MoNoPa10,MoNoPa12}. Here a DWR type of approach is used to assess the PDE error and the actual geometric error is estimated differently.
Furthermore the DWR method was used in \cite{KaDeGa14, Ka13} in the context of one shot methods for fuel ignition problems, the viscous Burgers
equation and aerodynamic shape optimization.
Optimal design in Navier Stokes flow is considered in \cite{BrLiUl12}. Here the DWR method for optimal control problems leads to separate terms for the
primal, the adjoint and the control residual.
Similarly \cite{Wo10} addresses a problem in free material optimization based on the variable thickness sheet model.
The paper is organized as follows. In section~\ref{sec:basics} we will describe in condensed form the necessary background on linearized elasticity, shape optimization, the sequential laminates construction in shape optimization, and its numerical discretization.
A posteriori error estimates based on the dual weighted residual approach are derived in section~\ref{sec:dwr}.
In section~\ref{sec:apriori} we discuss the estimation of weighting term using a priori regularity
while section~\ref{sec:localapprox} focuses on their numerical approximation. Details concerning the implementation
are given in section~\ref{sec:impl} and numerical results are finally presented in section~\ref{sec:results}.
\section{Elastic shape optimization and microscopic lamination}\label{sec:basics}
We consider an elastic object given as a simple connected domain $\Omega \subseteq D \subset \mathds{R}^2$
where $D$ is a circumjacent working domain and we at first suppose that $D$ is filled with elastic material (hard and soft)
and we aim at the optimization of the distribution of the two phases.
A Dirichlet boundary part $\Gamma_D \subset \partial D \cap \partial \Omega$ and a Neumann boundary
$\Gamma_N \subset \partial D \cap \partial \Omega$ subject to a sufficiently regular surface load $g$ are both supposed to be fixed and not subject to optimization.
The remaining boundary $\partial \Omega \setminus (\Gamma_D \cup \Gamma_N)$ can be varied.
The applied forcing causes stresses inside the material which maintain an inner equilibrium. It can be described
by the system of partial differential equations of linearized elasticity. Let $\Sigma$ denote the set of
feasible stresses in accordance to the PDE, i.e.
\begin{equation}\label{eq:feasible}
\Sigma\! :=\! \left\lbrace \sigma\!:\! D \!\rightarrow\! \mathds{R}^{2\times 2}_{\text{sym}} \; \big| \;\!
\div \{ \sigma \} \!= \!0 \text{ in }D, \, \sigma n \!=\! g \text{ on }\Gamma_{\!N}, \, u \!=\! 0 \text{ on }\Gamma_{\!D},
\, \sigma \!= \!C[q] \eps{u} \right\rbrace
\end{equation}
Here $u: D \rightarrow \mathds{R}^2$ denotes the displacement, $n$ the outward pointing normal on $\Gamma_N$,
$\eps{u} = \frac12 \left( \ensuremath{\mathrm{D}}u + \ensuremath{\mathrm{D}}u^\top \right)$ the symmetrized strain tensor, and
$C[q]: D \rightarrow \mathds{R}^{2^4}$ the elasticity tensor. The elasticity tensor is a forth order tensor with the symmetry properties
$(C[q])_{ijkl} = (C[q])_{jikl} = (C[q])_{ijlk} = (C[q])_{klij}$ and the ellipticity condition
$C[q] \, \xi : \xi \geq c \, \xi \cdot \xi$. Usually it can be characterized locally by a small parameter vector $q$.
We consider here material which behaves isotropic and is described by Lam\'e parameters $\lambda$ and $\mu$ and the strain stress relation
$\sigma = 2 \mu \eps{u} + \lambda \tr\eps{u} \mathds{1}$.\\
For later purposes we take into account the weak displacement formulation of \eqref{eq:feasible}, i.\,e.\
$u \in H^1_{\Gamma_D}$, the space of $L^2$ vector valued functions with weak derivatives in $L^2(\Omega)$
and vanishing trace on $\Gamma_D$, solves the variational problem
\begin{equation}\label{eq:weak}
a(q;u,\varphi) = l(\varphi) \;\;\forall\,\varphi \in H^1_{\Gamma_D}
\end{equation}
with the quadratic form $a(q;u,\varphi) = \int_D C[q] \, \eps{u} : \eps{\varphi} \;\mathsf{d\mathbf{x}}$ and the linear form
$l(\varphi) = \int_{\Gamma_N}\, C[q] \, \eps{u} n \cdot \varphi \;\mathrm{da(\mathbf{x})} = \int_{\Gamma_N}\, g \cdot \varphi \;\mathrm{da(\mathbf{x})} \,$.
A classical shape optimization problem now amounts to the task of distributing a hard material whose elastic behavior is
described by an elasticity tensor $A$. To this end, we consider a characteristic function $\chi \in L^\infty(D,\lbrace 0,1 \rbrace)$ of the hard phase
with $\int_D \chi \;\mathsf{d\mathbf{x}} = \Theta$ reflecting the volume constraint. The elasticity tensor on $D$ is then
defined as $C[q] = \chi A[q] + (1-\chi) B[q]$ where $B[q]$ is an elasticity tensor for a soft material filling
the remaining space.
To assess the performance of a given shape for a prescribed loading a suitable cost functional has to be taken into account.
We restrict ourselves to the compliance cost as a global measure of rigidity. For a displacement $u= u[\chi]$ being the unique solution
of \eqref{eq:weak} for given $\chi$ it is given by
\begin{equation} \label{eq:compliance}
J[u[\chi] ,\chi ] := \int_{\Gamma_N}\, g \cdot u[\chi] \;\mathrm{da(\mathbf{x})} = l(u[\chi] ) = a(q;u[\chi] ,u[\chi] ) \,.
\end{equation}
The cost functional $ J[u[\chi] ,\chi ]$ should thereby be minimized with respect to $\chi$ subject to a constraint on the volume of the hard material phase.
Ultimately one is interested in the degenerate case $B=0$, where there is void outside of
$\Omega=\{x\in D \;|\; \chi(x) = 1\}$.
This minimizing problem is ill-posed and on a minimizing sequence the onset of microstructures can be observed.
Thus one investigates a relaxed formulation of this shape optimization problem. We refer to
\cite{AlBoFr97,Al02} for a concise summary and references therein for further details. Relaxation leads to the following reformulation:
\begin{equation} \label{eq:relaxed}
\min\limits_{\sigma \in \Sigma} \, \int_D \min\limits_{0 \leq \theta \leq 1} \,
\min\limits_{C^\ast[q] \in G_\theta^B} {C^\ast[q]}^{-1} \sigma : \sigma + l \, \theta \;\mathsf{d\mathbf{x}}.
\end{equation}
Reading from left to right, for fixed stresses
$\sigma$ on $D$ solving \eqref{eq:feasible}, a fixed point $x \in D$ and a fixed local density $\theta(x)$, an
elasticity tensor $C^\ast[q]$ minimizing the elastic energy density is to be found
among all effective homogenized elasticity tensors resulting from a mixture of materials $A$ and $B$.
Additionally the Lagrangian multiplier $l$ takes into account the amount of total material used. It can be
shown that it is a decreasing function of $l$ which facilitates an adaptation to enforce a fixed prescribed
volume at any time.
For the set $G_\theta^B$ of homogenized tensors, also referred to as G-closure, no algebraic characterization is known. However there exist
sharp bounds on the elastic energy density, called Hashin-Shtrikman bounds. Theses bounds are attained within the subclass
of sequentially laminated materials. The construction starts on a scale $\epsilon_1$ by layering rigid material $A$ and
soft material $B$ with a certain ratio and along a certain direction. By means of homogenization effective material properties
can be computed and used in the next iteration in place of material $B$. Mechanically this takes place on a substantially larger scale $\epsilon_2 \gg \epsilon_1$.
In two dimensions and for the compliance cost it is known that the construction on these two iterations is sufficient, cf. \cite{KoLi88,AlKo93a}.
It is moreover possible to pass from the weak material $B$ to void \cite{AlKo93,AlBoFr97}. \\
The sequential laminates microstructure is locally characterized by three parameters (cf.~Fig.~\ref{fig:laminatesSketch}):
the inclination of the main lamination direction $\alpha(x)$, the ratio of material spent in the second lamination stage $m(x)$,
and the overall local material density $\theta(x)$. The second lamination direction is always orthogonal to the first one and the second
ratio parameter is $1-m(x)$.
\begin{figure}[!ht]
\hfill
\begin{minipage}{.22\linewidth}
\setlength{\unitlength}{\linewidth}
\includegraphics[width=\linewidth]{laminates_marked.pdf}
\begin{picture}(0,0)
\put(.08,.68){$\alpha$}
\put(-.05,.19){$m$}
\put(.71,.13){$\epsilon_1$}
\put(.22,.04){$\epsilon_2$}
\end{picture}
\end{minipage}
\hfill
\begin{minipage}{.25\linewidth}
\includegraphics[width=\linewidth]{ldomain_pull_laminates_1e7_lvl4_doerflerD30_grid_iter024.png}
\end{minipage}
\hfill
\begin{minipage}{.25\linewidth}
\includegraphics[width=\linewidth]{ldomain_pull_laminates_1e7_lvl4_doerflerD30_density_iter024.png}
\end{minipage}
\hfill
\begin{minipage}{.25\linewidth}
\includegraphics[width=\linewidth]{ldomain_pull_laminates_1e7_lvl4_doerflerD30_mises_iter024.png}
\end{minipage}
\hfill\mbox{}
\caption[]{Sketch of twofold sequential lamination.
Numerically optimized result for an L-shaped domain fixed at the bottom with downwards pointing load at the center region on the right hand side.
An adaptively refined grid is shown along with the density lamination parameter $\theta$ plotted and the von Mises stress using the color coding \raisebox{1pt}{\includegraphics[height=4pt]{colorbarRatio}}.}
\label{fig:laminatesSketch}
\end{figure}
All three parameters directly depend on the eigenvalues $\lambda_1$, $\lambda_2$
and the eigenvectors $\tau^{(1)}$, $\tau^{(2)}$ of the local stress tensor $\sigma$ and can be computed explicitly via the mapping
$x \mapsto q[\sigma](x) = \left( \alpha[\sigma](x), m[\sigma](x),\theta[\sigma](x) \right)^\top$, see \cite{AlBoFr97}:
\begin{equation}\label{eq:explicitparameters}
\begin{aligned}
\alpha[\sigma] &= \arctan\left( \frac{\tau^{(1)}_y(\sigma)}{\tau^{(1)}_x(\sigma)} \right) \,, \qquad
m[\sigma] = \frac{|\lambda_2(\sigma)|}{|\lambda_1(\sigma)|+|\lambda_2(\sigma)|} \,, \\
\theta[\sigma] &= \min\left\{ 1, \sqrt{ \frac{2 \mu + \lambda}{4 \mu (\mu+\lambda) \, l} } ( |\lambda_1(\sigma)|+|\lambda_2(\sigma)| ) \right\} \,.
\end{aligned}
\end{equation}
Here $\lambda$ and $\mu$ denote the Lam\'e parameters of the rigid isotropic material $A$ and $\tau^{(i)}_x$, $\tau^{(i)}_y$ the first and
second component of the two dimensional vector $\tau^{(i)}$.
Given these parameters the effective homogenized elasticity tensor $C^\ast[q](x)$ can likewise be computed explicitly:
\begin{equation}\label{eq:effectivetensor}
\begin{aligned}
C^\ast_{mnop}[q] &= R[\alpha] \, \bar{C}[m,\theta] := Q_{mi}[\alpha] \, Q_{nj}[\alpha] \, Q_{ok}[\alpha] \, Q_{pl}[\alpha] \; \bar{C}_{ijkl}(m,\theta) \,, \\[6pt]
\bar{C}_{1111}[m,\theta] &= \frac{ 4 \kappa \mu (\kappa+\mu) \theta (1-\theta (1-m))(1-m) } {4 \kappa \mu m (1-m) \theta^2 + (\kappa+\mu)^2 (1-\theta)} \,, \\
\bar{C}_{2222}[m,\theta] &= \frac{ 4 \kappa \mu (\kappa+\mu) \theta (1-\theta m)m } {4 \kappa \mu m (1-m) \theta^2 + (\kappa+\mu)^2 (1-\theta)} \,, \\
\bar{C}_{1122}[m,\theta] &= \frac{ 4 \kappa \mu \lambda \theta^2 m (1-m) } {4 \kappa \mu m (1-m) \theta^2 + (\kappa+\mu)^2 (1-\theta)} \,.
\end{aligned}
\end{equation}
Thereby $R$ denotes the linear mapping which transforms the tensor $\bar{C}$ given in reference configuration into the appropriate coordinate frame
spanned by the eigenvectors of the stress tensor. In its definition $Q$ are $2 \times 2$ rotation matrices and the Einstein
summation convention is used. The bulk modulus is defined as $\kappa = \lambda + \mu$. The tensor $\bar{C}$ is complemented using the symmetry relations and filling the remaining entries with $0$. This yields a singular elasticity tensor in the sense that a strain with
just off diagonal entries, corresponding to shearing deformations, will lead to a vanishing stress. \\
In the following we will assume that the density $\theta$ is always bounded away from $0$ by a small threshold.
Thus we deal with the hard\/soft configuration. For the ease of notation
we extend the Neumann boundary to $\Gamma_N = \partial D \setminus \Gamma_D$ and prescribe
homogeneous boundary conditions $g = 0$ on the added parts.
For the discretization we use a finite element ansatz and assume that we are given a
triangulation $\mathcal{T}$ of $D$ with elements $T \in \mathcal{T}$ and denote by $h$ the piecewise constant mesh size function.
Let $\mathcal{V}^{(2)}_h$ be the space of piecewise bi-quadratic and continuous vector valued functions on which we ask for a solution $u_h$ of the discrete
weak problem $a(q_h;u_h,\varphi_h) = l(\varphi_h)$ for all $\varphi_h \in \mathcal{V}^{(2)}_h$, or more explicitly
\begin{equation}\label{eq:discrete}
\begin{aligned}
& \int_D C^\ast[q_h] \, \eps{u_h} : \eps{\varphi_h} \;\mathsf{d\mathbf{x}}
= \int_{\Gamma_N}\, g \cdot \varphi_h \;\mathrm{da(\mathbf{x})} \;\;\forall\,\varphi_h \in \mathcal{V}^{(2)}_h \,.
\end{aligned}
\end{equation}
Associated to $u_h$ are stresses $\sigma_h := \sigma[u_h](x) = C^\ast[q_h] \, \eps{u_h}$
for which lamination parameters are computed via formula \eqref{eq:explicitparameters}.
For brevity of notation we write $q_h = (\alpha_h,m_h,\theta_h) := q[\sigma_h]$.
Let us emphasize that the parameter function varies from point to point.
More details on the numerical realization will be given in section~\ref{sec:impl}.
\section{A posteriori error estimates based on the dual weighted residual approach} \label{sec:dwr}
In this section we will derive estimates of the error in the compliance objective \eqref{eq:compliance} when using optimal
sequential laminated microstructures in shape optimization. Let $u$ be the solution of the continuous problem \eqref{eq:weak}
involving the elasticity tensor $C^\ast[q]$ with optimal lamination parameters $q$ defined by \eqref{eq:explicitparameters}
and $u_h$ be a discrete solution satisfying \eqref{eq:discrete}.
We are interested in the a posteriori control of the numerical approximation error in the cost functional.
To this end we pick up the dual weighted residual approach \cite{BeKaRa00}. In that respect we will treat
the rotation parameter $\alpha$ differently from the other two parameters $m$ and $\theta$, which will from now on be denoted by
$\vartheta := (m,\theta)$.
Let us begin with a useful observation for the elastic energy density, namely that
it is invariant with respect to the rotation of the microstructure. Intuitively this follows from
the fact that the lamination construction yields an orthotropic material which is always aligned with the main stress directions.
\begin{lemma}[Invariance of the elastic energy density w.\,r.\,t.\ rotations] \label{lemma:invariance}
Let $x \in D$ be fixed and $C^\ast[\alpha,\vartheta]$ be an effective elasticity tensor corresponding to an optimal
rank 2 sequential laminate at $x$. Let $u[\alpha,\vartheta]$ be the solution of \eqref{eq:weak}.
Then the local elastic energy density is invariant w.\,r.\,t.\ the rotation parameter $\alpha$, i.\,e.
$$\tdiff{\alpha} \left( C^\ast[\alpha,\vartheta] \, \eps{u[\alpha,\vartheta]} : \eps{u[\alpha,\vartheta]} \right) (x) = 0\,.$$
\end{lemma}
\textbf{Proof.}
With $\sigma[\alpha,\vartheta] := C^\ast[\alpha,\vartheta] \, \eps{u}$ the elastic energy density can be rewritten as
\[
C^\ast[\alpha,\vartheta] \, \eps{u[\alpha,\vartheta]} : \eps{u[\alpha,\vartheta]} = {C^\ast}^{-1}[\alpha,\vartheta] \, \sigma[\alpha,\vartheta] : \sigma[\alpha,\vartheta] \,.
\]
Here $\alpha$ and $\vartheta$ enter the effective homogenized tensor $C^\ast[\alpha,\vartheta]$ by means of \eqref{eq:effectivetensor}.
Since $C^\ast[\alpha,\vartheta]$ represents an optimal material the elastic energy density attains the lower Hashin-Shtrikman bound,
cf.~\cite[Theorem~2.3.35]{Al02}, and we obtain
\[
{C^\ast}^{-1}[\alpha,\vartheta] \, \sigma[\alpha,\vartheta] : \sigma[\alpha,\vartheta]
= A^{-1} \sigma[\alpha,\vartheta] : \sigma[\alpha,\vartheta]+ \frac{(\kappa+\mu)\theta}{4\kappa \mu (1-\theta)} ( |\lambda_1| + |\lambda_2| )^2 \,.
\]
Here the eigenvalues $\lambda_1$, $\lambda_2$ of $\sigma$ do not depend on the rotation.
In fact, the rotation parameter $\alpha$ is derived from the orientation of the eigenvectors.
Thus $\alpha$ only appears via $\sigma$ in the first term of the right hand side.
This term also represents an elastic energy density but this time involving the isotropic constituent $A$.
Because of the isotropy of $A$ altering the alignment of $\sigma$, e.\,g.\ by rotating to reference configuration, leaves this term unchanged.
Altogether this means that the right hand side does not depend on the rotation parameter. This proofs the claim.
\qed \\
We now present the main result.
\begin{theorem}[Weighted a posteriori error estimate] \label{theorem:estimates}
For a given continuous solution $u$ to \eqref{eq:weak} with a microstructured material given by the optimal lamination parameter function $q$ and a numerical
approximation $u_h$ the following error estimate holds for the compliance objective \eqref{eq:compliance}:
\[
|J[q;u]-J[q_h;u_h]| \leq \sum_T \eta_T(u_h,q_h) + \mathcal{R} \,.
\]
Here $\mathcal{R}$ is a higher order remainder term involving higher order derivatives and the local error indicators $\eta_T$ of first order are defined by
\[
\eta_T(u_h,q_h) := \rho_T^{(u)} \omega_T^{(u)} + \rho_{\partial T}^{(u)} \omega_{\partial T}^{(u)}
+ \fr12 \rho_T^{(m)} \omega_T^{(m)} + \fr12 \rho_T^{(\theta)} \omega_T^{(\theta)}
\quad \text{with}
\]
\begin{equation*}
\begin{aligned}
\rho_T^{(u)} &= \left\Vert -\div \left\{ \sigma_h \right\} \right\Vert_{0,2,T} \,, &
\omega_T^{(u)} &= \left\Vert u-\tilde{u}_h \right\Vert_{0,2,T} \,, \\
\rho_{\partial T}^{(u)} &=
\left\lbrace
\begin{array}{l l}
\Vert \textstyle\frac{1}{2} \left[ \sigma_h n \right]\Vert_{0,2,\partial T}; & \partial T \cap \partialD = \emptyset \\
\Vert \sigma_h n - g\Vert_{0,2,\partial T}; & \partial T \subset \Gamma_N
\end{array}
\right. \,, &
\omega_{\partial T}^{(u)} &= \left\Vert u-\tilde{u}_h \right\Vert_{0,2,\partial T} \,, \\
\rho_T^{(m)} &= \left\Vert R[\alpha_h] \, \bar{C}_{,m}\left[m_h,\theta_h\right] \; \eps{u_h[\alpha_h]} : \eps{u_h[\alpha_h]}\right\Vert_{0,1,T} \,, &
\omega_T^{(m)} &= \left\Vert m-\tilde{m}_h \right\Vert_{0,\infty,T} \,, \\
\rho_T^{(\theta)} &= \left\Vert R[\alpha_h] \, \bar{C}_{,\theta}\left[m_h,\theta_h\right] \; \eps{u_h[\alpha_h]} : \eps{u_h[\alpha_h]}\right\Vert_{0,1,T} \,, &
\omega_T^{(\theta)} &= \left\Vert \theta-\tilde{\theta}_h \right\Vert_{0,\infty,T} \,,
\end{aligned}
\end{equation*}
where $\tilde{u}_h \in \mathcal{V}^{(2)}_h$ and $\tilde{m}_h$, $\tilde{\theta}_h$ are chosen arbitrarily.
\end{theorem}
\textbf{Proof.}
The first step is to use the Lagrangian $\mathcal{L}(q;u,p) := J[q;u] + a(q;u,p) - l(p)$ for decoupled $q$, $u$, and $p$ and take into account
the PDE constraint, namely that the displacement $u$ has to be a solution of the linearized elasticity system \eqref{eq:weak}.
The first order optimality conditions are
\begin{align}
\rho(q;u)(v) := \mathcal{L}_{,p}(q;u,p)(v) &= a(q;u,v) - l(v) \,, \label{primal}\\
\rho^\ast(q;u,p)(v) := \mathcal{L}_{,u}(q;u,p)(v) &= J_{,u}[q;u](v) + a(q;v,p) \,, \label{dual} \\
\rho^q(q;u,p)(v) := \mathcal{L}_{,q}(q;u,p)(v) &= J_{,q}[q;u](v) + a_{,q}(q;u,p)(v) - l_{,q}(p)(v) \,, \label{control}
\end{align}
where $v$ is a universal variable for an arbitrary test function from the corresponding space.
Equation \eqref{primal} is just the definition of the elastic solution $u$.
Equation \eqref{dual} defines the dual solution, which in case of compliance is $p=-u$.
As we assumed $u$ and $u_h$ to be solutions of the continuous and the discrete problem, respectively,
$(q,u,-u)$ and $(q_h,u_h,-u_h)$ are stationary points of the Lagrangian. Thus we have
\[
\mathcal{L}(q;u,p)-\mathcal{L}(q_h;u_h,p_h) = J[q;u]-J[q_h;u_h] \,.
\]
Estimating the error can therefore be done by considering the difference in the Lagrangian.\\
Next we consider a suitable first order expansion of the Lagrangian.
Here we will take special care of the rotation parameter $\alpha$ to be able
to use Lemma~\ref{lemma:invariance} above.
The Lagrangian thus only depends on $u$ and the controls $q=(\alpha,\vartheta)$ because the dual solution coincides with the negative primal solution.
Hence, in the following we will skip the dual solution as a parameter.
We now study the dependence of the elastic solution $u[\alpha]$ on the rotation parameter $\alpha$.\\
As a shortcut notation we write $e_u(s) = u[\alpha]-u_h[\alpha_h + s e_\alpha]$, $e_\alpha = \alpha -\alpha_h$, and $e_{\vartheta} = \vartheta-\vartheta_h$
for the difference between the exact and the discrete displacement and controls. Here $u_h$ is given for the continuous spatially varying parameter function
$\alpha_h + s e_\alpha$ as the solution to the corresponding problem \eqref{eq:discrete}.
We now express the error in the Lagrangian as a 1D integral and observing $\left. u_h[\alpha_h+s e_\alpha] + s e_u(s) \right|_{s=1} = u[\alpha]$
and $\left. u_h[\alpha_h+s e_\alpha] + s e_u(s) \right|_{s=0}$ $= u_h[\alpha_h]$ we obtain
\begin{eqnarray*}
e_\mathcal{L} &:=& \mathcal{L}(\alpha,\vartheta;u[\alpha])-\mathcal{L}(\alpha_h,\vartheta_h;u_h[\alpha_h])\\
&=& \int_0^1 \tdiff{s} \mathcal{L} (\alpha_h + s e_\alpha, \vartheta_h + s e_{\vartheta}, u_h[\alpha_h+s e_\alpha] + s e_u(s) ) \,\mathrm{d}s\,.
\end{eqnarray*}
Now, we use the trapezoidal rule $\int_0^1 f(s) \mathrm{d}s = \frac12 ( f(0) + f(1) ) + \frac12 \int_0^1 f''(s) s(1-s)\, \mathrm{d}s$ with
$f(s) = \tdiff{s} \mathcal{L}(\alpha_h + s e_\alpha, \vartheta_h + s e_{\vartheta}, u_h[\alpha_h+s e_\alpha] + s e_u(s) )$.
The term $f(1)$ vanishes because $(\alpha,\vartheta,u[\alpha])$ is assumed to be a stationary point and
$e=(e_\alpha,e_{\vartheta},u[\alpha] - u_h[\alpha])^T$ is a feasible test direction, i.e. $\nabla \mathcal{L}(\alpha,\vartheta,u[\alpha])(e) = 0$.
Thus, we obtain
$$
e_\mathcal{L} = \frac12 \nabla \mathcal{L} (\alpha_h, \vartheta_h, u_h[\alpha_h])
\left( \begin{array}{c} e_\alpha \\ e_{\vartheta} \\ {u_h}_{,\alpha}[\alpha_h](e_\alpha) + e_u(0) \end{array} \right) + \mathcal{R}\,
$$
where $\mathcal{R} := \frac12 \int_0^1 \frac{\mathsf{d}^3}{\mathsf{d} s^3} \mathcal{L}
( \alpha_h + s e_\alpha, \vartheta_h + s e_{\vartheta}, u_h[\alpha_h+s e_\alpha] + s e_u(s)) \, s \, (1-s) \mathrm{d}s\,$.
Next, we expand the lower order term and achieve
\begin{eqnarray*}
&& \frac12 \nabla \mathcal{L} (\alpha_h, \vartheta_h, u_h[\alpha_h])
\left( \begin{array}{c} e_\alpha \\ e_{\vartheta} \\ {u_h}_{,\alpha}[\alpha_h](e_\alpha) + e_u(0) \end{array} \right) \notag \\
&& = \frac12 \mathcal{L}_{,u} (\alpha_h,\vartheta_h;u_h[\alpha_h]) ( u[\alpha] - u_h[\alpha_h] )
+ \frac12 \mathcal{L}_{,\vartheta} (\alpha_h,\vartheta_h;u_h[\alpha_h]) ( \vartheta-\vartheta_h ) \notag \\
&& \quad + \frac12 \mathcal{L}_{,\alpha} (\alpha_h,\vartheta_h;u_h[\alpha_h]) ( \alpha -\alpha_h )
+ \frac12 \mathcal{L}_{,u} (\alpha_h,\vartheta_h;u_h[\alpha_h]) ( {u_h}_{,\alpha}[\alpha_h](\alpha -\alpha_h)) \notag \\
&& = \frac12 \mathcal{L}_{,u} (\alpha_h,\vartheta_h;u_h[\alpha_h]) ( u[\alpha] - u_h[\alpha_h] )
+ \frac12 \mathcal{L}_{,\vartheta} (\alpha_h,\vartheta_h;u_h[\alpha_h]) ( \vartheta-\vartheta_h ) \notag \\
&& \quad + \frac12 \tdiff{\alpha} \mathcal{L} (\alpha,\vartheta_h;u_h[\alpha])\big|_{\alpha=\alpha_h} ( \alpha -\alpha_h ) \,. \notag
\end{eqnarray*}
By definition we have
\[
\tdiff{\alpha} \mathcal{L} (\alpha,\vartheta_h;u_h[\alpha])\big|_{\alpha=\alpha_h}
= \left[\tdiff{\alpha} \int_D C^\ast\left[\alpha,\vartheta_h \right] \eps{u_h[\alpha] } : \eps{u_h[\alpha]} \;\mathsf{d\mathbf{x}}\right]_{\alpha=\alpha_h}
\]
and thus the term vanishes by virtue of Lemma~\ref{lemma:invariance}.
Finally we consider arbitrary discrete test functions $\tilde{u}_h \in \mathcal{V}^{(2)}_h$ and $\tilde{\vartheta}_h$
resulting from an arbitrary stress by means of formula \eqref{eq:explicitparameters}
and use the fact that $(\alpha_h,\vartheta_h,u_h[\alpha_h])$ is a discrete stationary point of the Lagrangian and thus $\nabla \mathcal{L} (\alpha_h, \vartheta_h, u_h[\alpha_h])$ vanishes in directions $(0, \tilde{\vartheta}_h -\vartheta_h, \tilde{u}_h[\alpha_h] - u_h[\alpha_h] )$. This gives
\begin{eqnarray*}
e_\mathcal{L} \!\!
&=& \!\! \frac12 \mathcal{L}_{,u} (\alpha_h,\vartheta_h;u_h[\alpha_h]) ( u[\alpha] - \tilde{u}_h[\alpha_h] )
+ \frac12 \mathcal{L}_{,\vartheta} (\alpha_h,\vartheta_h;u_h[\alpha_h]) ( \vartheta - \tilde{\vartheta}_h ) + \mathcal{R} \,.\notag
\end{eqnarray*}
Furthermore, we consider the weak formulation \eqref{eq:weak} and the compliance cost functional
\eqref{eq:compliance}. For the \emph{primal residual} we get:
\begin{eqnarray*}
&& \mathcal{L}_{,u} (\alpha_h,\vartheta_h;u_h[\alpha_h]) ( u[\alpha] - \tilde{u}_h[\alpha_h] ) \\
&&= -2 \int_D C^\ast\left[\alpha_h,\vartheta_h \right] \eps{u_h[\alpha_h] } : \eps{ u[\alpha] - \tilde{u}_h[\alpha_h]} \;\mathsf{d\mathbf{x}} + 2 \int_{\Gamma_N}\, g \cdot (u[\alpha] - \tilde{u}_h[\alpha_h]) \\
&&= 2 \, \Big( \sum_j \int_{T_j} \div \left\{ C^\ast\left[\alpha_h,\vartheta_h \right] \eps{u_h[\alpha_h]} \right\} \cdot (u[\alpha] - \tilde{u}_h[\alpha_h]) \;\mathsf{d\mathbf{x}} \\
&& \qquad \qquad - \int_{\partial T_j} C^\ast\left[\alpha_h,\vartheta_h \right] \eps{u_h[\alpha_h]} n \cdot (u[\alpha] - \tilde{u}_h[\alpha_h]) \;\mathrm{da(\mathbf{x})} \\
&& \qquad \qquad + \int_{\partial T_j \cap \Gamma_N} g \cdot (u[\alpha] - \tilde{u}_h[\alpha_h]) \;\mathrm{da(\mathbf{x})} \Big)
\end{eqnarray*}
This leads to the postulated residual terms with
\[
\begin{aligned}
\rho_T^{(u)} &= \left\Vert \div \left\{ \sigma[u_h[\alpha_h]] \right\} \right\Vert_{0,2,T}, & \!\!
\omega_T^{(u)} &= \left\Vert u[\alpha] - \tilde{u}_h[\alpha_h] \right\Vert_{0,2,T}, \\
\rho_{\partial T}^{(u)} &=
\left\lbrace
\begin{array}{l l}
\Vert \textstyle\frac{1}{2} \left[ \sigma[u_h[\alpha_h]] \cdot n \right]\Vert_{0,2,\partial T}\! & \!\partial T_j \cap \partialD = \emptyset \\
\Vert \sigma[u_h[\alpha_h]] \cdot n - g\Vert_{0,2,\partial T} \! & \!\partial T_j \subset \Gamma_N
\end{array},
\right. & \!\!
\omega_{\partial T}^{(u)} &= \left\Vert u[\alpha] - \tilde{u}_h[\alpha_h] \right\Vert_{0,2,\partial T }.
\end{aligned}
\]
For the \emph{control residual} we get using \eqref{eq:effectivetensor}
\begin{eqnarray*}
&& \mathcal{L}_{,\vartheta} (\alpha_h,\vartheta_h;u_h[\alpha_h]) ( \vartheta - \tilde{\vartheta}_h ) \\
&&= \sum_j \int_{T_j} C_{,\vartheta}\left[\alpha_h,\vartheta_h \right] ( \vartheta - \tilde{\vartheta}_h ) \eps{u_h[\alpha_h]} : \eps{u_h[\alpha_h]} \;\mathsf{d\mathbf{x}} \\
&&= \sum_j \int_{T_j} R[\alpha_h] \, \bar{C}_{,m}\left[m_h,\theta_h\right] (m-\tilde{m}_h) \; \eps{u_h[\alpha_h]} : \eps{u_h[\alpha_h]} \\
&& \qquad \qquad + R[\alpha_h] \, \bar{C}_{,\theta}\left[m_h,\theta_h\right] (\theta-\tilde{\theta}_h) \; \eps{u_h[\alpha_h]} : \eps{u_h[\alpha_h]} \;\mathsf{d\mathbf{x}}
\end{eqnarray*}
Concerning the last expression we know that the controls $m_h$, $\theta_h$ being the laminate parameters are bounded.
Differentiation of the entries of the effective tensor \eqref{eq:effectivetensor} w.\,r.\,t.\ $m$ and $\theta$ again yields bounded expressions
within the attainable range of the parameters. The tensors $\bar{C}_{,m}$ and $\bar{C}_{,\theta}$ are therefore bounded and the
whole integrand is bounded in $L_1$.
We finally obtain the desired residual estimate with
\[
\begin{aligned}
\rho_T^{(m)} &= \left\Vert R[\alpha_h] \, \bar{C}_{,m}\left[m_h,\theta_h\right] \; \eps{u_h[\alpha_h]} : \eps{u_h[\alpha_h]}\right\Vert_{0,1,T} \,, &
\omega_T^{(m)} &= \left\Vert m-\tilde{m}_h \right\Vert_{0,\infty,T} \,, \\
\rho_T^{(\theta)} &= \left\Vert R[\alpha_h] \, \bar{C}_{,\theta}\left[m_h,\theta_h\right] \; \eps{u_h[\alpha_h]} : \eps{u_h[\alpha_h]}\right\Vert_{0,1,T} \,, &
\omega_T^{(\theta)} &= \left\Vert \theta-\tilde{\theta}_h \right\Vert_{0,\infty,T} \,.
\end{aligned}
\]
This proofs the claim.
\qed
\section{Estimation of weights using a priori regularity} \label{sec:apriori}
For the error estimates of Theorem~\ref{theorem:estimates} the residual terms $\rho_T$ can be computed straightforwardly
from the discrete solution $u_h$. The weights $\omega_T$, however, depend on the exact (unknown) solution $u$.
Using a priori known regularity those terms can be estimated. For the primal solution classical interpolation estimates exist.
For the laminate parameters it is possible to estimate the error by exploiting the explicit
relation to the primal solution \eqref{eq:explicitparameters}.
The following estimates especially ensure robustness of the error estimates for a sufficiently smooth continuous solution $u$:
\begin{corollary}[A priori estimates of the weights]
Under the assumption that $u \in W^{2,\infty}$ the following estimates for the weighting terms hold:
\begin{equation}
\begin{aligned}
\omega_T^{(u)} &= \left\Vert u-\tilde{u}_h \right\Vert_{0,2,T} && \lesssim h(T) \; |u|_{1,2,\omega_T} \,, \\
\omega_{\partial T}^{(u)} &= \left\Vert u-\tilde{u}_h \right\Vert_{0,2,\partial T} && \lesssim h(T)^{\frac12} \; |u|_{1,2,\omega_{\partial T}} \,, \\
\omega_T^{(m)} &= \left\Vert m-\tilde{m}_h \right\Vert_{0,\infty,T} && \lesssim h(T) |\ln h(T)| \; |u|_{2,\infty,T} \,, \\
\omega_T^{(\theta)} &= \left\Vert \theta-\tilde{\theta}_h \right\Vert_{0,\infty,T} && \lesssim h(T) |\ln h(T)| \; |u|_{2,\infty,T} \,.
\end{aligned}
\end{equation}
Here, $h(T)$ is the edge length of the square element $T$ and $\lesssim$ denotes the smaller or equal inequality up to a constant factor depending solely on the triangulation.
\end{corollary}
\textbf{Proof.} For the primal error weights we choose
$\tilde{u}_h = \Ih{2} u$, where $\Ih{2}$ denotes the piecewise bi-quadratic
Lagrangian interpolation.
Then standard interpolation estimates \cite[Theorem 3.1.5]{Ciarlet1978} give
\begin{align*}
& \Vert u - \Ih{2} u \Vert_{0,2,T} \lesssim h(T) |u|_{1,2,T} \,, \\
& \Vert u - \Ih{2} u \Vert_{0,2,\partial T} \lesssim h(T)^{\frac12} |u|_{1,2,\partial T} \,.
\end{align*}
Next, we estimate the error with respect to the control parameter $m$. The estimate for $\theta$ is completely analogous.
We choose $\tilde{m}_h = m[C^\ast \eps{\Ih{2} u}] =: \mathbf{m}(\lambda_1(\sigma_h),\lambda_2(\sigma_h))$,
i.e. $\tilde{m}_h$ is obtained by the pointwise evaluation of the interpolated solution $u$, the computation of the
corresponding stress $\sigma_h=C^\ast \eps{\Ih{2} u}]$ and its eigenvalues and finally inserting them into formulae \ref{eq:explicitparameters}.
Then, $\sup_{x \in T} | \mathbf{m}(\lambda_1(\sigma),\lambda_2(\sigma))-$ $\mathbf{m}(\lambda_1(\sigma_h),\lambda_2(\sigma_h)) |$
is attained for some $\bar{x} \in T$ with stresses $\bar{\sigma}=\sigma(\bar x)$, $\bar{\sigma}_h=\sigma_h(\bar x)$.
From the Lipschitz continuity of the function $\mathbf{m}$ we deduce
\[
\left|\mathbf{m}(\lambda_1(\bar{\sigma}),\lambda_2(\bar{\sigma}))-\mathbf{m}(\lambda_1(\bar{\sigma}_h),\lambda_2(\bar{\sigma}_h))\right|
\lesssim \left(|\lambda_1(\bar{\sigma})-\lambda_1(\bar{\sigma}_h)|^2+|\lambda_2(\bar{\sigma}) - \lambda_2(\bar{\sigma}_h)|^2 \right)^{\frac12} \!.
\]
Furthermore, the eigenvalues are Lipschitz continuous functions of the underlying matrices.
In particular the Wielandt-Hoffmann inequality \cite{HoWi53} for $p=2$ yields
\begin{align*}
&\left(|\lambda_1(\bar{\sigma})-\lambda_1(\bar{\sigma}_h)|^2+|\lambda_2( \bar{\sigma}) - \lambda_2(\bar{\sigma}_h)|^2 \right)^{\frac12}
\leq \Vert \bar{\sigma} - \bar{\sigma}_h \Vert_F \,.
\end{align*}
The elasticity tensor $C^\ast$ represents a bounded linear operator, thus
\[
\left\Vert \sigma - \sigma_h \right\Vert_{0,\infty,T}
= \left\Vert C (\eps{u} - \eps{\Ih{2} u} ) \right\Vert_{0,\infty,T}
\leq ||| C ||| \;
\Vert \eps{u} - \eps{\Ih{2} u} \Vert_{0,\infty,T}
\]
with $||| C |||$ representing the associated operator norm of the tensor $C$. Passing from the symmetrized strain tensor
to the full Jacobian yields
\[
\Vert \eps{u} - \eps{\Ih{2} u} \Vert_{0,\infty,T} \lesssim |u - \Ih{2} u |_{1,\infty,T} \,.
\]
Finally, by standard $L^\infty$ estimates \cite[Theorem 3.3.7]{Ciarlet1978}, we achieve
\[
|u - \Ih{2} u |_{1,\infty,T} \lesssim h(T) |\ln h(T)| \; |u|_{2,\infty,T} \,.
\]
\qed
\begin{remark}
The presented argument would not work for a weighting term like $\left\Vert \alpha-\tilde{\alpha}_h \right\Vert_{0,\infty,T}$
involving the rotation parameter as it depends on the eigenvectors of the local stress and the eigenvectors
do not depend continuously on the matrix.
In fact this motivates the elimination of the term involving the rotation parameter in Theorem~\ref{theorem:estimates}.
\end{remark}
\section{Numerical treatment} \label{sec:localapprox}
For a practical numerical scheme for the weighting terms $\omega_T$ that were a priori estimated in section~\ref{sec:apriori} we ask for an
approximation based on the numerical solution $u_h$.
One approach would be to exploit higher regularity of the solution by computing a second discrete solution using a higher order scheme.
However this is usually computationally demanding. To keep the cost for the evaluation of the error estimates low \cite{BeRa97} suggests to merely interpolate
the discrete solution to a higher order polynomial.
As we use bi-quadratic finite elements to overcome checkerboard instabilities, cf.~\cite{BoJo98}, we therefore construct polynomials of degree
four on patches of four neighboring elements from the local degrees of freedom of the discrete elastic displacement $u_h$.
In the discrete problem \eqref{eq:discrete} the coefficient functions $q$ and thus $C^\ast$ vary from point to point.
However it is known, see \cite[Theorem 4.1.6]{Ciarlet1978}, that the convergence order of the finite element scheme is unaltered
if the chosen quadrature order is high enough, which is the case in our computations.
To this end the parameters $\alpha$, $m$ can easily be evaluated at the quadrature points. The same holds true for the density parameter as well.
However it needs to be piecewise constant on each element to avoid
checkerboard instabilities, see \cite{BoJo98}. This is ensured by an element wise averaging of all
values obtained on quadrature points.
In detail we proceed as follows.
In case of the ratio parameter $m$ (and the rotation parameter $\alpha$ which however was eliminated from the error estimate)
the interpolated discrete deformation $u_h$ can be used again to locally obtain parameters $\alpha$ and $m$ for fixed $\theta$.
For the density parameter the piecewise constant approximation can be used to construct a bilinear profile on a patch of four neighboring
elements. Let us summarize the approximation of the weights in the following remark:
\begin{remark}[Approximation of local weights]
To approximate the weights for the primal and the control error indicator on every element $T$ and to derive a suitable error indicator
we consider a patch $\mathcal{S}$ of the four neighboring elements including $T$ and proceed as follows:
\begin{itemize}
\item[-] For the primal error indicator $5 \times 5$ degrees of freedom on each element of the patch are used to construct a bi-quartic
polynomial on the patch.
This interpolation $\Ih{4} u_h$ is then compared to the original approximation leading to
\[
\omega_T^{(u)} \approx \left\Vert u_h-\Ih{4}u_h \right\Vert_{0,2,T}\,, \qquad
\omega_{\partial T}^{(u)} \approx \left\Vert u_h-\Ih{4}u_h \right\Vert_{0,2,\partial T}\,.
\]
\item[-]
For the control error indicator the ratio parameter $m$ is
computed from the interpolated elastic solution above and compared to the original value;
for the density parameter $\theta$ four piecewise constant values are used
to construct a bilinear profile on the patch which is
then compared to the original values. In explicit
\[
\omega_T^{(m)} \approx \left\Vert m[u_h] - m[\Ih{4}u_h] \right\Vert_{0,\infty,T}\,, \qquad
\omega_T^{(\theta)} \approx \left\Vert \theta_h - \Ih{1}\theta_h \right\Vert_{0,\infty,T}\,.
\]
\end{itemize}
\end{remark}
\section{Implementation} \label{sec:impl}
For our computations we use an adaptive regular quadrilateral mesh provided by the \texttt{QuocMesh} library
\footnote{\url{http://numod.ins.uni-bonn.de/software/quocmesh}}.
Refinement is done via
uniform refinement of cells and handling of constrained hanging nodes. To obtain the optimal sequential lamination microstructure we
reimplemented the alternating algorithm suggested in \cite{AlBoFr97}.
The effective tensors are regularized by setting $C_{44} = 10^{-2}$. To ensure coercivity we restrict $m$ to the range
$[\varepsilon,1-\varepsilon]$ and $\theta$ to $[\varepsilon,1]$ with $\varepsilon=10^{-3}$.
The algorithm is terminated once the change in the compliance cost value is below $10^{-7}$.
As already mentioned we use bi-quadratic elements to overcome checkerboard instabilities.
For the numerical integration it turned out to be sufficient to use a Gauss quadrature rule of order 5.
The lamination parameters for the interpolated solution cannot be directly computed from the pointwise stress
$C^\ast \, \eps{\Ih{4} u_h}$ as it in turn depends on the unknown parameters.
Here we use a Newton method to solve the equation
\[
\bar{C}[m(\lambda_i),\bar{\theta}] \, \eps{\Ih{4} u_h} - Q(\alpha) \begin{pmatrix} \lambda_1 & \\ & \lambda_2 \end{pmatrix} Q(\alpha)^\top = 0
\]
for the rotation parameter $\alpha$ and the eigenvalues $\lambda_i$ of the stress (from which $m$ can be computed).
The density $\bar{\theta}$ is fixed to the original value found at the current point.
For the adaptive scheme we use a D\"orfler marking strategy \cite{Do96} with a fraction of $40\%$ of the total error to be associated with
the set of elements to be refined.
\section{Numerical results} \label{sec:results}
As the first example we consider a {\it carrier plate under shearing} computed on the unit square., cf.~Fig.~\ref{fig:carrier}.
The volume constraint was set to $33\%$.
First we ran the alternating algorithm until convergence on uniform meshes of level $l \in \{2,\ldots,10\}$.
The obtained values can be used to extrapolate the asymptotic value of the compliance cost by assuming the following
dependence on the grid width:
\[
J[u_h] = J^\ast + c \, h^p \,.
\]
Using a least squares fit we found the parameters $J^\ast = 1{.}8399$, $c = 1{.}7645$, and $p = 1{.}0484$.
$J^\ast$ is used to replace the exact cost value when assessing error reduction in the following.
We ran the discussed adaptive scheme for the carrier plate scenario, see Figures~\ref{fig:carrier} and \ref{fig:carrier2}.
\begin{figure}[!ht]
\includegraphics[width=.18\linewidth]{test_lvl4_1e6_grid_iter004.png}
\hfill
\includegraphics[width=.18\linewidth]{test_lvl4_1e6_grid_iter007.png}
\hfill
\includegraphics[width=.18\linewidth]{square_lvl4_1e6_iter008_qDenResWei.png}
\hfill
\includegraphics[width=.18\linewidth]{test_lvl4_1e6_grid_iter008.png}
\hfill
\includegraphics[width=.18\linewidth]{test_lvl4_1e6_grid_iter010.png}\\[3pt]
\includegraphics[width=.18\linewidth]{test_lvl4_1e6_grid_iter013.png}
\hfill
\includegraphics[width=.18\linewidth]{test_lvl4_1e6_grid_iter016.png}
\hfill
\includegraphics[width=.18\linewidth]{test_lvl4_1e6_grid_iter019.png}
\hfill
\includegraphics[width=.18\linewidth]{square_lvl4_1e6_iter020_qDenResWei.png}
\hfill
\includegraphics[width=.18\linewidth]{test_lvl4_1e6_grid_iter020.png}
\caption{Adaptive meshes after 4, 7, 8, 10, 13, 16, 19, and 20 refinement steps obtained for the carrier plate scenario.
Color plots show the error indicator after 7 and 19 steps, leading to the subsequent grid.}
\label{fig:carrier}
\end{figure}
\begin{figure}[!ht]
\hfill
\includegraphics[width=.4\linewidth]{square_error}
\hfill\mbox{}
\caption{
Difference to extrapolated value $J^\ast$ plotted
over the number of elements, once for a uniform refinement (green) and the adaptive refinement strategy (red).
}
\label{fig:carrier2}
\end{figure}
Next, we take into account a {\it cantilever scenario}. Here we consider the domain $D = [0,2] \times [0,1]$
with Dirichlet boundary condition on the left hand side and a downward pointing force in the middle on the right hand side. The
volume constraint is set to $50\%$, cf. Figure~\ref{fig:cantilever}.
\begin{figure}[!ht]
\includegraphics[width=.3\linewidth]{cantilever-04-24_grid_iter004.png}
\hfill
\includegraphics[width=.3\linewidth]{cantilever-04-24_grid_iter008.png}
\hfill
\includegraphics[width=.3\linewidth]{cantilever-04-24_grid_iter012.png}\\[3pt]
\includegraphics[width=.3\linewidth]{cantilever-04-24_grid_iter016.png}
\hfill
\includegraphics[width=.3\linewidth]{cantilever-04-24_grid_iter020.png}
\hfill
\includegraphics[width=.3\linewidth]{cantilever-04-24_grid_iter024.png}\\
\mbox{\rule{\linewidth}{.4pt}}\\[12pt]
\includegraphics[width=.3\linewidth]{cantilever-04-24_density_iter004.png}
\hfill
\includegraphics[width=.3\linewidth]{cantilever-04-24_density_iter008.png}
\hfill
\includegraphics[width=.3\linewidth]{cantilever-04-24_density_iter012.png}\\[3pt]
\includegraphics[width=.3\linewidth]{cantilever-04-24_density_iter016.png}
\hfill
\includegraphics[width=.3\linewidth]{cantilever-04-24_density_iter020.png}
\hfill
\includegraphics[width=.3\linewidth]{cantilever-04-24_density_iter024.png}\\
\caption{At different refinement steps of the adaptive method applied to the cantilever example the current mesh (top) and the corresponding density $\theta$ (bottom) are displayed (from left to right: step 4, 8, 12, 16, 20, and 24).}
\label{fig:cantilever}
\end{figure}
As a third scenario we investigate a {\it bridge} within the domain $D = [0,2] \times [0,1]$ in Figure~\ref{fig:bridge}.
We prescribe roller boundary conditions on short strips at the lower boundary in the left and right corners, i.\,e.\
the displacement in $y$-direction is enforced to be $0$, compare for instance \cite{Al02}.
Additionally the node in the lower left corner is subject
to full homogeneous Dirichlet boundary conditions to ensure a unique solution. The lower boundary in between is loaded
in vertical direction. The volume constraint is set to $33\%$.
\begin{figure}[!hb]
\hfill
\includegraphics[width=.4\linewidth]{bridge_pull_laminates_1e7_lvl4_doerflerD30_grid_iter022}
\hfill
\includegraphics[width=.4\linewidth]{bridge_pull_laminates_1e7_lvl4_doerflerD30_density_iter022}
\hfill\mbox{}
\caption{Adaptively refined grid and corresponding density $\theta$ for the bridge scenario.}
\label{fig:bridge}
\end{figure}
\section{Conclusion} \label{sec:sum}
In this paper we have applied the dual weighted residual approach to an elastic multi scale shape optimization
problem using sequential laminates as an underlying optimal microstructure. The microstructure
enters the PDE via effective material properties whose explicit formulae allow to compute
sensitivities of the cost functional with respect to the describing parameters of the microstructure.
The derived goal-oriented error estimate enables to appropriately steer the adaptive refinement
leading to a substantial reduction of the number of degrees of freedom compared to uniform meshes and the same accuracy.
In particular the goal oriented error estimation outperforms a residual type estimator solely for the elastic displacement problem.
The main ingredient is the error term incorporating the sensitivity of the elastic energy density w.\,r.\,t.\ to the local material density.
This is in accordance with the observation that the rigidity of a shape is mostly improved by an optimal distribution of the available material.
The adaptive scheme leads to sharply resolved interfaces, separating regions with
almost no material from regions with bulk material or an intermediate material density.\\
\section*{Acknowledgments}
This work has been supported by the \emph{Deutsche Forschungsgemeinschaft} through
\emph{Collaborative Research Centre 1060 -- The Mathematics of Emergent Effects}.
\bibliographystyle{alphadin}
|
\section{Introduction}
\label{sec:intro}
In the years 1928 -- 1931 Dirac introduced the concept of antimatter\cite{dir28, dir30,dir31}. In 1932 Anderson showed that these theoretical concepts were real with the detection of the positron\cite{and33}. Then in 1934 Mohorovi{\v c}i{\'c} published a paper on the possibility of observing positronium (Ps) from astrophysical sources via its recombination lines\cite{moh34}. This prescient paper went largely unnoticed, and the existence of Ps was predicted independently and separately by Pirenne \cite{pir44}, Ruark\cite{rua45}, Landau (unpublished work referred to in \citet{ali45}) and Wheeler\cite{whe46}.
To date, the observational challenge laid down by Mohorovi{\v c}i{\'c}, viz.\ to detect astrophysical sources of Ps from its recombination lines, has not been met. However the possibility of detection of Ps recombination lines has been raised on subsequent occasions inspired by advances in technology\cite{mcc84,burd97,ell09}.
Despite the non-detection of Ps recombination lines, there is clear evidence of its existence in the Galaxy from its annihilation signature\cite{lev78}; see the recent review by \citet{pra11}.
The observations of annihilating electrons and positrons in our Galaxy have raised several important questions for astrophysics, the foremost being what are the sources of the positrons? Can they be explained with conventional astrophysics (e.g.\ high energy sources) or are more exotic sources required (e.g.\ annihilating dark matter)? Why is the annihilation radiation centered so strongly on the nucleus of the Galaxy with relatively little disc emission? How far do positrons disperse through the Galaxy before annihilation?
In this paper we address the question of what other observable signatures there are, arising from other leptonic atoms. For example, it is possible in principle to form atoms of $\mu^{-}$ -- $\mu^{+}$\ and $\tau^{-}$ -- $\tau^{+}$. However, in these cases the observable signatures are complicated by the fact that the constituent particles are themselves unstable, and may decay prior to annihilation or transitions\cite{avi79}. It is important to note that even improbable signatures may be significant if the rate of production of the atom is high enough; this situation is often very relevant in astrophysical situations. For example the hydrogen 21 cm emission line that results from the spin-flip transition between the hyperfine splitting of the $^{2}$S ground state has been hugely exploited in radio astronomy since the 1950s\cite{ewe51,mull51}, although it has a transition probability of only $2.9 \times 10^{-15}$ s$^{-1}$.
Even faint signatures would be of interest. First, they would be the first such observations of $\mu^{-}$ -- $\mu^{+}$\ or $\tau^{-}$ -- $\tau^{+}$, either in astrophysics or in the laboratory. Secondly they would place constraints on the energetics and rate of production of $\mu$ or $\tau$ in astrophysical sources. Thirdly, the rate of $\mu^{-}$ -- $\mu^{+}$\ or $\tau^{-}$ -- $\tau^{+}$ production, compared with estimates of $\mu$ or $\tau$ production rates, would place constraints on the physical conditions during recombination.
Such measurements could be very useful. Taking the example of Ps, the rate, distribution and energetics of formation are all well measured, but the identity of positron sources still remains unclear. This is largely because it is difficult to reproduce the distribution of Ps emission, and also because there are many \e+\ candidates, which are difficult to distinguish observationally. Observations, or even non-detection, of $\mu^{-}$ -- $\mu^{+}$\ or $\tau^{-}$ -- $\tau^{+}$\ could place further constraints on the competing models for the origin of Galactic positrons.
The study of such processes is important, even if they are unlikely. Anomalous spectral signatures are routinely discovered, for example the recent detection of an unidentified emission line in the X-ray spectra of clusters of galaxies at 3.5~keV\cite{boy14,bul14}. We note that this line is not attributable to $\mu^{-}$ -- $\mu^{+}$\ or $\tau^{-}$ -- $\tau^{+}$, although these species are expected to produce X-ray emission lines, as we show. Nevertheless, in such cases it is important to rule out prosaic, but poorly understood processes before accepting more exotic physics. The formation of $\mu^{-}$ -- $\mu^{+}$\ and $\tau^{-}$ -- $\tau^{+}$\ in astrophysical environments has so far received very little attention, but should not be rejected \emph{a priori} on the basis of the short lifetimes of the $\mu$ and $\tau$.
The production of $\mu^{-}$ -- $\mu^{+}$\ and $\tau^{-}$ -- $\tau^{+}$\ in electron-positron colliders is difficult due to the rapid decay times of the particles, although methods have been proposed for the creation of $\mu^{-}$ -- $\mu^{+}$\cite{brod09}. It may be possible to bypass these difficulties if large numbers of leptonium atoms are produced in high energy astrophysical environments. This is an important possibility since neither $\mu^{-}$ -- $\mu^{+}$\ nor $\tau^{-}$ -- $\tau^{+}$\ have ever been observed, and together they are the most compact pure QED systems that could exist. An astrophysical detection would be more difficult to interpret than a collider detection, but would nevertheless constitute an important basic test of QED.
We wish to make a first step in addressing this challenge by calculating the expected observable signatures of astrophysical sources.
Consider a source of particle - anti-particle lepton pairs. Their possible fates are, (i) the particle immediately decays, if it is $\mu^{\pm}$ or $\tau^{\pm}$, (ii) it directly annihilates with its antiparticle, (iii) it radiatively recombines with a free particle, then annihilates, (iv) it forms a new atom via charge-exchange, then annihilates.
Ps clearly has the highest probability of forming, since its constituent particles are stable, and can travel through the interstellar medium until they form Ps through charge-exchange or radiative recombination, although they may also directly annihilate in-flight or when thermalized. The probabilities for Ps formation have been discussed in detail\cite{mcc84,gou89,wal96}, and Ps formation at the Galactic centre is confirmed by the observation of the 511 keV annihilation signature and the three-photon triplet annihilation continuum\cite{jea06}, from which a Ps formation fraction of $\approx 97$ per cent is inferred. See \citet{pra11} for a recent review of Ps astrophysics.
The formation of any atom having a $\mu$ or $\tau$ lepton is much less likely, since the decay times are so short that no travel of the particles is possible. Thus, we only consider the formation of true muonium and true tauonium (hereafter referred to as M and T\footnote{The nomenclature for such systems is somewhat confusing. A bound state of a particle and its anti-particle is called an onium, however the name muonium is already given to the combination of an anti-muon with an electron to form a hydrogen like atom, and by analogy tauonium refers to an anti-tauon and an electron. Therefore the systems $\mu^{-}$ -- $\mu^{+}$\ and $\tau^{-}$ -- $\tau^{+}$\ are usually referred to as \emph{true} muonium and \emph{true} tauonium, however to avoid confusion, and for brevity, we prefer to use the symbols M and T}) \emph{in situ}, i.e.\ the particle - anti-particle pairs must be created with an energy less than the ionisation energy, so that the particle forms immediately. A summary of the basic properties of M, T, and Ps, to which frequent comparisons are made, is given in Table~\ref{tab:summ}; the calculations or references for these values can be found throughout the paper, in the sections given in the final column of the table.
\begin{table*}
\caption{Summary of the main properties of Ps, M and T.}
\label{tab:summ}
\begin{tabular}{llllll}
&& Ps & M & T & Section\\ \hline
\multicolumn{2}{l}{Rest mass/annihilation energy (MeV)} & 0.511& 105.66& 1406.6 &\\
\multicolumn{2}{l}{Ionisation energy (eV)} &6.8&1784.1&23751.4 & \S\ref{sec:mechanisms}\\
\multicolumn{2}{l}{Bohr radius (m)} &$1.058 \times 10^{-10}$ &$5.199 \times 10^{-13}$& $3.044 \times 10^{-14}$ & \S\ref{sec:decay_rad} \\
\multicolumn{2}{l}{Decay time of constituent particles (s)} &$\infty$&$2.197 \times 10^{-6}$ & $2.874 \times 10^{-13}$& \S\ref{sec:mutaudecay}\\
$\gamma\gamma$ Annihilation time (ground state) (s)& Singlet&$1.2 \times 10^{-10}$&$6.0 \times 10^{-13}$& $3.6 \times 10^{-14}$&\S\ref{sec:decay_ann}\\
& Triplet &$1.4 \times 10^{-7}$& $6.7 \times 10^{-10}$ &$4.0 \times 10^{-11}$ & \S\ref{sec:decay_ann}\\
e$^{-}$e$^{+}$ Annihilation time (ground state) (s)& Triplet &--&$1.8 \times 10^{-12}$& $1.1 \times 10^{-13}$& \S\ref{sec:decay_ann}\\
Recombination time ($nL \rightarrow n'L'$) (s) & $2\ 1 \rightarrow 1\ 0$ (Lyman $\alpha$)&$3.191\times 10^{-9}$ &
$1.543\times 10^{-11}$ & $9.18\times 10^{-13}$ & \S\ref{sec:decay_rad} \\
&$3\ 1 \rightarrow 1\ 0$ (Lyman $\beta$) & $1.195\times 10^{-8}$ & $ 5.780\times 10^{-11}$ & $3.44\times 10^{-12}$& \S\ref{sec:decay_rad} \\
& $3\ 1 \rightarrow 2\ 0$ & $8.905\times 10^{-8}$ &$ 4.307\times 10^{-10}$ & $2.56\times 10^{-11}$ & \S\ref{sec:decay_rad} \\
&$3\ 0 \rightarrow 2\ 1$ & $3.166\times 10^{-7}$ & $1.531\times 10^{-9} $&$ 9.11\times 10^{-11}$ & \S\ref{sec:decay_rad} \\
& $3\ 2 \rightarrow 2\ 1$ & $3.092\times 10^{-8}$ &$ 1.495\times 10^{-10} $& $8.89\times 10^{-12}$ & \S\ref{sec:decay_rad} \\
& $4\ 1 \rightarrow 2\ 0$ &$2.068\times 10^{-7} $&$ 9.999\times 10^{-10} $& $5.95\times 10^{-11}$& \S\ref{sec:decay_rad} \\
& $4\ 0 \rightarrow 2\ 1$ &$7.753\times 10^{-7} $&$ 3.750\times 10^{-9} $& $2.23\times 10^{-10} $& \S\ref{sec:decay_rad} \\
& $4\ 2 \rightarrow 2\ 1$ &$9.692\times 10^{-8}$ &$ 4.687\times 10^{-10} $&$ 2.79\times 10^{-11}$& \S\ref{sec:decay_rad} \\
Recombination energies (keV) & Lyman $\alpha$ &$5.102 \times 10^{-3}$&1.055&17.7 & \S\ref{sec:decay_rad} \\
&Lyman $\beta$ &$6.047 \times 10^{-3}$&1.250& 21.0& \S\ref{sec:decay_rad} \\
& Balmer $\alpha$ &$9.448 \times 10^{-4}$&0.195&3.3& \S\ref{sec:decay_rad} \\
&Balmer $\beta$ &$1.276 \times 10^{-3}$&0.264& 4.4& \S\ref{sec:decay_rad}
\end{tabular}
\end{table*}
To calculate the likelihood of forming M and T, we begin by reviewing the possible mechanisms for pair-production which are relevant in astrophysical environments (section~\ref{sec:mechanisms}). We then use the relevant cross-sections for pair-production, to calculate the pair-production spectra, and from these calculate the fraction of pairs which have sufficiently low energy to form an onium immediately, i.e.\ the fraction of pairs produced with kinetic energy less than the ionisation energy of the atom (section~\ref{sec:pairprodspec}). These pairs may either directly annihilate, or radiatively recombine, and we calculate the branching fraction for these processes in section~\ref{sec:gould}. For those that do radiatively recombine, we calculate the cross-sections for radiative recombination, and hence the fractions which recombine into the $nL$th quantum level in section~\ref{sec:wallyn}. Then in section~\ref{sec:decay} we calculate the different decay channels and in section~\ref{sec:branch} the resulting branching ratios for an onium in the $n$th level, and thus the total branching fractions for specific observational signatures. The signatures themselves are discussed in section~\ref{sec:signatures}. This brings us to a position where, for a given rate of pair-production, we can estimate the luminosity of each of the different decay channels, and we apply these calculations to specific astrophysical sources in section~\ref{sec:astrophys}. We give our conclusions in section~\ref{sec:discuss}.
\section{Pair-production processes in astrophysical environments}
\label{sec:mechanisms}
The formation of leptonium other than Ps is not complicated by possibilities of escape, in-flight formation, charge-exchange etc.; the lifetimes of the particles themselves are so short (\S~\ref{sec:decay}), that if an atom is not formed immediately after the creation of a particle, then the particle itself will decay before any of these processes can occur. Ultra-relativistic particles may live longer due to time dilation, but as for \e+\cite{bus79}, the annihilation signature will give rise to very weak continuum emission.
We consider the following production processes for leptons, $l$,
\begin{eqnarray}
\pi^{-} \rightarrow l^{-} + \overline{\nu}_{l}, \label{eqn:piminusmu}\\
\pi^{+} \rightarrow l^{+} + \nu_{l} \label{eqn:piplusmu}, \\
\gamma + \gamma \rightarrow l^{-} + l^{+}, \label{eqn:mupair}\\
e^{-} + e^{+} \rightarrow l^{-} + l^{+} \label{eqn:mueepair}.
\end{eqnarray}
The pion decay processes will occur in collisions of highly energetic ($>200$ MeV) cosmic rays with protons in the ISM\cite{gue05}. The third process requires very high photon energies and will occur in the high energy environments surrounding compact objects, such as black-holes and neutron stars\cite{gue05}. The last process may occur if the \e+\ have large kinetic energy, and collide with e$^{-}$ in the ISM, e.g.\ if an e$^{\pm}$ jet from an AGN or microquasar collides with an interstellar cloud of gas or a star.
We can immediately dismiss the pion decay processes as irrelevant for $\tau$ leptons, since the $\tau$ are more massive than the $\pi^{\pm}$. We can also dismiss the pion decay processes as irrelevant for the formation of M, on two grounds. First, only a single particle is produced, and the probability of radiative recombination with another free $\mu$ before it decays is negligible. Secondly, the energetics are unfavorable. Since the reaction involves a single particle of known mass decaying into two particles, the energetics are precisely determined in the zero-momentum frame; the $\mu$ is created with a kinetic energy of 4.12~MeV, which is much higher than the ionisation energy of M, $\approx 1.4$~keV. Thus the created $\mu$ must first lose energy, then collide with its antiparticle, before either particle decays, the probability of which is negligible.
This leaves us with photon-photon pair production and electron-positron pair production. In both these processes the $\mu$ (or $\tau$) will be created with typical energies much higher than the ionisation energy of M (or T) ($\approx 1.4$~keV for M, and $\approx 23.7$~keV for T). Only the small fraction of pairs whose total kinetic energy in the zero momentum (z.m.) frame is less than the ionisation energy can form an onium, i.e.,
\begin{eqnarray}
T_{1}+T_{2} \le E_{\rm ion}, \label{eqn:oniumenergy} \\
2 (\gamma-1) m c^{2} \le \frac{\mu q^{4}}{32 \pi^{2} \epsilon_{0}^{2} \hbar^{2}} ,
\end{eqnarray}
where the right hand side is the ionisation energy in S.I. units, $m$ is the mass of the lepton, $\mu$ is the reduced mass of the atom, $q$ is the electron charge, and $\gamma$ is the Lorentz factor of the produced pairs in the z.m.\ frame. Therefore the maximum Lorentz factor in the z.m.\ frame that the particles can have and still form an onium is,
\begin{equation}
\label{eqn:gammalimit}
\gamma_{\rm lim}=1+\frac{q^{4}}{128 \pi^{2} c^{2} \epsilon_{0}^{2} \hbar^{2}}=1+6.656 \times 10^{-6}.
\end{equation}
Thus, only very low energy pairs can produce an onium, and $\gamma_{\rm lim}$ is independent of mass. We now estimate the fraction of pairs produced with $\gamma < \gamma_{\rm lim}$, for various formation mechanisms.
\section{The pair-production spectrum}
\label{sec:pairprodspec}
Throughout this section we will refer to $\mu$ and M, but the arguments are identical for $\tau$ and T.
\subsection{Pair production by photon-photon annihilation}
\label{sec:gg}
For two photons with energy $E_{1}$ and $E_{2}$ colliding at an angle $\theta$, pair production will occur if,
\begin{equation}
\label{eqn:pairprod}
E_{1}E_{2} \ge \frac{2 \left( m c^{2}\right)^{2}}{1-\cos \theta},
\end{equation}
where $m=2\mu$ is the mass of an individual particle. The total energy of either produced particle, $E_{0}$ is given by,
\begin{equation}
\label{eqn:pairprodenergy}
E_{0}^{2}=\gamma^{2} m^{2} c^{4} = \frac{E_{1}E_{2} (1 - \cos \theta)}{ 2},
\end{equation}
where $\gamma$ is the Lorentz factor of either of the created particles in the z.m.\ frame.
Many authors\cite{gou67,bon71,aga83,cop90,boe97} have evaluated the pair-production spectrum (usually of e$^{\pm}$) based on various assumptions about the distributions of photon energies. Here we repeat these calculations, except we integrate the resulting spectra in order to obtain the fraction of produced pairs which have sufficiently low energy to form M, i.e.\ those which obey equation~\ref{eqn:gammalimit}.
The total cross-section for the reaction $\gamma + \gamma \rightarrow \mu^{-} + \mu^{+}$ is\cite{jau55},
\begin{eqnarray}
\sigma &= &F_{\rm c} \frac{q^{4}}{32 \pi \epsilon_{0}^{2} m^{2} c^{4}} \nonumber \\
&&(1-\beta^{2})\left((3-\beta^{4})\ln\frac{1+\beta}{1-\beta} - 2\beta(2-\beta^{2})\right) \label{eqn:ggsigma}
\end{eqnarray}
\begin{equation}
F_{\rm c} =\frac{\pi\frac{\alpha}{\beta}}{1 - {\rm e}^{-\pi\frac{\alpha}{\beta}}}. \label{eqn:ssc}
\end{equation}
where $\beta=u/c$, and $u$ is the velocity of the outgoing particles in z.m.\ frame, and $F_{\rm c}$ is the Sommerfeld-Sakharov correction to the cross-section, due to the attractive Coulomb force between the created $\mu^{\pm}$, which applies near the threshold energy\cite{sak48,sak91,vol03}.
Now consider a source of photons whose number density per unit energy is given by $n_{1}(E_{1})$. Let these photons interact with an isotropic photon gas whose number density per unit energy is given by $n_{2}(E_{2})$. The fraction of photons in the isotropic gas which lie within a differential cone at angle $\theta$ and width ${\rm d}\theta$ to the incoming photon is $\frac{1}{2} \sin \theta {\rm d}\theta$.
The reaction rate in the lab frame, i.e.\ the number of pairs created per unit volume per unit time is thus given by,
\begin{eqnarray}
R_{\gamma \gamma} &= &
\int_{0}^{\infty} \int_{-1}^{1} \int_{E_{1 {\rm min}}}^{E_{1 {\rm max}}} \nonumber \\ &&n_{1}(E_{1}) n_{2}(E_{2}) \sigma(\beta) \frac{c}{2}(1 - \cos \theta) \nonumber \\
&&{ {\rm d}E_{1}\ \rm d}\cos\theta\ {\rm d}E_{2} \label{eqn:rgg}
\end{eqnarray}
%
where the factor $c(1-\cos \theta)$ is the relative velocity of the incoming photons, and the lower limit on the integral over $E_{1}$,
%
\begin{equation}
E_{1 {\rm min}} =\frac{2 m^{2} c^{4}}{E_{2}(1 - \cos \theta)}
\end{equation}
%
from equation~\ref{eqn:pairprod}, ensures pair production is possible. The upper limit on $E_{1}$ is either
%
\begin{equation}
E_{1 {\rm max}} =E_{1 {\rm M}}=\frac{2 \gamma_{\lim}^{2} m^{2} c^{4}}{E_{2}(1 - \cos \theta)}
\end{equation}
to specify only the produced pairs which satisfy equation~\ref{eqn:gammalimit}, or
%
\begin{equation}
E_{1 {\rm max}} =E_{1 \infty}=\infty
\end{equation}
for all produced pairs.
The fraction of pairs which can produce M, $f_{\gamma \gamma}$, is thus given by the ratio of R integrated over both these limits.
We have evaluated equation~\ref{eqn:rgg} numerically for different cases of photon energy distribution. Because we are interested in the ratio of $R$ for different limits in $\gamma$, and because $E_{1}$ and $E_{2}$ are related to $\gamma$ via equation~\ref{eqn:pairprodenergy}, the form for the second distribution $n(E_{2})$ does not matter; integrating over $E_{2}$ leads to a different constant which cancels when taking the ratio. Thus, $f_{\gamma \gamma}$ is identical for a specific power-law interacting with any isotropic photon gas. The fraction $f_{\gamma \gamma}$ for different power-law indices are given in Table~\ref{tab:ggpl}.
\begin{table}
\center
\caption{The fraction of pairs produced via photon-photon annihilation with sufficiently low energy to form an onium, for photon distributions with a power-law energy density with index $-\alpha$.}
\label{tab:ggpl}
\begin{tabular}{lc}
$\alpha$ & $f_{\gamma \gamma}$\footnote{Since the fraction $f_{\gamma \gamma}$ is independent of mass this is the fraction for any onium, Ps, M or T.} \\ \hline
1& $2.0 \times 10^{-7}$\\
1.5 & $3.8 \times 10^{-7}$\\
2 & $6.1 \times 10^{-7}$\\
2.5& $8.9 \times 10^{-7}$
\end{tabular}
\end{table}
\subsection{Pair production by electron-positron annihilation}
\label{sec:ee}
The total cross-section for the process $e^{-} + e^{+} \rightarrow \mu^{-} + \mu^{+}$ is given by\cite{per92},
\begin{equation}
\sigma =F_{\rm c} \frac{\pi \alpha^{2} \hbar^{2} \beta \left(\beta^{4}-4 \beta^{2}+3\right)}{6 M^{2} c^{2}}
\end{equation}
where $M$ is the muon mass, and $F_{\rm c}$ is the Sommerfeld-Sakharov factor given by equation~\ref{eqn:ssc}.
The fraction of those muons produced which have sufficiently low energy that they could immediately form M now given by
\begin{eqnarray}
R_{\rm ee} &= &
\int_{0}^{\infty} \int_{-1}^{1} \int_{E_{1 {\rm min}}}^{E_{1 {\rm max}}} \nonumber \\ &&n_{1}(E_{1}) n_{2}(E_{2}) \sigma(\beta) \frac{v}{2} \nonumber \\
&&{ {\rm d}E_{1}\ \rm d}\cos\theta\ {\rm d}E_{2} \label{eqn:ree}
\end{eqnarray}
%
where $v$ is the relative speed of the electron and positron in the lab frame and $n_{1}(E_{1})$ and $n_{2}(E_{2})$ are now the distributions of the lab frame number density per unit energy of the electrons and positrons.
The lab frame energies $E_{1}$ and $E_{2}$ are related to the Lorentz factors of the produced pairs by equation~\ref{eqn:eemmen}, and the relative velocity in the lab frame, $v$, is given by equation~\ref{eqn:eemmv}. Therefore changing variables to integrate over $\gamma$ instead of $E_{1}$, we can calculate $R_{{\rm ee}_{\gamma_{\rm lim}}}$ in the limit $\gamma_{\rm max} = \gamma_{\rm lim}$, and $R_{{\rm ee}_{\infty}}$ in the limit $\gamma_{\rm max} = \infty$, and thus the fraction of produced pairs with sufficiently low energy to form M is given by,
%
\begin{equation}
\label{eqn:fMee}
f_{\rm ee} = \frac{R_{{\rm ee}_{\gamma_{\rm lim}}}}{R_{{\rm ee}_{\infty}}}.
\end{equation}
%
As for photon-photon annihilation the exact energy density of $E_{2}$ does not matter, since it cancels when taking the ratio, and the result is independent of the mass of the produced particles. We have calculated $f_{\rm ee}$ for various power-law indices, where we evaluated the integrals numerically, and we show the results in Table~\ref{tab:eemm}.
\begin{table}
\center
\caption{The fraction of pairs produced via electron-positron annihilation with sufficiently low energy to form an onium, for electron distributions with a power-law energy density with index $-\alpha$.}
\label{tab:eemm}
\begin{tabular}{lc}
$\alpha$ & $f_{\rm ee}$\footnote{Since the fraction $f_{\rm ee}$ is independent of mass this is the fraction for any onium, Ps, M or T.} \\ \hline
1& $5.7 \times 10^{-7}$\\
1.5 & $9.1 \times 10^{-7}$\\
2 & $1.3 \times 10^{-6}$ \\
2.5& $1.7 \times 10^{-6} $
\end{tabular}
\end{table}
\section{Cross-sections for radiative recombination and direct annihilation}
\label{sec:gould}
In the previous section we calculated the fraction of pairs produced, either via photon-photon annihilation, or by electron positron annihilation, which have sufficiently low energy to form Ps, M or T immediately.
However, the fraction of these pairs which will form an onium is less still. As for Ps\cite{gou89} there is a still a probability that some of these pairs will directly annihilate; though in the case of $\mu$ and $\tau$ we can neglect charge exchange since the probability of the particles decaying before meeting a neutral atom is large.
We have calculated the total reaction rates for radiative recombination and direct annihilation following \citet{gou89} (see also \citet{gou72}). Figure~\ref{fig:rrda} shows the fraction of pairs which radiatively recombine, $g_{\rm rr}$, or directly annihilate, $g_{\rm da}$, as a function of temperature. For e$^{\pm}$ radiative recombination dominates below $\approx 10^{6}$ K, and for $\mu$ and $\tau$, radiative recombination dominates at all astrophysically relevant temperatures.
\begin{figure}
\centering \includegraphics[scale=0.6]{rrdivda.eps}
\caption{The fraction of thermalised pairs which will either radiatively recombine (R.R.) or directly annihilate (D.A.), neglecting charge-exchange, as a function of temperature, for e$^{\pm}$ (continuous line), $\mu^{\pm}$ (dashed line), and $\tau^{\pm}$ (dotted line).}
\label{fig:rrda}
\end{figure}
\subsection{Cross-sections for radiative recombination onto the $nL$th level}
\label{sec:wallyn}
For those pairs which do radiatively recombine to form an onium, we require the fraction which recombine into the $nL$th level, such that we may then determine the subsequent decay.
\citet{wal96} give the cross-sections for the formation of Ps via radiative recombination as a function of the quantum levels $n$ and $L$ and the relative energy of of the electron and positron. We have repeated these calculations for Ps, M and T, and show the results in Figure~\ref{fig:sigman}. Thus we may calculate the fraction of oniums which form in the $nL$th energy level via radiative recombination as
\begin{equation}
\label{eqn:fsigmanL}
f_{\sigma_{nL}} = \frac{\sigma(nL)}{\sum\limits_{n=1}^{n=\infty} \sum\limits_{L=0}^{L=n-1} \sigma(nL)}.
\end{equation}
In practice we cannot sum to $n=\infty$, and instead choose a suitably high value of $n=n_{\rm lim}$, such that the fraction $f_{\sigma_{nL}}$ asymptotes to a constant value.
We computed $f_{\sigma_{nL}}$ for various limits, $n_{\rm lim}$, and found that $n_{\rm lim} = 300$, 500 or 1000 is sufficient for Ps, M, or T respectively. Using these limits we have calculated the fractions $f_{\sigma_{nL}}$ for Ps, M and T as a function of temperature, and give the results for the first six energy levels in Table~\ref{tab:sigmanfract}.
\begin{figure}
\subfigure[The total cross section for levels $nL$]{
\centering \includegraphics[scale=0.58]{crossplotnL.eps}
}
\subfigure[The total cross section for level $n$, summed over all the $L$ sub-levels.]{
\centering \includegraphics[scale=0.58]{crossplot.eps}
}
\caption{(a ) The cross section for the radiative recombination of Ps (black), M (red) and T (blue), onto the quantum levels $nL$ for $n=1$ to 3 as a function of the relative energy of the combining particles. The lines are coded (1,0) thick, (2,0) normal, (2,1) thin, (3,0) dashed, (3,1) dotted, (3,2) dot-dashed. (b) The total cross sections for level $n$, from $n=1$ to 10, summed over all the $L$ sub-levels.}
\label{fig:sigman}
\end{figure}
\begin{table*}
\caption{The fraction of those atoms which radiatively recombine to form Ps, M or T, which do so in the level $nL$.}
\label{tab:sigmanfract}
\begin{tabular}{ll|lllll|lllll|lllll}
&&\multicolumn{15}{c}{Temperature (K)} \\
&&\multicolumn{5}{c}{Ps}&\multicolumn{5}{c}{M}&\multicolumn{5}{c}{T} \\
n & L & $10^{3}$& $10^{4}$& $10^{5}$& $10^{6}$& $10^{7}$ & $10^{3}$& $10^{4}$& $10^{5}$& $10^{6}$& $10^{7}$ & $10^{3}$& $10^{4}$& $10^{5}$& $10^{6}$& $10^{7}$ \\ \hline
1 & 0 & 0.296 & 0.459 & 0.702 & 0.814 & 0.830 & 0.153 & 0.193
& 0.263 & 0.396 & 0.628 & 0.123 & 0.145 & 0.182 & 0.244 &
0.358 \\
2 & 0 & 0.043 & 0.068 & 0.100 & 0.105 & 0.104 & 0.022 & 0.028
& 0.039 & 0.059 & 0.093 & 0.018 & 0.021 & 0.027 & 0.036 &
0.053 \\
2 & 1 & 0.116 & 0.143 & 0.068 & 0.009 & 0.001 & 0.062 & 0.078
& 0.105 & 0.140 & 0.102 & 0.050 & 0.058 & 0.073 & 0.097 &
0.133 \\
3 & 0 & 0.015 & 0.023 & 0.031 & 0.031 & 0.031 & 0.008 & 0.010
& 0.013 & 0.020 & 0.029 & 0.006 & 0.007 & 0.009 & 0.012 &
0.018 \\
3 & 1 & 0.044 & 0.055 & 0.025 & 0.003 & 0.000 & 0.023 & 0.030
& 0.040 & 0.053 & 0.038 & 0.019 & 0.022 & 0.028 & 0.037 &
0.051 \\
3 & 2 & 0.048 & 0.037 & 0.003 & 0.000 & 0.000 &
0.027 & 0.034 & 0.045 & 0.046 & 0.010 & 0.022 & 0.026 &
0.032 & 0.042 & 0.049 \\
4 & 0 & 0.007 & 0.010 & 0.013 & 0.013 & 0.013 & 0.004 & 0.004
& 0.006 & 0.009 & 0.013 & 0.003 & 0.003 & 0.004 & 0.006 &
0.008 \\
4 & 1 & 0.021 & 0.026 & 0.011 & 0.001 & 0.000 & 0.011 & 0.014
& 0.019 & 0.025 & 0.017 & 0.009 & 0.011 & 0.013 & 0.018 &
0.024 \\
4 & 2 & 0.030 & 0.023 & 0.002 & 0.000 & 0.000 &
0.017 & 0.021 & 0.028 & 0.029 & 0.006 & 0.014 & 0.016 &
0.020 & 0.026 & 0.030 \\
4 & 3 & 0.020 & 0.007 & 0.000 & 0.000 &
0.000 & 0.013 & 0.016 & 0.020 & 0.013 &
0.001 & 0.010 & 0.012 & 0.015 & 0.019 & 0.017 \\
5 & 0 & 0.004 & 0.006 & 0.007 & 0.007 & 0.007 & 0.002 & 0.003
& 0.003 & 0.005 & 0.007 & 0.002 & 0.002 & 0.002 & 0.003 &
0.005 \\
5 & 1 & 0.012 & 0.014 & 0.006 & 0.001 & 0.000 & 0.006 & 0.008
& 0.011 & 0.014 & 0.009 & 0.005 & 0.006 & 0.007 & 0.010 &
0.013 \\
5 & 2 & 0.018 & 0.014 & 0.001 & 0.000 & 0.000 &
0.010 & 0.013 & 0.017 & 0.017 & 0.004 & 0.008 & 0.010 &
0.012 & 0.016 & 0.019 \\
5 & 3 & 0.017 & 0.007 & 0.000 & 0.000 &
0.000 & 0.011 & 0.014 & 0.017 & 0.012 &
0.001 & 0.009 & 0.011 & 0.013 & 0.017 & 0.015 \\
5 & 4 & 0.008 & 0.001 & 0.000 & 0.000 & 0.000 & 0.006 & 0.008 & 0.009 &
0.003 & 0.000 & 0.005 & 0.006 & 0.007 & 0.009 & 0.005 \\
6 & 0 & 0.002 & 0.003 & 0.004 & 0.004 & 0.004 & 0.001 & 0.002
& 0.002 & 0.003 & 0.004 & 0.001 & 0.001 & 0.001 & 0.002 &
0.003 \\
6 & 1 & 0.007 & 0.008 & 0.003 & 0.000 & 0.000 & 0.004 & 0.005
& 0.007 & 0.008 & 0.005 & 0.003 & 0.004 & 0.005 & 0.006 &
0.008 \\
6 & 2 & 0.012 & 0.009 & 0.001 & 0.000 & 0.000 &
0.007 & 0.008 & 0.011 & 0.011 & 0.002 & 0.005 & 0.006 &
0.008 & 0.010 & 0.012 \\
6 & 3 & 0.013 & 0.005 & 0.000 & 0.000 &
0.000 & 0.008 & 0.010 & 0.013 & 0.009 &
0.000 & 0.007 & 0.008 & 0.010 & 0.012 & 0.011 \\
6 & 4 & 0.009 & 0.001 & 0.000 & 0.000 &0.000 & 0.007 & 0.009 & 0.010 &
0.004 & 0.000 & 0.006 & 0.007 & 0.008 & 0.010 & 0.006 \\
6 & 5 & 0.003 & 0.000 & 0.000 & 0.000 & 0.000 & 0.003 & 0.004 & 0.004 &
0.001 & 0.000 & 0.002 & 0.003 & 0.004 & 0.004
& 0.001 \\
\end{tabular}
\end{table*}
\section{Decay channels}
\label{sec:decay}
We have now calculated the fraction of pairs produced via photon-photon (\S~\ref{sec:gg}), or electron-positron (\S~\ref{sec:ee}), annihilation which have sufficiently low energy to form Ps, M or T, and of those which do, the fraction which radiatively recombine or directly annihilate (\S~\ref{sec:gould}), and of those which radiatively recombine to form an onium, the fraction which form in the $nL$th level (\S~\ref{sec:wallyn}).
An onium in the $nL$th level may decay via annihilation of its constituent particles, if $L=0$; it may radiatively transition to a lower quantum level if $n >1$; or if it is M or T, either one of the individual constituent particles may decay from any level. We wish to calculate the branching fractions for these various decay channels, and we begin by calculating the lifetimes for each process.
\subsection{Annihilation}
\label{sec:decay_ann}
Invariance under charge-conjugation leads to the following selection rule for the decay of an onium into $n$ photons,
\begin{equation}
(-1)^{l+s} = (-1)^{n},
\end{equation}
where $l$ and $s$ are the orbital angular momentum and spin quantum numbers respectively\cite{yan50,wol52}. Furthermore, since the electron and positron (or $\mu^{-}$ and $\mu^{+}$ etc.) only overlap in the $L=0$ state, annihilation into photons is only possible (barring negligible higher order decay processes) from the singlet $^{1}S_{0}$ state or from the triplet $^{3}S_{1}$ state. The $^{1}S_{0}$ will decay into an even number of photons, with 2 being the most probable. The $^{3}S_{1}$ state will decay into an odd number of photons, with 3 being the most probable, since 1 is forbidden due to conservation of momentum. However, for M and T the triplet state can also decay via one photon, which then decays into e$^{\pm}$ pairs, i.e.\ M~$\rightarrow \gamma^{*} \rightarrow {\rm e}^{+}{\rm e}^{-}$.
The decay rate of an onium in the $n$th level of the singlet state is given by\cite{dir30b},
\begin{equation}
\Gamma_{1} = \frac{1}{n^{3}} \frac{\alpha^{5} \mu c^{2}}{\hbar},
\end{equation}
where $\mu$ is the reduced mass of the onium.
For triplet states the lowest order decay rate is \cite{ore49,berk80},
\begin{equation}
\Gamma_{3}=\frac{1}{n^{3}}\frac{2}{9\pi}(\pi^{2}-9)\left(\frac{2 \mu c^{2}}{\hbar}\right)\alpha^{6}.
\end{equation}
For M and T the decay of the triplet state into electron-positron pairs has a decay rate given by\cite{brod09},
\begin{equation}
\Gamma_{e^{+}e^{-}} = \frac{\alpha^{5} \mu c^{2}}{3 \hbar n^{3}}.
\end{equation}
\subsection{Decay}
\label{sec:mutaudecay}
Both the $\mu$ and the $\tau$ lepton are intrinsically unstable and will decay to lighter particles.
We take the decay rates to be, $T_{\mu} =1/\Gamma_{\mu} = 2.197 \times 10^{-6}$~s and $T_{\tau} = 1/\Gamma_{\tau}=2.874 \times 10^{-13}$~s \cite{per92}.
\subsection{Radiation}
\label{sec:decay_rad}
The coefficient for a radiative transition of a hydrogenic atom from the state $n' L'$ to a lower state $nL$, $A_{n'L',nL}$ is given by\cite{gre57,pen64,bro71}
\begin{equation}
A_{n'L',nL} = 4 \pi^{2} \nu^{3} \left(\frac{ 8 \pi \alpha a_{0}^{2}}{3 c^{2}}\right) \frac{{\rm Max}(L,L')}{(2L'+1)} \left|p(n'L',nL)\right|^{2},
\end{equation}
where $\nu$ is the frequency of the transition,
\begin{equation}
\nu = c R_{\infty} \left(\frac{1}{n^{2}} - \frac{1}{n'^{2}}\right),
\end{equation}
$R_{\infty}$ is the Rydberg constant for the particular onium,
\begin{equation}
R_{\infty}=\frac{(\mu e^{4})}{8 \epsilon_{0}^{2} c h^{3}}
\end{equation}
and $a_{0}$ is the Bohr radius of the onium,
\begin{equation}
a_{0}=\frac{4 \pi \epsilon_{0} \hbar^{2}}{\mu e^{2}}.
\end{equation}
The dipole matrix elements are given by\cite{gor29},
\begin{widetext}
\begin{eqnarray}
|p(n'L-1,nL|^{2} &=& \Biggl( \frac{(-1)^{n'-1}}{4(2L-1)!} \sqrt{\frac{(n+L)!(n'+L-1)!}{(n-L-1)!(n'-L)!}}
\frac{(4nn')^{L+1}}{(n+n')^{n+n'}} (n-n')^{n+n'-2L-2} \Biggr. \nonumber \\
&& \times \biggl( \ _{2}F_{1}\left[-n+L+1,-n'+L,2L,\frac{-4nn'}{(n-n')^{2}}\right] \biggr. \nonumber \\
&& -\left(\frac{n-n'}{n+n'}\right)^{2} \ _{2}F_{1} \left[-n+L-1,-n'+L,2L,\frac{-4 nn'}{(n-n')^{2}}\right] \biggl. \biggr) \Biggl.\Biggr)^{2},
\end{eqnarray}
\end{widetext}
which is correct when $L>L'$; if $L<L'$ then $nL$ and $n'L'$ are swapped. These can be adapted for any onium simply by replacing the reduced mass, $\mu$\cite{wal96}.
The radiative coefficient for a transition from a level $n'L'$ to \emph{any} lower level is thus given by,
\begin{equation}
A_{n'L'} = \sum\limits_{n=1}^{n=n'-1} \sum\limits_{L=L' \pm 1} A_{n'L',nL}.
\end{equation}
\section{Branching ratios}
\label{sec:branch}
\subsection{Branching ratios for a particular level, $nL$}
Consider leptonium in the level $nL$. It may decay via annihilation (two photon, three photon or $e^{\pm}$, depending on the state and the type of atom), radiative transition, or the decay of either constituent particle (for M and T). The decay rates for these channels were calculated in the previous section. Thus we may now calculate the probabilities for each process.
Recalling that annihilation can only take place from the states with $L=0$, and also that the annihilation mechanisms differ for singlet states, which decay into two photons, and triplet states, which decay either into three photons, or into $e^{\pm}$ pairs via a single photon, the total decay constant for leptonium in a level $nL$ via \emph{any} decay process is given by,
\begin{eqnarray}
A_{nL_{\rm sing}} = &A_{nL} +\delta(L)\Gamma_{1} + 2 \Gamma_{\mu,\tau},& ^{1}S_{0}\label{eqn:AnL1}\\
A_{nL_{\rm trip}} = &A_{nL} +\delta(L)\Gamma_{3}+\delta(L)\Gamma_{ee} + 2 \Gamma_{\mu,\tau},& ^{3}S_{1},\label{eqn:AnL3}
\end{eqnarray}
%
where $\delta(L)$ is a function which is 1 when $L=0$, and zero otherwise.
Therefore the probability of making a particular radiative transition from a level $n'L'$ to a lower level $nL$ is given by,
\begin{eqnarray}
P_{n'L',nL_{\rm sing}} &= &\frac{1}{4}\frac{A_{n'L',nL}}{A_{n'L'_{\rm sing}}} \\
P_{n'L',nL_{\rm trip}} &= & \frac{3}{4}\frac{A_{n'L',nL}}{A_{n'L'_{\rm trip}}}, \\
P_{n'L',nL_{\rm tot}} &= &P_{n'L',nL_{\rm sing}} + P_{n'L',nL_{\rm trip}}
\end{eqnarray}
where the leading factors of $1/4$ and $3/4$ are the probabilities that the onium will form in the singlet or triplet state, respectively.
Similarly the probability for leptonium in the level $n'L'$ to decay via two-photon annihilation is,
\begin{equation}
P_{n'L'_{\gamma\gamma}} = \frac{1}{4}\frac{\delta(L')\Gamma_{1}}{A_{n'L'_{\rm sing}}}.
\end{equation}
Likewise for three photon annihilation,
\begin{equation}
P_{n'L'_{\gamma\gamma\gamma}} = \frac{3}{4}\frac{\delta(L')\Gamma_{3}}{A_{n'L'_{\rm trip}}},
\end{equation}
and annihilation into an electron-positron pair
\begin{equation}
P_{n'L'_{\rm ee}} = \frac{3}{4}\frac{\delta(L')\Gamma_{ee}}{A_{n'L'_{\rm trip}}}.
\end{equation}
Finally the probability that either constituent particle will decay is
\begin{equation}
P_{n'L'_{\rm decay}} = 2\left(\frac{\Gamma_{\mu,\tau}}{4}\frac{1}{ A_{n'L'_{\rm sing}}} + \frac{3}{4}\frac{\Gamma_{\mu,\tau} }{A_{n'L'_{\rm trip}}}\right)
\end{equation}
where $\Gamma_{\mu,\tau}$ is the decay rate of the $\mu$ or $\tau$ as appropriate.
\subsection{Total branching ratios}
The previous subsection gave the branching ratios for the various decays from a particular state $nL$. We now need to consider the probability of leptonium being in the state $nL$ in the first place. There are two possibilities for populating the state: it may recombine directly into the state $nL$, or it may recombine into a higher state and cascade down into $nL$. That is, we assume Case A recombination\cite{ost06}; because leptonium is short-lived, there is no resonant scattering of emission lines from Ps, M or T, through the excitation and re-emission of surrounding Ps, M or T atoms. For the same reason, we also assume that collisional excitation of the atoms is negligible.
\subsubsection{Cascade probabilities}
The probability of cascading from a level $n'L'$ to a lower level $nL$ via \emph{all} cascade paths, $C_{n'L',nL}$, is calculated using an iterative procedure\cite{pen64}, such that,
\begin{widetext}
\begin{equation}
C_{n'L',nL_{\rm sing,trip}} =
\begin{cases}
1,& n=n' \wedge L=L' \\
\sum\limits_{n''=n}^{n'-1} \sum\limits_{L''=L' \pm 1} P_{n'L',n''L''_{\rm sing,trip}} C_{n''L'',nL_{\rm sing,trip}},& n\neq n' \vee L\neq L' ,
\end{cases}
\end{equation}
\end{widetext}
where the subscripts `sing' and `trip' are used to specify the calculations for the singlet state or triplet state, an unfortunate but necessary complication since the decay paths are different for each.
\subsubsection{Population of the state, $nL$}
Meanwhile the state is depopulated by the processes described in section~\ref{sec:decay}. Therefore, in equilibrium the population of any state, $N_{nL}$ is given by\cite{sea59,pen64},
\begin{eqnarray}
\sum\limits_{n'=n}^{\infty}\sum\limits_{L'=0}^{n'-1} N_{\rm pairs} f_{\rm ion} g_{\rm rr} f_{\sigma_{nL}} C_{n'L',nL_{\rm sing,trip}} \nonumber \\
= N_{nL_{\rm sing,trip}} A_{nL_{\rm sing,trip}},
\end{eqnarray}
where $N_{\rm pairs}$ is the rate at which pairs are produced,
$f_{\rm ion}$ is the fraction of pairs produced with energy less than the ionization energy, i.e.\ $f_{\gamma \gamma}$ or $f_{ee}$ in sections~\ref{sec:gg} and \ref{sec:ee} above,
$g_{\rm rr}$ is the fraction of those pairs which radiatively recombine (\S~\ref{sec:gould}, Figure~\ref{fig:rrda}),
$f_{\sigma_{nL}}$ is the fraction of those pairs which form in level $nL$, (\S~\ref{sec:wallyn}, equation~\ref{eqn:fsigmanL}),
and again the equation has two forms reflecting the different decay paths of the singlet and triplet states.
Let us then define $f_{nL}$ as the relative population of the level $nL$, i.e.,
\begin{eqnarray}
f_{nL_{\rm sing,trip}}&=&\frac{N_{nL_{\rm sing,trip}}}{N_{\rm pairs} f_{\rm ion} g_{\rm rr} }\nonumber \\
& =& \frac{\sum\limits_{n'=n}^{\infty} \sum\limits_{L'=0}^{n'-1}f_{\sigma_{n'L'}} C_{n'L',nL_{\rm sing,trip}} }{A_{nL_{\rm sing,trip}}}, \label{eqn:fnL}
\end{eqnarray}
which is the fraction of leptonium in the level $nL$ for a population whose formation and decay rates are in equilibrium.
\subsubsection{Branching ratios}
The final step in calculating the total branching ratios is simply to multiply the fractional population each state, $f_{nL}$, by the appropriate decay constant, and sum over all levels. Thus,
\begin{widetext}
\begin{align}
P_{\rm Ly\alpha} &=&& A_{21,10}\left(\frac{1}{4} f_{21_{\rm sing}} + \frac{3}{4} f_{21_{\rm trip}}\right), & {\rm Lyman}\ \alpha \label{eqn:pla}\\
P_{\rm Ly\beta} &= && A_{31,10}\left(\frac{1}{4} f_{31_{\rm sing}} + \frac{3}{4} f_{31_{\rm trip}}\right), &{\rm Lyman}\ \beta \\
P_{\rm Ba\alpha} &=&& A_{31,20}\left(\frac{1}{4} f_{31_{\rm sing}} + \frac{3}{4} f_{31_{\rm trip}}\right) + A_{30,21}\left(\frac{1}{4} f_{30_{\rm sing}} + \frac{3}{4} f_{30_{\rm trip}}\right), &{\rm Balmer}\ \alpha \\
P_{\rm Ba\beta} &=&& A_{41,20}\left(\frac{1}{4} f_{41_{\rm sing}} + \frac{3}{4} f_{41_{\rm trip}}\right) + A_{40,21}\left(\frac{1}{4} f_{40_{\rm sing}} + \frac{3}{4} f_{40_{\rm trip}}\right), & {\rm Balmer}\ \beta \\
P_{\gamma\gamma} &=&&\sum\limits_{n=0}^{\infty} \frac{1}{4}f_{n0_{\rm sing}} \Gamma_{1}(n), &{\rm Two\ photon\ annihilation} \label{eqn:pgamgam}\\
P_{\gamma\gamma\gamma} &=&&\sum\limits_{n=0}^{\infty} \frac{3}{4}f_{n0_{\rm trip}} \Gamma_{3}(n), &{\rm Three\ photon\ annihilation} \\
P_{\rm ee} &=&&\sum\limits_{n=0}^{\infty} \frac{3}{4}f_{n0_{\rm trip}} \Gamma_{ee}(n), &{\rm Electron-positron\ annihilation} \label{eqn:pelpos}\\
P_{\rm decay} &=&& \sum\limits_{n=0}^{\infty}\sum\limits_{L=0}^{n-1} 2 \Gamma_{\mu,\tau} \left(\frac{1}{4} f_{nL_{\rm sing}} + \frac{3}{4} f_{nL_{\rm trip}}\right),& {\rm Decay\ of\ either\ particle} \label{eqn:pdec}
\end{align}
\end{widetext}
All of the equations~\ref{eqn:pla} -- \ref{eqn:pdec}, contain either one or two infinite sums. All include the infinite sum in equation~\ref{eqn:fnL} which accounts for recombination into any level up $n=\infty$ followed by cascade to the relevant level. Equations~\ref{eqn:pgamgam} -- \ref{eqn:pdec} also include the sum over all levels up to $n=\infty$ which all contribute to the relevant decay process. In practice we cannot compute these infinite sums, and instead compute equations~\ref{eqn:pla} -- \ref{eqn:pdec} up to certain limits on each sum. We then alter these limits, and fit to the resulting curve or surface, and extrapolate our results to $n=\infty$. The limits are $n=11$ -- $15$, for the capture-cascade sum, and $n=6$ -- $10$ for the sum over the decays.
For example, Figure~\ref{fig:MLyafit} shows the fit to the branching ratio for M emitting Lyman $\alpha$ radiation at a gas temperature of $T=10^{4}$~K. We find a good fit to all curves using a function of the form
\begin{equation}
\label{eqn:ffit1}
b = a + c {\rm e}^{- \frac{n}{d}},
\end{equation}
where $a$, $c$, and $d$ are all parameters to be fit.
Figure~\ref{fig:eefit} shows the two dimensional fit over both limits for annihilation into electron-positron pairs for M at $T=10^{4}$~K. Again, we find a good fit with a functional form
\begin{equation}
\label{eqn:ffit2}
b = a + c {\rm e}^{- \frac{n_{1}}{d}}+ f {\rm e}^{- \frac{n_{2}}{g}},
\end{equation}
where where $a$, $c$, $d$, $f$ and $g$ are all parameters to be fit, and the $n_{1}$ and $n_{2}$ are the different limiting values of $n$ in the sums.
\begin{figure}
\centering \includegraphics[scale=0.58]{lyafit.eps}
\caption{The branching ratio of Ly $\alpha$ emission from M at a gas temperature of $T=10^{4}$~K as a function of the limiting value of the sum over $n$ in equation~\ref{eqn:fnL}, shown by the points. We find good fits with a function of the form given in equation~\ref{eqn:ffit1}, shown by the curve; in this case giving a limiting value as $n \rightarrow \infty$ of 0.45.}
\label{fig:MLyafit}
\end{figure}
\begin{figure}
\centering \includegraphics[scale=0.31]{eefit.eps}
\caption{The branching ratio of decay into an electron-positron pair from M at a gas temperature of $T=10^{4}$~K as a function of the limiting value of the sums over $n$ in equations~\ref{eqn:fnL} and \ref{eqn:pelpos}, shown by the points. We find good fits with a function of the form given in equation~\ref{eqn:ffit2}, shown by the curve; in this case giving a limiting value as $n \rightarrow \infty$ of 0.59.}
\label{fig:eefit}
\end{figure}
We have checked the accuracy of the extrapolations in the following way. The branching ratios for the two photon or three photon decay of Ps, should be 0.25 and 0.75 by definition, since there is no other way for Ps to decay. We find the values given in Table~\ref{tab:Psbranch}, which agree to within 2 per cent in all cases.
We wish to highlight one further caveat to these branching ratios. The extrapolations for photon decays, and electron-positron decays are not large, and become less important as $n$ increases, since they can only occur when $L=0$, which becomes less common for higher $n$. However, the decay of either particle can occur from any $L$ level, and the extrapolations are large and therefore uncertain. Therefore, rather than rely on the extrapolations for decay, we set the probability to be one minus the sum of the other decays. In any case, the decay of either particle does not have a clear observational signature and it is the probabilities for the other processes which are most important for this study. This uncertainty in our results could be lessened through calculations to higher levels, but would require significant computing resources which are unavailable to us at this time.
The results for the branching ratios are given in Tables~\ref{tab:Psbranch},~\ref{tab:Mbranch} and \ref{tab:Tbranch}. Since Ps is well studied and understood, we do not discuss it further\cite{wal96,pra11}.
M has significant probabilities of Lyman $\alpha$ and Balmer $\alpha$ radiation before decaying. The decay processes are dominated by two-photon annihilation for para-M and electron-positron annihilation for ortho-M, as well as significant probabilities for either $\mu^{\pm}$ to decay if $T < 10^{5}$~K.
The much shorter lifetimes of $\tau$ means that the decay processes of T are dominated by the decay of either $\tau^{\pm}$, with only a small probability of either electron-positron or two photon annihilation. The probabilities of emitting Lyman $\alpha$ radiation before decay are 1 -- 2 per cent, and $\approx 0$ for other emission lines.
\begin{table*}
\caption{Total branching ratios for Ps as a function of temperature.}
\label{tab:Psbranch}
\begin{tabular}{lllllll}
&&&&& Two & Three \\
&&&&&photon&photon \\
T (K) & Ly$\alpha$ & Ly$\beta$ & Balmer $\alpha$ & Balmer $\beta$ &decay&decay\\\hline
1000 & 0.46 & 0.08 & 0.24 & 0.06 & 0.24 & 0.73 \\
10000 & 0.29 & 0.07 & 0.09 & 0.03 & 0.25 & 0.74 \\
100000 & 0.11 & 0.03 & 0.03 & 0.01 & 0.25 & 0.75 \\
1000000 & 0.04 & 0.01 & 0.02 & 0.01 & 0.25 & 0.75 \\
10000000 & 0.03 & 0.01 & 0.02 & 0.01 & 0.25 & 0.75 \\
\end{tabular}
\end{table*}
\begin{table*}
\caption{Total branching ratios for M as a function of temperature.}
\label{tab:Mbranch}
\begin{tabular}{lllllllll}
&&&&&Two&Three&&\\
&&&&&photon&photon&Electron-positron&Decay of\\
T (K) & Ly$\alpha$ & Ly$\beta$ & Balmer $\alpha$ & Balmer $\beta$ & decay & decay &decay &either particle\\ \hline
1000 & 0.37 & 0.04 & 0.27 & 0.04 & 0.16 & 0.00 & 0.48 & 0.36\\
10000 & 0.45 & 0.05 & 0.31 & 0.05 & 0.20 & 0.00 & 0.59 & 0.21\\
100000 & 0.46 & 0.07 & 0.26 & 0.06 & 0.23 & 0.00 & 0.70 & 0.07 \\
1000000 & 0.32 & 0.07 & 0.12 & 0.04 & 0.25 & 0.00 & 0.74 & 0.01 \\
10000000 & 0.13 & 0.04 & 0.02 & 0.01 & 0.25 & 0.00 & 0.74 & 0.00 \\
\end{tabular}
\end{table*}
\begin{table*}
\caption{Total branching ratios for T as a function of temperature.}
\label{tab:Tbranch}
\begin{tabular}{lllllllll}
&&&&&Two&Three&&\\
&&&&&photon&photon&Electron-positron&Decay of\\
T (K) & Ly$\alpha$ & Ly$\beta$ & Balmer $\alpha$ & Balmer $\beta$ & decay & decay &decay &either particle\\ \hline
1000 & 0.01 & 0.00 & 0.00 & 0.00 & 0.03 & 0.00 & 0.06 & 0.91 \\
10000 & 0.01 & 0.00 & 0.00 & 0.00 & 0.03 & 0.00 & 0.07 & 0.90 \\
100000 & 0.01 & 0.00 & 0.00 & 0.00 & 0.04 & 0.00 & 0.09 & 0.87 \\
1000000 & 0.01 & 0.00 & 0.00 & 0.00 & 0.06 & 0.00 & 0.12 & 0.83 \\
10000000 & 0.02 & 0.00 & 0.00 & 0.00 & 0.08 & 0.00 & 0.17 & 0.75 \\
\end{tabular}
\end{table*}
\section{Observable signatures}
\label{sec:signatures}
Tables~\ref{tab:Mbranch} and~\ref{tab:Tbranch} give the branching ratios for possible radiative recombination lines and decays of M and T. We now discuss the observational signatures of these processes.
The energies of radiative recombination lines and of 2-photon annihilation signatures and ionization are given in Table~\ref{tab:summ}. In addition to the 2-photon annihilation signature there will be a continuum of radiation from the 3-photon decay of the singlet state, with energies from zero to the energy of the 2-photon decay. The decay mechanisms for both $\tau$ and $\mu$ are numerous and there is no one clear observable signature that we can associate with either of these, nor is there any single identifiable signature of annihilation into e$^{\pm}$.
All these signatures are observable with current instrumentation if they are sufficiently bright. For example Fermi-LAT is sensitive to $\gamma$-rays between 30 MeV to 300GeV and could observe the annihilation of M and T;
INTEGRAL IBIS is sensitive to the T Lyman lines;
XMM-Newton is sensitive to X-rays between 0.1 and 15 keV, and can observe the recombination lines of M, and the T Balmer lines.
\section{Expected signatures from astrophysical sources}
\label{sec:astrophys}
We are finally in a position to estimate the expected observable signatures of M and T from astrophysical sources.
Consider a source of particle anti-particle pairs being produced at a rate $r$ per second.
The fraction of these pairs which will produce M or T is given by
\begin{equation}
f_{\rm onium} = r f_{\rm ion} g_{\rm rr}
\end{equation}
where
$f_{\rm ion}$ is the fraction of pairs produced with energy less than the ionization energy, i.e.\ $f_{\gamma \gamma}$ or $f_{ee}$ in sections~\ref{sec:gg} and \ref{sec:ee} above, and
$g_{\rm rr}$ is the fraction of those pairs which radiatively recombine (\S~\ref{sec:gould}, Figure~\ref{fig:rrda}). Therefore the luminosity of any particular observable signature in ph~s$^{-1}$ is
\begin{equation}
\label{eqn:lum}
L = p\ f_{\rm onium}\ b
\end{equation}
where $b$ is the branching ratio of the particular signature (Tables~\ref{tab:Mbranch} and \ref{tab:Tbranch}), and $p$ is a factor which specifies the number of photons produced by the process, i.e.\ $p=1$ for recombination lines, $p=2$ for two-photon annihilation and $p=3$ for three-photon annihilation. These luminosities can be easily converted to units of erg~s$^{-1}$ using the energies of the emitted photons given in Table~\ref{tab:summ}.
We will now estimate the expected fluxes from particular sources. These should be compared to the limiting sensitivities of current instruments. The M recombination lines are all in the soft X-ray band (0.5 -- 2 keV), for which a 100 ksec observation has a $\approx 4\sigma$ limiting sensitivity\cite{has01} of $f_{\rm X}\sim 3.1 \times 10^{-16}$ erg cm$^{-2}$ s$^{-1}$. The T recombination Balmer lines are in the hard X-ray band (2 -- 10 keV), for which a 100 ksec observation has a $\approx 4\sigma$ limiting sensitivity of $f_{\rm X}\sim 1.4 \times 10^{-15}$ erg cm$^{-2}$ s$^{-1}$. The T Lyman lines are at the low energy limit of the SPI spectrometer on board INTEGRAL. The limiting 4$\sigma$ sensitivity for an exposure time of $10^{5}$ s is $\approx 3 \times 10^{-13}$ erg cm$^{-2}$ s$^{-1}$, using data from the INTEGRAL webpages. The M and T annihilation lines are detectable by FERMI-LAT, for which the limiting $4\sigma$ sensitivity for an exposure time of $10^{5}$ s is $\approx 4 \times 10^{-13}$ erg cm$^{-2}$ s$^{-1}$ for M and $\approx 1 \times 10^{-13}$ erg cm$^{-2}$ s$^{-1}$ for T, using data from the FERMI webpages.
\subsection{Jets}
The powerful relativistic jets of active galactic nuclei (AGN) and microquasars are a probable source of pair production. These jets have strong radio emission with a power-law spectrum which is interpreted as being due to synchrotron emission from relativistic $e^{-}$ gyrating around the magnetic field lines of the AGN. The jets must be electrically neutral, otherwise they would cause a potential difference to build-up, which would oppose and eventually stop the jet.
It is unknown whether the positive component of jets consists of positrons, protons, or a mixture of both. Observational studies have had to rely on indirect methods of searching for the presence of \e+, such as estimates of the bulk kinetic energy contained in jets, which have been used to argue for both $e^{-}$-p plasma\cite{cel93} and $e^{-}$-\e+\ plasma\cite{rey96,war98,hir05}.
Theoretically there are good reasons to expect that jets contain some fraction of \e+. A pair plasma has the advantage of explaining $\gamma$-ray jets\cite{bla95} and the very high Lorentz factors ($\Gamma > 5$) required to account for superluminal bulk velocities of jets\cite{beg84}.
It is possible that some Ps may form in the jet, or similarly, that some e$^{\pm}$ pairs may collide with the required energy to form M and T as outlined in \S~\ref{sec:ee}. However, in the observers frame, the velocity of the oniums thus formed will be highly relativistic, causing relativistic Doppler broadening and beaming of any emitted radiation\cite{boe96,mar07}. Such a broadened signal may still be possible to observe\cite{boe96}, but would be difficult to interpret, and hardly constitutes an unambiguous test of the presence of exotic onium atoms such as M or T.
Therefore we do not consider onium formation in the jet itself as a likely candidate for detection.
However, if a jet collides with stationary object such as a gas cloud, or a star, then the leptonium formed in the collision will give rise to emission that is in principle observable. Such collisions may occur in the radio jets of AGN\cite{gom00}, or if the jets of microquasars are mis-aligned and hit the secondary companion\cite{gue06}.
Let us consider some illustrative examples. Previously\cite{ell09}, we have calculated the expected positron contents of jets, by applying the arguments of \citet{mar07} and \citet{mar83} , which were developed in order to search for 511 keV Ps annihilation radiation from 3C 120, to the empirical measurements of \citet{ghi93}.
We found a maximum of $\approx 10^{49}$~\e+~s$^{-1}$ are produced in the jets of blazars, whereas quasars produce $\approx 10^{46}$ -- $10^{48}$~\e+~s$^{-1}$ in their jets. The jets of microquasars are expected to have a positron flux of $\approx 10^{41}$~\e+~s$^{-1}$\cite{gue06}.
We assume that the object being hit by the jet is dense enough to stop \emph{all} these positrons. Let us assume a spectral index of $\alpha = 1.5$ and a temperature of $T=10^{6}$~K for the positrons in the jet. Thus the intrinsic luminosities of the various observational signatures can be calculated, and are given in Table~\ref{tab:blazar} for a blazar. These are then used to calculate the flux of the signal as a function of redshift, as shown in Figure~\ref{fig:blazar}. Similar results are shown in Table~\ref{tab:mqso} and Figure~\ref{fig:mqso}, for a mis-aligned microquasar with a spectral index of $\alpha = 1.5$ and a temperature of $T=10^{5}$~K for the positrons in the jet.
\begin{table}
\caption{The intrinsic luminosities for a blazar jet which produces $10^{49}$~\e+~s$^{-1}$, with spectral index $ \alpha=1.5$ and $T=10^{6}$~K.}
\label{tab:blazar}
\begin{tabular}{lll}
& \multicolumn{2}{c}{Luminosity (ph s$^{-1}$) }\\
& M & T \\ \hline
Ly $\alpha$ &$2.7\times 10^{42}$ &$1.2\times 10^{41}$\\
Ly $\beta$ & $6.0\times 10^{41}$&$1.3\times 10^{40}$\\
Ba $\alpha$ & $9.8\times 10^{41}$&$7.7\times 10^{39}$\\
Ba $\beta$ & $3.3\times 10^{41}$&$1.6\times 10^{39}$\\
Two photon & $4.1\times 10^{42}$&$9.6\times 10^{41}$
\end{tabular}
\end{table}
\begin{figure}
\begin{center}
\subfigure[M]{
\includegraphics[scale=0.6]{Mbllumplot.eps}
}
\subfigure[T]{
\includegraphics[scale=0.6]{Tbllumplot.eps}
}
\caption{The flux as a function of redshift for a blazar jet which produces $10^{49}$~\e+~s$^{-1}$, with spectral index $\alpha=1.5$ and $T=10^{6}$~K, colliding with a gas cloud. The lines show two-photon annihilation (red), Ly $\alpha$ (black) Ly $\beta$ (dashed, black), Ba $\alpha$ (blue) and Ba $\beta$ (dashed, blue).}
\label{fig:blazar}
\end{center}\end{figure}
\begin{table}
\caption{The intrinsic luminosities for a microquasar jet which produces $10^{41}$~\e+~s$^{-1}$, with spectral index $\alpha=1.5$ and $T=10^{5}$~K.}
\label{tab:mqso}
\begin{tabular}{lll}
& \multicolumn{2}{c}{Luminosity (ph s$^{-1}$) }\\
& M & T \\ \hline
Ly $\alpha$ &$3.9\times 10^{34}$ &$8.7\times 10^{32}$\\
Ly $\beta$ & $5.8\times 10^{33}$&$9.7\times 10^{31}$\\
Ba $\alpha$ & $2.3\times 10^{34}$&$5.9\times 10^{31}$\\
Ba $\beta$ & $4.8\times 10^{33}$&$1.2\times 10^{32}$\\
Two photon & $4.0\times 10^{34}$&$7.2\times 10^{33}$
\end{tabular}
\end{table}
\begin{figure}
\begin{center}
\subfigure[M]{
\includegraphics[scale=0.6]{Mmqsolumplot.eps}
}
\subfigure[T]{
\includegraphics[scale=0.6]{Tmqsolumplot.eps}
}
\caption{The flux as a function of redshift for a microquasar jet which produces $10^{41}$~\e+~s$^{-1}$, with spectral index$=-1.5$ and $T=10^{5}$~K, colliding with a gas cloud. The lines show two-photon annihilation (red), Ly $\alpha$ (black) Ly $\beta$ (dashed, black), Ba $\alpha$ (blue) and Ba $\beta$ (dashed, blue).}
\label{fig:mqso}
\end{center}\end{figure}
For our illustrative examples, no lines from blazars are above the relevant current 4$\sigma$ detection limits for a 100~ks exposure for either M or T. However, a microquasar jet which produces $10^{41}$~\e+~s$^{-1}$, with spectral index$=-1.5$ and $T=10^{5}$~K, colliding with a gas cloud, would be detectable in principle to the distances given in Table~\ref{tab:mqsojet}. The recombination lines would only be detectable at very close distances, but the two photon annihilation should be detectable at 4$\sigma$ in a 100~ksec exposure to $\approx 0.4$~kpc and $\approx 1$~kpc for M and T respectively.
\begin{table}
\caption{The distance to which a microquasar jet which produces $10^{41}$~\e+~s$^{-1}$, with spectral index$=-1.5$ and $T=10^{5}$~K, colliding with a gas cloud, would be detectable at 4 $\sigma$ in 100 ksec.}
\label{tab:mqsojet}
\begin{tabular}{lcc}
& \multicolumn{2}{c}{Distance (kpc)} \\
&M&T\\ \hline
Ly $\alpha$ &0.04 & $9 \times 10^{-4}$\\
Ly $\beta$ & 0.01 & $3 \times 10^{-4}$\\
Ba $\alpha$ & 0.01 & 0.001\\
Ba $\beta$ &0.007 &$ 7 \times 10^{-4}$\\
Two photon & 0.40 & 1.13
\end{tabular}
\end{table}
\subsection{Accretion discs}
We now consider pairs produced in the accretion disc itself through photon-photon annihilation. The number density of pairs thus produced in the optically thick region of the disc is
\begin{equation}
N_{\gamma\gamma} \sim \frac{1}{\sigma_{\rm T} R},
\end{equation}
where $\sigma_{\rm T}$ is the Thompson cross-section for $\mu$ or $\tau$ and $R$ is the radius of the optically thick disc\cite{belo99}. If we assume $R \approx 2GM/c^{2}$, i.e.\ the disc has approximately the Schwarzschild radius, and the thickness of the disc is $\approx R$, then taking masses of $M_{\rm AGN} = 10^{6}$ M$_{\odot}$ for the mass of a black hole in an AGN, and $M_{\mu{\rm QSO}} = 10$ M$_{\odot}$ for the mass of a black hole in a microquasar, then the pair yields are as given in Table~\ref{tab:accdisc}.
\begin{table}
\caption{The number of pairs produced through photon-photon annihilation in an accretion disc.}
\label{tab:accdisc}
\begin{tabular}{lll}
& AGN & Microquasar \\ \hline
$\mu$ & $1.7 \times 10^{52}$ & $1.7 \times 10^{42}$ \\
$\tau$ & $5.0 \times 10^{54}$ & $5.0 \times 10^{44}$ \\
\end{tabular}
\end{table}
Now, assuming a spectral index of $\alpha = 1.5$, then we may estimate the branching ratios as in equation~\ref{eqn:lum}, whereupon we find the luminosities in ph~s$^{-1}$, as listed in Table~\ref{tab:accdisclum}. The flux of the AGN as a function of redshift, and the flux of the microquasar as a function of distance, are shown in Figures~\ref{fig:accdiscagn} and Figures~\ref{fig:accdiscmqso}.
\begin{table}
\caption{The intrinsic luminosities for pair production via photon annihilation in accretion discs surrounding a microquasar and an AGN, assuming solar masses of $10$ and $10^{6}$ M$_{\odot}$ respectively, temperatures of $10^{5}$~K and $10^{6}$~K, and a spectral index of $\alpha=1.5$.}
\label{tab:accdisclum}
\begin{tabular}{lllll}
& \multicolumn{2}{c}{Microquasar} & \multicolumn{2}{c}{AGN} \\
& \multicolumn{2}{c}{Luminosity (ph s$^{-1}$) } & \multicolumn{2}{c}{Luminosity (ph s$^{-1}$) } \\
& M & T &M&T\\ \hline
Ly $\alpha$ & $2.9 \times 10^{35}$ &$1.8 \times 10^{36}$ & $2.0 \times 10^{45}$ & $2.4 \times 10^{46}$\\
Ly $\beta$ & $4.2 \times 10^{34}$ &$2.0 \times 10^{35}$ &$4.4 \times 10^{44}$ & $2.7 \times 10^{45}$\\
Ba $\alpha$ & $1.7 \times 10^{35}$ & $1.2 \times 10^{35}$ & $7.2 \times 10^{44}$ & $1.6 \times 10^{45}$ \\
Ba $\beta$ & $3.6 \times 10^{34}$ &$2.5 \times 10^{34}$ & $2.4 \times 10^{44}$ & $3.2 \times 10^{44}$\\
Two photon & $2.9 \times 10^{35}$ & $1.5 \times 10^{37}$ & $3.0 \times 10^{45}$ & $2.0 \times 10^{47}$
\end{tabular}
\end{table}
\begin{figure}
\begin{center}
\subfigure[M]{
\includegraphics[scale=0.6]{Magnlumplot_disc.eps}
}
\subfigure[T]{
\includegraphics[scale=0.6]{Tagnlumplot_disc.eps}
}
\caption{The flux as a function of redshift for an accretion disc around an AGN with a $10^{6}$~M$_{\odot}$ black hole, and spectral index $\alpha=1.5$ and $T=10^{6}$~K. The lines show two-photon annihilation (red), Ly $\alpha$ (black) Ly $\beta$ (dashed, black), Ba $\alpha$ (blue) and Ba $\beta$ (dashed, blue).}
\label{fig:accdiscagn}
\end{center}\end{figure}
\begin{figure}
\begin{center}
\subfigure[M]{
\includegraphics[scale=0.6]{Mmqsolumplot_disc.eps}
}
\subfigure[T]{
\includegraphics[scale=0.6]{Tmqsolumplot_disc.eps}
}
\caption{The flux as a function of redshift for an accretion disc around a microquasar with a $10$~M$_{\odot}$ black hole, and spectral index $\alpha=1.5$ and $T=10^{5}$~K. The lines show two-photon annihilation (red), Ly $\alpha$ (black) Ly $\beta$ (dashed, black), Ba $\alpha$ (blue) and Ba $\beta$ (dashed, blue).}
\label{fig:accdiscmqso}
\end{center}\end{figure}
For these illustrative examples, and a 4$\sigma$ detection limit for a 100~ks exposure, the two-photon annihilation line of M would be observable in the accretion disc of an AGN at $z<0.025$, while the two-photon annihilation line of T would be observable out to $z<0.9$. Neither the M nor the T recombination lines would be detectable.
Meanwhile, for microquasars the M and T lines would be detectable at 4$\sigma$ in a 100 ksec exposure to the distances given in Table~\ref{tab:mqsodisc}. The recombination lines are detectable at close distances while the two photon annihilation should be detectable to 1.1 or 51~kpc for M and T respectively.
\begin{table}
\caption{The distance to which an accretion disc around a microquasar with a $10$~M$_{\odot}$ black hole, and spectral index $\alpha=1.5$ and $T=10^{5}$~K, would be detectable at 4 $\sigma$ in 100 ksec.}
\label{tab:mqsodisc}
\begin{tabular}{lcc}
& \multicolumn{2}{c}{Distance (kpc)} \\
&M&T\\ \hline
Ly $\alpha$ & 0.11&0.04 \\
Ly $\beta$ & 0.05& 0.01\\
Ba $\alpha$ & 0.04 & 0.06 \\
Ba $\beta$ &0.02 & 0.03 \\
Two photon &1.1 & 51
\end{tabular}
\end{table}
\section{Conclusions}
\label{sec:discuss}
True muonium and true tauonium are the most compact pure QED systems, but have never been observed. Unlike Ps, for which there are extensive observations in our own Galaxy\cite{pra11}, both the formation and decay of M and T are affected by the intrinsic instability of the $\mu$ and $\tau$ leptons. We have investigated the likelihood of their formation in astrophysical environments, and the prospects for their observation.
The probability of formation is small, $\sim 10^{-7}$ from photon-photon annihilation or electron-positron annihilation. The probability is small for two reasons: (i) the lifetimes of the $\mu$ and $\tau$ are intrinsically short (\S~\ref{sec:mutaudecay}), and thus M and T can only form from the products of pair production processes, such that the pairs immediately recombine, (ii) even then, only those pairs with a total kinetic energy less than the ionisation energy can form leptonium, and since pair production usually takes place in high energy process, these pairs constitute a small fraction of the total.
Nevertheless, high energy astrophysical environments are capable of producing copious numbers of $\mu$ and $\tau$ pairs, and the cross-section for radiative recombination dominates that of direct annihilation (\S~\ref{sec:gould}). Thus even the small fraction of pairs with energy low enough to
form M or T can lead to a significant flux.
The decay of M and T are hastened by the short lifetimes of the $\mu$ and $\tau$ leptons, which may be one reason why the possibility of astrophysical observations has not received much attention. Here, we have carefully calculated the probabilities of the various observational signatures. We have calculated the cross-sections for recombination onto the $nL$th level (\S~\ref{sec:wallyn}), and the resulting branching ratios for Lyman $\alpha$, Lyman $\beta$, Balmer $\alpha$, Balmer $\beta$, two photon annihilation, three photon annihilation, annihilation into e$^{\pm}$ pairs, and the decay of either $\mu$ or $\tau$. Although the decay of either $\tau$ does dominate for T, there is still a small probability for observing recombination lines or two-photon annihilation. For M the situation is more hopeful, with significant branching ratios for Lyman $\alpha$, Balmer $\alpha$ and two-photon annihilation.
In section~\ref{sec:astrophys}, we made estimates of the fluxes of M and T recombination and annihilation signatures for the cases of blazar jet-cloud interactions, jet-star interactions in mis-aligned microquasars, and within the accretion discs of AGN and microquasars. These were compared to the current detection limits of X-ray and $\gamma$-ray observatories. The expected signatures from AGN jet-cloud interactions were all below current detections limits.
However, M and T formation within microquasar jet-star interactions, or within the accretion discs of both AGN and microquasars were estimated to yield signatures brighter than the detection limits, with those from microquasars offering the brightest estimates due to their proximity. Actual observations would be further complicated by the intrinsic backgrounds and also emission from the object in question, which may have other significant line emission (e.g.\ see \citet{mar02} for a spectrum of SS 433).
These examples are only illustrative. Other sources may also be significant. For example, 511 keV electron-positron annihilation radiation was recently discovered in terrestrial $\gamma$-ray flashes due to lightning strikes\cite{bri11}. These same flashes could lead to M or T formation, as well as the possibility of observing Ps recombination lines.
As a different example we note in passing that the recent detection of an unidentified line at $\approx 3.5$~keV in galaxy clusters\cite{boy14,bul14} cannot be explained by the T Balmer $\alpha$ line at 3.3 keV, which is ruled out by the constraints on the energy.
In summary, the astrophysics of pair-production in any high energy source could lead to possible M and T formation. This paper provides the tools to estimate the fluxes of the M and T recombination and annihilation signatures once the rate of pair production is known.
\begin{acknowledgments}
We thank the referees for useful comments which have improved this paper. We thank M.~Voloshin for useful advice regarding the Sommerfeld-Sakharov correction. We thank M.~Colless for advice on the nomenclature for leptonium which improved the lucidity of the paper.
\end{acknowledgments}
\begin{widetext}
|
\section{Introduction}
Recently Cheng, Hwang and Yang \cite{MR2262784}, \cite{MR2481053}, have considered the functional
\begin{equation}
\label{eq:F}
\mathcal{F}(u)=\int_\Omega |\nabla u+\vec{F}|+\int_\Omega fu,
\end{equation}
on a domain $\Omega\subset{\mathbb{R}}^{2n}$, where $\vec{F}$ is a vector field and $f\in L^\infty(\Omega)$. In case $\vec{F}(x,y)=(-y,x)$, the integral $\int_\Omega |\nabla u+\vec{F}|$ is the sub-Riemannian area of the horizontal graph of the function $u$ in the Heisenberg group ${\mathbb{H}}^n$. Among several interesting results, they proved in \cite[Thm.~A]{MR2481053} that, in case $n=1$, $u\in C^1(\Omega)$ is an stationary point of $\mathcal{F}$ and $f\in C^0(\Omega)$, the integral curves of the vector field $((\nabla u+\vec{F})/|\nabla u+\vec{F}|)^\bot$, defined in the set $|\nabla u+\vec{F}|\neq 0$, are of class $C^2$. The geometric meaning of their result is that the projection of the characteristic curves of the graph of $u$ are of class $C^2$. A stationary point $u$ of $\mathcal{F}$ satisfies weakly the prescribed mean curvature equation
\begin{equation}
\label{eq:pmc}
\divv\bigg(\frac{\nabla u+\vec{F}}{|\nabla u+\vec{F}|}\bigg)=f.
\end{equation}
Theorem~A in \cite{MR2481053} is well-known for $C^2$ minimizers and generalizes a previous result by Pauls \cite[Lemma~3.3]{MR2225631} for $H$-minimal surfaces with components of the horizontal Gauss map in the class $W^{1,1}$. For lipschitz continuous vanishing viscosity minimal graphs, it was proven by Capogna, Citti and Manfredini \cite[Cor.~1.6]{MR2583494}.
In order to extend this result to arbitrary surfaces, it is natural to replace $\mathcal{F}$ by the sub-Riemannian prescribed mean curvature functional
\begin{equation}
\label{eq:JOm}
\mathcal{J}(E,B)=P(E,B)+\int_{E\cap B} f,
\end{equation}
where $E$ is a set of locally finite sub-Riemannian perimeter in $\Omega$, $P(E,B)$ is the relative sub-Riemannian perimeter of $E$ in a bounded open set $B\subset\Omega$, and $f\in L^\infty(\Omega)$. If $E\subset{\mathbb{H}}^n$ is the subgraph of a function $t=u(x,y)$ in the Heisenberg group ${\mathbb{H}}^n$, then $\mathcal{J}(E)$ coincides with \eqref{eq:F} taking $\vec{F}(x,y)=(-y,x)$.
The notion of sub-Riemannian perimeter used in sub-Riemannian geometry was first introduced by Capogna, Danielli and Garofalo \cite{MR1312686} for Carnot-Carath\'eodory spaces. General properties and existence of sets with minimum perimeter were proved later by Garofalo and Nhieu \cite{MR1404326}. A rather complete theory of finite perimeter sets in the Heisenberg group ${\mathbb{H}}^n$ following De Giorgi's original arguments was developed by Franchi, Serapioni and Serra-Cassano \cite{MR1871966}, and later extended to step $2$ Carnot groups \cite{MR1984849} by the same authors. The recent monograph \cite{MR2312336} provides a quite complete survey on recent progress on the subject.
We have defined the prescribed mean curvature functional following Massari \cite{MR0355766}, who considered minimizers of $\mathcal{J}$ for the Euclidean relative perimeter. He obtained existence and regularity results for this problem and observed that, in case $E$ is the subgraph of a Lipschitz function $u$ defined on an open bounded set $D\subset{\mathbb{R}}^{n-1}$, the function $u$ satisfies weakly the prescribed mean curvature equation
\[
\divv\bigg(\frac{\nabla u}{\sqrt{1+|\nabla u|^2}}\bigg)(x)=f(x,u(x))
\]
for $x\in D$. In case $\partial E\cap\Omega$ is a hypersurface of class $C^2$ then the mean curvature of $\partial E$ at a point $p\in\partial E$ equals $g(p)$. See also Maggi \cite[pp.~139--140]{MR2976521}.
\mbox{}
The aim of this paper is to extend Cheng, Hwang and Yang's regularity result for characteristic curves \cite[Thm.~A]{MR2481053} from $C^1$ horizontal graphs satisfying weakly the mean curvature equation in the first Heisenberg group ${\mathbb{H}}^1$ to surfaces of class $C^1$ with prescribed mean curvature in arbitrary three-dimensional contact sub-Riemannian manifolds.
In this setting, the Euclidean perimeter is replaced by the sub-Riemannian one and the integral of the function $f$ is computed using Popp's measure \cite[\S~10.6]{MR1867362}, \cite{MR3108867}. The minimizing condition will be replaced by a stationary one. Our ambient space will be a three-dimensional contact manifold with a sub-Riemannian metric defined on its horizontal distribution. In particular, no assumption on the existence of a pseudo-hermitian structure is made. We shall prove in Theorem~\ref{th:main}
\begin{quotation}
Let $E\subset\Omega$ be a set with $C^1$ boundary and prescribed mean curvature $f\in C^0(\Omega)$ in a domain $\Omega\subset M$ of a three-dimensional contact sub-Riemannian manifold. Then characteristic curves in $\partial E$ are of class $C^2$.
\end{quotation}
We remark that \cite[Thm.~A]{MR2481053} states that the projection of characteristic curves to the plane $t=0$ is of class $C^2$, but together with \cite[(2.22)]{MR2165405} this implies that the characteristic curves themselves are $C^2$. We thank J.-H. Cheng for pointing out this fact.
While the proof of \cite[Thm.~A]{MR2481053} was based on the integral formula \cite[(2.3)]{MR2481053}, see also (3.7) in \cite[Remark~3.4]{gr-aim}, the proof of Theorem~\ref{th:main} is purely variational and follows by localizing the first variation of perimeter along a characteristic curve. A much weaker version of Theorem~\ref{th:main} was given in \cite[Thm.~3.5]{gr-aim}, where it was proven that the regular part of an area-stationary surface of class $C^1$ in the sub-Riemannian Heisenberg group ${\mathbb{H}}^1$ is foliated by horizontal geodesics. Theorem~\ref{th:main} provides a new result even for the case of the first Heisenberg group ${\mathbb{H}}^1$.
The regularity of characteristic curves proven in Theorem~\ref{th:main} allows us to define in Section~\ref{sec:mc} a mean curvature function $H$ in the regular part of $\partial E$, that coincides with $f$. As a consequence of the definition of the mean curvature, we shall prove in Proposition~\ref{prop:C^k} that characteristic curves are of class $C^{k+2}$ in case $f$ is of class $C^k$ when restricted to a characteristic direction. This holds, e.g., when $f\in C^k(\Omega)$ of $f\in C_\mathbb{H}^k(\Omega)$, the space of functions with continuous horizontal derivatives of order $k$, $k\ge 1$. This class contains $C^1(\Omega)$ when $k\ge 2$. Critical points of the perimeter, eventually under a volume constraint, and $C^1$ boundary, have constant prescribed mean curvature as shown in Section~\ref{sec:pmc}. Hence Theorem~\ref{th:main} applies to these sets and implies that the regular parts of their boundaries are foliated by $C^\infty$ characteristic curves, see Proposition~\ref{prop:critical}.
We have organized this paper into several sections. In the second one we provide the necessary background on contact sub-Riemannian manifolds and sets of finite perimeter, and we recall the first variation formula for $C^1$ surfaces following \cite{Gaphd}. In Section~\ref{sec:pmc} we introduce the definition of set of locally finite perimeter with prescribed mean curvature and prove that a set with $C^1$ boundary and area-stationary under a volume constraint has constant prescribed mean curvature. The main result, Theorem~\ref{th:main}, is proven in Section~\ref{sec:main}. The consequences on the mean curvature and higher regularity for characteristic curves will appear in Section~\ref{sec:mc}.
\section{Preliminaries}
\subsection{Contact sub-Riemannian manifolds}
In this paper we shall consider a $3$-dimensional $C^\infty$ manifold $M$ with contact form $\omega$ and a sub-Riemannian metric $g_\mathcal{H}$ defined on its \emph{horizontal distribution} $\mathcal{H}:=\text{ker}(\omega)$. By definition, $d\omega|_{\mathcal{H}}$ is non-degenerate. We shall refer to $(M,\omega,g_\mathcal{H})$ as a \emph{$3$-dimensional contact sub-Riemannian manifold}. It is well-known that $\omega\wedge d\omega$ is an orientation form in $M$. Since
\[
d\omega(X,Y)=X(\omega(Y))-Y(\omega(X))-\omega([X,Y]),
\]
the horizontal distribution $\mathcal{H}$ is completely non-integrable. The \emph{Reeb vector field} $T$ in $M$ is the only one satisfying
\begin{equation}
\label{eq:reeb}
\omega(T)=1,\qquad \mathcal{L}_T\omega=0,
\end{equation}
where $\mathcal{L}$ is the Lie derivative in $M$.
A canonical contact structure in Euclidean $3$-space ${\mathbb{R}}^3$ with coordinates $(x,y,t)$ is given by the contact one-form $\omega_0:=dt+xdy-ydx$. The associated contact manifold is the Heisenberg group ${\mathbb{H}}^1$. Darboux's Theorem \cite[Thm.~3.1]{MR1874240} (see also \cite{MR2979606}) implies that, given a point $p\in M$, there exists an open neighborhood $U$ of $p$ and a diffeomorphism $\phi_p$ from $U$ into an open set of ${\mathbb{R}}^{3}$ satisfying $\phi_p^*\omega_0=\omega$. Such a local chart will be called a \emph{Darboux chart}. Composing the map $\phi_p$ with a contact transformation of ${\mathbb{H}}^1$ also provides a Darboux chart. This implies we can prescribe the image of a point $p\in U$ and the image of a horizontal direction in $T_pM$.
The metric $g_\mathcal{H}$ can be extended to a Riemannian metric $g$ on $M$ by requiring $T$ to be a unit vector orthogonal to $\mathcal{H}$. The Levi-Civita connection associated to $g$ will be denoted by $D$. The integral curves of the Reeb vector field $T$ are \emph{geodesics} of the metric $g$. This property can be easily checked since condition $\mathcal{L}_T\omega=0$ in \eqref{eq:reeb} implies $\omega([T,X])=0$ for any\ $X\in\mathcal{H}$. Hence, for any horizontal vector field $X$, we have
\[
\escpr{X,D_TT}=-\escpr{D_TX,T}=-\escpr{D_XT,T}=0.
\]
We trivially have $\escpr{T,D_TT}=0$, and so we get $D_TT=0$, as claimed.
The Riemannian volume element in $(M,g)$ will be denoted by $dM$. It coincides with Popp's measure \cite[\S~10.6]{MR1867362}, \cite{MR3108867}. The volume of a set $E\subset M$ with respect to the Riemannian metric $g$ will be denoted by $|E|$.
\subsection{Torsion and the sub-Riemannian connection}
The following is taken from \cite[\S~3.1.2]{Gaphd}. In a contact sub-Riemannian manifold, we can decompose the endomorphism $X\in TM\rightarrow D_X T$ into its antisymmetric and symmetric parts, which we will denoted by $J$ and $\tau$, respectively,
\begin{equation}\label{eq:jtau}
\begin{split}
2\escpr{J(X),Y}&=\escpr{D_X T,Y}-\escpr{D_Y T,X},\\
2\escpr{\tau(X),Y}&=\escpr{D_X T,Y}+\escpr{D_Y T,X}.
\end{split}
\end{equation}
Observe that $J(X),\tau(X)\in\mathcal{H}$ for any vector field $X$, and that $J(T)=\tau(T)=0$. Also note that
\begin{equation}\label{eq:J}
2\escpr{J(X),Y}=-\escpr{[X,Y],T}, \qquad X,Y\in\mathcal{H}.
\end{equation}
We will call $\tau$ the \emph{$($contact$)$ sub-Riemannian torsion}. We note that our $J$ differs from the one defined in \cite[(2.4)]{MR3044134} by the constant $g([X,Y],T)$, but plays the same geometric role and can be easily generalized to higher dimensions, \cite[\S~3.1.2]{Gaphd}.
Now we define the \emph{$($contact$)$ sub-Riemannian connection} $\nabla$ as the unique metric connection, \cite[eq.~(I.5.3)]{MR2229062}, with torsion tensor $\tor(X,Y)=\nabla_XY-\nabla_YX-[X,Y]$ given by
\begin{equation}\label{def:tor}
\tor(X,Y):=\escpr{X,T}\,\tau(Y)-\escpr{Y,T}\,\tau(X)+2\escpr{J(X),Y}\,T.
\end{equation}
From \eqref{def:tor} and Koszul formula for the connection $\nabla$ it follows that $T$ is a parallel vector field for the sub-Riemannian connection. In particular, their integral curves are geodesics for the connection $\nabla$.
If $X\in\mathcal{H}$, $p\in M$, and $X_p\neq 0$, then $J(X_p)\neq 0$: as $d\omega|_{\mathcal{H}}$ is non-degenerate, there exists $Y\in\mathcal{H}$ such that $d\omega_p(X_p,Y_p)\neq 0$. From \eqref{eq:jtau} we have $2\,g(J(X_p),Y_p)=-g([X,Y]_p,T_p)$, different from $0$ since $\omega_p([X,Y]_p)=-d\omega(X_p,Y_p)\neq 0$.
The standard orientation of $M$ is given by the $3$-form $\omega\wedge d\omega$. If $X_p$ is horizontal, then the basis $\{X_p,J(X_p),T_p\}$ is positively oriented. To check this, observe first that the sign of $(\omega\wedge d\omega)(X,J(X),T)$ equals the sign of $d\omega(X,J(X))$, and we have
\[
d\omega(X,J(X))=-\omega([X,J(X)])=-g([X,J(X)],T)=g(\tor(X,J(X)),T)=2\,g(J(X),J(X))>0.
\]
\subsection{Perimeter and $C^1$ surfaces}
A set $E\subset M$ has \emph{locally finite perimeter} if, for any bounded open set $B\subset M$, we have
\[
P(E,B):=\sup\bigg\{\int_{E\,\cap\,B}\divv U\,dM: U\ \text{horizontal},\ \text{supp}(U)\subset B, ||U||_\infty\le 1\bigg\}<+\infty.
\]
The quantity $P(E,B)$ is the \emph{relative perimeter} of $E$ in $B$.
Assuming $\Sigma=\partial E$ is a surface of class $C^1$, the relative perimeter of $E$ in a bounded open set $B\subset M$ coincides with the sub-Riemannian area of $\Sigma\cap B$, given by
\begin{equation}
\label{eq:areaC1}
A(\Sigma\cap B)=\int_{\Sigma\,\cap\, B}|N_h|\,d\Sigma.
\end{equation}
Here $N$ is the Riemannian unit normal to $\Sigma$, $N_h$ is the horizontal projection of $N$ to the horizontal distribution, and $d\Sigma$ is the Riemannian area measure, all computed with respect the Riemannian metric $g$, see \cite{MR2312336}. The quantity $|N_h|$ vanishes in the \emph{singular set} $\Sigma_0\subset\Sigma$ of points $p\in\Sigma$ where the tangent space $T_p\Sigma$ coincides with the horizontal distribution $\mathcal{H}_p$. The \emph{horizontal unit normal} at $p\in\Sigma\setminus\Sigma_0$ is defined by $(\nu_{h})_p:=(N_h)_p/|(N_h)_p|$. At every point $p\in\Sigma\setminus\Sigma_0$, the intersection $\mathcal{H}_p\cap T_p\Sigma$ is one-dimensional and generated by the characteristic vector field $Z:=J(\nu_{h})/|J(\nu_{h})|$. The vector $S_p$ is defined for $p\in\Sigma\setminus\Sigma_0$ by $S_p:=g(N_p,T_p)\,(\nu_{h})_p-|(N_h)_p|\,T_p$. The tangent space $T_p\Sigma$, $p\in\Sigma\setminus\Sigma_0$, is generated by $\{Z_p,S_p\}$.
\subsection{The first variation of the sub-Riemannian perimeter for $C^1$ surfaces}
Given a set $E$ with $C^1$ boundary, we can use the flow $\{\varphi_s\}_{s\in{\mathbb{R}}}$ of a vector field $U$ with compact support in $B$ to produce a variation of $\Sigma\cap B$. The Riemannian area formula gives the following expression of the sub-Riemannian area of $\Sigma_s:=\varphi_s(\Sigma\cap B)$,
\[
A(\Sigma_s)=\int_\Sigma |N^s_h|\,\Jac(\varphi_s)\,d\Sigma,
\]
where $N^s$ is a unit normal to $\Sigma_s$. Fix $p\in\Sigma\setminus\Sigma_0$ and the orthonormal basis $\{e_1,e_2\}=\{Z_p,S_p\}$ in $T_p\Sigma$. We consider extensions $E_1$, $E_2$ of $Z_p$, $S_p$, respectively, along the integral curve of $U$ passing through $p$. The vector fields $E_1(s)$, $E_2(s)$ are invariant under the flow of $U$ and generate the tangent plane to $\Sigma_s$ at the point $\varphi_s(p)$. The vector $(E_1\times E_2)/|(E_1\times E_2)|$ is normal to $\Sigma_s$. Here $\times$ denotes the cross product with respect to a volume form $\eta$ for the metric $g$ inducing the same orientation as $\omega\wedge d\omega$, i.e. $g(w,u\times v)=\eta(w,u,v)$. It is easy to check that $|(E_1\times E_2)|(s)=\Jac(\varphi_s)(p)$, and that
\[
V(p,s):=(E_1\times E_2)_h(s)=\big(\escpr{E_1,T}\,(T\times E_2)-\escpr{E_2,T}\,(E_1\times T)\big)(s).
\]
Hence
\[
A(\Sigma_s)=\int_\Sigma |V(p,s)|\,d\Sigma(p),
\]
and we get
\[
\frac{d}{ds}\bigg|_{s=0} |V(s,p)|=\frac{\escpr{\nabla_{U_p} V,V_p}}{|V_p|}.
\]
Since $\{(\nu_{h})_p,Z_p,T_p\}$ is positively oriented, observe that $V_p=|(N_h)_p|\,(\nu_h)_p$. On the other hand,
\begin{align*}
\nabla_{U_p} V=\escpr{\nabla_{U_p} E_1,T_p}\,(T_p\times (E_2)_p)-\escpr{\nabla_{U_p}E_2,T_p}\,&((E_1)_p\times T_p)
\\
&-\escpr{(E_2)_p,T_p}\,(\nabla_{U_p} E_1\times T_p),
\end{align*}
and so
\begin{equation*}
\frac{\escpr{\nabla_{U_p}V,V_p}}{|V_p|}=-\escpr{\nabla_{U_p}E_2,T_p}+|(N_h)_p|\,\escpr{\nabla_{U_p} E_1\times T_p,(\nu_{h})_p}.
\end{equation*}
Since
\begin{align*}
\escpr{\nabla_{U_p}E_2,T_p}&=\escpr{\nabla_{(E_2)_p}U+\tor(U_p,(E_2)_p),T_p}
\\
&=S_p(\escpr{U,T})+\escpr{\tor(U_p,S_p),T_p}
\\
&=S_p(\escpr{U,T})+2\,\escpr{J(U_p),S_p},
\end{align*}
and
\begin{align*}
\escpr{\nabla_{U_p} E_1\times T_p,(\nu_{h})_p}&=\escpr{(\nabla_{(E_1)_p}U+\tor(U_p,(E_1)_p))\times T_p,(\nu_{h})_p}
\\
&=\eta((\nu_{h})_p,\nabla_{(E_1)_p}U+\tor(U_p,(E_1)_p),T_p)
\\
&=+\escpr{\nabla_{Z_p}U+\tor(U_p,Z_p),Z_p}
\\
&=+\escpr{\nabla_{Z_p}U,Z_p}+\escpr{U_p,T_p}\,\escpr{\tau(Z_p),Z_p},
\end{align*}
we conclude that the first variation of the sub-Riemannian perimeter is given by
\begin{equation}
\label{eq:1starea}
\begin{split}
\frac{d}{ds}\bigg|_{s=0} A(\Sigma_s)=\int_{\Sigma\,\cap\,B}\big\{-S(g(U,T))-2\,g(J(U),S)&+|N_h|\,g(\nabla_ZU,Z)
\\
&+|N_h|\,\escpr{U,T}\,\escpr{\tau(Z),Z}\big\}\,d\Sigma.
\end{split}
\end{equation}
This formula was obtained in \cite[Lemma~3.4]{MR3044134}.
\section{Sets with prescribed mean curvature}
\label{sec:pmc}
The reader is referred to \cite[(12.32) and Remark~17.11]{MR2976521} for background and references in the Euclidean case. Consider a domain $\Omega\subset M$, and a function $f:\Omega\to{\mathbb{R}}$. We shall say that a set of locally finite perimeter $E\subset\Omega$ has \emph{prescribed mean curvature $f$ on $\Omega$} if, for any bounded open set $B\subset\Omega$, $E$ is a critical point of the functional
\begin{equation}
\label{eq:a-hv}
P(E,B)-\int_{E\cap B} f,
\end{equation}
where $P(E,B)$ is the relative perimeter of $E$ in $B$, and the integral on $E\cap B$ is computed with respect to the canonical Popp's measure on $M$, see \cite{MR1867362} and \cite{MR3108867}. The admissible variations for this problem are the flows induced by vector fields with compact support in $B$.
If $\Sigma=\partial E$ is a surface of class $C^1$ in $\Omega$, then $\Sigma$ has prescribed mean curvature $f$ if it is a critical point of the functional
\begin{equation}\label{eq:prescribedfunctional}
A(\Sigma\cap B) -\int_{E\cap B} f,
\end{equation}
for any bounded open set $B\subset\Omega$.
If $E$ is a critical point of the relative perimeter $P(E,B)$ in any bounded open set $B\subset\Omega$, then $E$ has zero or vanishing prescribed mean curvature.
Assume now that $E\subset\Omega$ is a set of locally finite perimeter with $C^1$ boundary $\Sigma$, and that $E$ is a \emph{critical point of the perimeter under a volume constraint}. This means $(d/ds)_{s=0} A(\varphi_s(\Sigma\cap B))=0$ for any flow associated to a vector field with compact support in $\Omega$ satisfying $(d/ds)_{s=0} |\varphi_s(E\cap B)|=0$. If the perimeter of $E$ in $\Omega$ is positive, then there exists a (horizontal) vector field $U_0$ with compact support in $\Omega$ so that $\int_{E\cap\Omega}\divv U_0\,dM>0$. By the Divergence Theorem,
\[
\int_{\Sigma\cap\Omega} g(U_0,N)\,d\Sigma\neq 0,
\]
where $N$ is the outer normal to $E$. Let $\{\psi_s\}_{s\in{\mathbb{R}}}$ be the flow associated to the vector field $U_0$ and define
\begin{equation}
\label{eq:defH0}
H_0:=\frac{d/ds\big|_{s=0}\, A(\psi_s(\Sigma))}{d/ds\big|_{s=0}|\psi_s(E)|}.
\end{equation}
Let $B\subset \Omega$ be a bounded open subset and $W$ a vector field with compact support in $B$ and associated flow $\{\varphi_s\}_{s\in{\mathbb{R}}}$. Choose $\lambda\in{\mathbb{R}}$ so that $W-\lambda U_0$ satisfies
\[
\frac{d}{ds}\bigg|_{s=0}|\varphi_s(E)|-\lambda\,\frac{d}{ds}\bigg|_{s=0}|\psi_s(E)|=\int_{\Sigma} g(W-\lambda U_0,N)\,d\Sigma=0.
\]
Then the flow associated to $W-\lambda U_0$ preserves the volume of $E\cap(B\cup B_0)$, where $B_0\subset\Omega$ is a bounded open set containing $\text{supp}(U_0)$. Let $Q(U)$ be the integral expression in \eqref{eq:1starea}. From our hypothesis and the linearity of \eqref{eq:1starea}, $Q(W-\lambda U_0)=0$. Hence $Q(W)=\lambda\,Q(U_0)$. From \eqref{eq:defH0} we get
\[
Q(W)=\lambda Q(U_0)=\lambda H_0\frac{d}{ds}\bigg|_{s=0}|\psi_s(E)|=H_0\frac{d}{ds}\bigg|_{s=0}|\varphi_s(E)|,
\]
and so $E$ has (constant) prescribed mean curvature $H_0$.
\section{Main result}
\label{sec:main}
In this Section we shall prove our main result
\begin{theorem}
\label{th:main}
Let $M$ be a $3$-dimensional contact sub-Riemannian manifold, $\Omega\subset M$ a domain, and $E\subset\Omega$ a set of prescribed mean curvature $f\in C^0(\Omega)$ with $C^1$ boundary $\Sigma$. Then the characteristic curves in $\Sigma$ are of class $C^2$.
\end{theorem}
\begin{proof}
Given any point $p\in\Sigma\setminus\Sigma_0$, consider a Darboux chart $(U_p,\phi_p)$ such that $\phi_p(p)=0$. The metric $g_{\mathcal{H}}$ can be described in this local chart by the matrix of smooth functions
\[
G=\begin{pmatrix}
g_{11} & g_{12}
\\
g_{21} & g_{22}
\end{pmatrix}
=
\begin{pmatrix}
g(X,X) & g(X,Y)
\\
g(Y,X) & g(Y,Y)
\end{pmatrix}.
\]
After a Euclidean rotation around the $t$-axis, which is a contact transformation in ${\mathbb{H}}^1$ \cite[p.~640]{MR2435652}, we may assume there exists an open neighborhood $B\cap\Sigma$ of $p\in \Sigma\setminus\Sigma_0$, where $B\subset{\mathbb{H}}^1$ is an open set containing $p$, so that $B\cap\Sigma$ is the intrinsic graph $G_u$ of a $C^1$ function $u:D\to{\mathbb{R}}$ defined on a domain $D$ in the vertical plane $y=0$. We can also assume that $E\cap B$ is the subgraph of $u$. The graph $G_u$ can be parameterized by the map $f_u:D\to{\mathbb{R}}^3$ defined by
\[
f_u(x,t):=(x,u(x,t),t-xu(x,t)),\qquad (x,t)\in D.
\]
The tangent plane to any point in $G_u$ is generated by the vectors
\begin{align*}
\tfrac{\partial}{\partial x}&\mapsto (1,u_x,-u-xu_x)=X+u_xY-2uT,
\\
\tfrac{\partial}{\partial t}&\mapsto (0,u_t,1-xu_t)=u_tY+T,
\end{align*}
and so the characteristic direction is given by $Z=\widetilde{Z}/|\widetilde{Z}|$, where
\[
\widetilde{Z}=X+(u_x+2uu_t)Y.
\]
If $\gamma(s)=(x(s),t(s))$ is a $C^1$ curve in $D$, then
\[
\Gamma(s)=(x(s),u(x(s),t(s)),t(s)-x(s)u(x(s),t(s)))\subset G_u
\]
is also $C^1$, and so
\[
\Gamma'(s)=x'\,(X+u_xY-2uT)+t'\,(u_tY+T)=x'X+(x'u_x+t'u_t)Y+(t'-2ux')T.
\]
In particular, horizontal curves in $G_u$ satisfy the ordinary differential equation $t'=2ux'$. Since $u\in C^1(D)$, we have uniqueness of characteristic curves through any given point in $G_u$.
A unit normal vector to $\Sigma$ is given by $\widetilde{N}/|\widetilde{N}|$, where
\[
\widetilde{N}=(X+u_xY-2uT)\times (u_tY+T).
\]
Here $\times$ is the cross product with respect to the Riemannian metric $g$ and a given volume form $\eta$ chosen so that $\eta(X,Y,T)>0$. Hence $g(w,u\times v)=\eta(w,u,v)$. If $\{e_1,e_2,e_3\}$ is an orthonormal basis so that $\eta(e_1,e_2,e_3)=1$ and $A$ is the matrix whose columns are the coordinates of $X$, $Y$, $T$ in the basis $\{e_1,e_2,e_3\}$, then $\eta(X,Y,T)=\det(A)$. On the other hand, as
\[
A^tA=\begin{pmatrix}
g_{11} & g_{12} & 0
\\
g_{21} & g_{22} & 0
\\
0 & 0 & 1
\end{pmatrix},
\]
we get $\det(A)^2=\det(G)$. Since $\det(A)>0$ we obtain $\det(A)=\det(G)^{1/2}$ and so
\[
\eta(X,Y,T)=\det(G)^{1/2}.
\]
Let $E_1=X+u_xY-2uT$, $E_2=u_tY+T$. We compute the scalar product of $\widetilde{N}=E_1\times E_2$ with $X$, $Y$, $T$ to obtain
\begin{align*}
g(X,E_1\times E_2)&=\eta(X,E_1,E_2)=\det
{\arraycolsep=0.3\arraycolsep\begin{pmatrix}
1 & 1 & 0
\\
0 & u_x & u_t
\\
0 & -2u & 1
\end{pmatrix}}
\,\eta(X,Y,T)=(u_x+2uu_t)\det(G)^{1/2}.
\\
g(Y,E_1\times E_2)&=\eta(Y,E_1,E_2)=\det
{\arraycolsep=0.3\arraycolsep\begin{pmatrix}
0 & 1 & 0
\\
1 & u_x & u_t
\\
0 & -2u & 1
\end{pmatrix}}
\,\eta(X,Y,T)=-\det(G)^{1/2}.
\\
g(T,E_1\times E_2)&=\eta(T,E_1,E_2)=\det
{\arraycolsep=0.3\arraycolsep\begin{pmatrix}
0 & 1 & 0
\\
0 & u_x & u_t
\\
1 & -2u & 1
\end{pmatrix}}
\,\eta(X,Y,T)=u_t \det(G)^{1/2}.
\end{align*}
Since $g(Y,E_1\times E_2)<0$ and $E\cap B$ is the subgraph of $u$, the vector field $E_1\times E_2$ points into the interior of $E$. If $\widetilde{N}=E_1\times E_2=\alpha X+\beta Y+\gamma T$, then
\[
\begin{pmatrix}
g_{11} & g_{12} & 0
\\
g_{21} & g_{22} & 0
\\
0 & 0 & 1
\end{pmatrix}
\begin{pmatrix}
\alpha \\ \beta \\ \gamma
\end{pmatrix}=\det(G)^{1/2}
\begin{pmatrix}
u_x+2uu_t \\ -1 \\ u_t
\end{pmatrix},
\]
whence
\begin{equation}
\label{eq:Ntilde}
\begin{split}
\begin{pmatrix} \alpha \\ \beta \end{pmatrix}&=\det(G)^{1/2}\, G^{-1}
\begin{pmatrix} u_x+2uu_t \\ -1 \end{pmatrix},
\\
\gamma&=\det(G)^{1/2} u_t.
\end{split}
\end{equation}
Let us compute now the sub-Riemmanian area of the intrinsic graph $G_u$. It is easy to check that $d\Sigma=|E_1\times E_2|\,dxdt$, i.e., that $\Jac(f_u)=|E_1\times E_2|$. Since $|N_h|=|(E_1\times E_2)_h|/|E_1\times E_2|$, then using \eqref{eq:Ntilde} and the explicit expression of the inverse matrix $G^{-1}$ we get
\begin{align*}
|N_h|\,\Jac(f_u)&=|(E_1\times E_2)_h|=\bigg(\big(\alpha\ \ \beta\big)\, G\,
\begin{pmatrix} \alpha \\ \beta \end{pmatrix}\bigg)^{\frac{1}{2}}
\\
&=\big((g_{22}\circ f_u)(u_x+2uu_t)^2+2\,(g_{12}\circ f_u)(u_x+2uu_t)+(g_{11}\circ f_u)\big)^{1/2}.
\end{align*}
Finally, from \eqref{eq:areaC1} we obtain
\begin{equation}
\label{eq:areagraph}
A(G_u)=\int_D \big(g_{22}(u_x+2uu_t)^2+2g_{12}(u_x+2uu_t)+g_{11}\big)^{1/2}dxdt,
\end{equation}
where, by abuse of notation, we have written $g_{ij}$ instead of the cumbersome notation $(g_{ij}\circ f_u)$.
Now we consider variations of $G_u$ by graphs of the form $s\mapsto u+sv$, where $v\in C_0^\infty(D)$ and $s$ is a real parameter close to $0$. This variation is obtained by applying the flow associated to the vector field $\tilde{v}Y$ to the graph $G_u$. The function $\tilde{v}$ is obtained by extending $v$ to be constant along the integral curves of the vector field $Y$, and multiplying by an appropriate function with compact support equal to $1$ in a neighborhood of $\Sigma$.
When $F$ is a function of $(x,y,t)$, we have
\[
\frac{d}{ds}\bigg|_{s=0} (F\circ f_{u+sv})(x,t)=\bigg(\frac{\partial F}{\partial y}-x\frac{\partial F}{\partial t}\bigg)_{f_u(x,t)}\,v(x,t)=Y_{f_u(x,t)} (F)\, v(x,t).
\]
So we get
\begin{equation*}
\frac{d}{ds}\bigg|_{s=0} A(G_{u+sv}
=\int_D \big(K_1v+M\,(v_x+2uv_t+2vu_t)\big)\,dxdt,
\end{equation*}
where the functions $K_1$ and $M$ are given by
\[
K_1=\frac{1}{2}\frac{Y(g_{22})(u_x+2uu_t)^2+2 Y(g_{12})(u_x+2uu_t)+Y(g_{11})}{(g_{22}(u_x+2uu_t)^2+2g_{12}(u_x+2uu_t)+g_{11})^{1/2}},
\]
and
\[
M=\frac{g_{22}(u_x+2uu_t)+g_{12}}{(g_{22}(u_x+2uu_t)^2+2g_{12}(u_x+2uu_t)+g_{11})^{1/2}}.
\]
Observe that the functions $K_1$ and $M$ are continuous. Since
\[
Z=\frac{X+(u_x+2uu_t)\,Y}{(g_{22}(u_x+2uu_t)^2+2g_{12}(u_x+2uu_t)+g_{11})^{1/2}},
\]
the function $M$ coincides with $g(Z,Y)\circ f_u$. A straightforward computation implies
\[
1=|Z|^2=\det(G)^{-1}\big(g_{22}\escpr{Z,X}^2-2g_{12}\escpr{Z,X}\escpr{Z,Y}+g_{11}\escpr{Z,Y}^2\big)
\]
and so
\begin{equation}
\label{eq:ZX}
\escpr{Z,X}=\frac{g_{12}g(Z,y)\pm (\det(G)(g_{22}-\escpr{Z,Y}^2))^{1/2}}{g_{22}}.
\end{equation}
By Schwarz's inequality $\escpr{Z,Y}^2\le \escpr{Y,Y}=g_{22}$. Inequality is strict since otherwise $Y$ and $Z$ would be collinear. Hence $\escpr{Z,X}$ has the same regularity as $\escpr{Z,Y}$ by \eqref{eq:ZX}.
The subgraph of $u$ can be parameterized by the map $(x,t,s)\to (x,s,t-xs)$. The Jacobian of this map is easily seen to be equal to $\det(G)$. Hence
\[
\frac{d}{ds}\bigg|_{s=0} \int_{\text{subgraph}\,G_{u+sv}} f=\int_D f\,\det(G)\,v\,dxdt.
\]
If $\Sigma$ has prescribed mean curvature $f$, this implies
\begin{equation}\label{eq:1stvariation smooth test function}
\int_D \big(Kv+M\,(v_x+2uv_t+2vu_t\big)\,dxdt=0,
\end{equation}
for any $v\in C_0^\infty(D)$, where the \emph{continuous} function $K$ is given by $K=K_1-f\,\det(G)$. By Remark \ref{remark:continuos approximation argument} below, \eqref{eq:1stvariation smooth test function} also holds for any $v\in C^0_0(D)$ for which $v_x+2uv_t$ exists and it is continuous.
Now we proceed as in the proof of Theorem~3.5 in \cite{gr-aim}. Assume the point $p\in G_u$ corresponds to the point $(a,b)$ in the $xt$-plane. The curve $s\mapsto (s,t(s))$ is (a reparameterization of the projection of) a characteristic curve if and only if the function $t(s)$ satisfies the ordinary differential equation $t'(s)=u(s,t(s))$. For $\varepsilon$ small enough, we consider the solution $t_\varepsilon$ of equation $t_\varepsilon'(s)=2u(s,t_\varepsilon(s))$ with initial condition $t_\varepsilon(a)=b+\varepsilon$, and define $\gamma_\varepsilon(s):=(s,t_\varepsilon(s))$, with $\gamma=\gamma_0$. We may assume that, for small enough $\varepsilon$, the functions $t_\varepsilon$ are defined in the interval $[a-r,a+r]$ for some $r>0$. The function $\partial t_\varepsilon/\partial\varepsilon$ satisfies
\begin{equation}
\label{eq:dteps}
\bigg(\frac{\partial t_\varepsilon}{\partial\varepsilon}\bigg)'(s)=2u_t(s,t_\varepsilon(s))\,\bigg(\frac{\partial t_\varepsilon}{\partial\varepsilon}\bigg)(s),\qquad \frac{\partial t_\varepsilon}{\partial\varepsilon}(a)=1.
\end{equation}
where $'$ is the derivative with respect to the parameter $s$.
We consider the parameterization
\[
F(\xi,\varepsilon):= (\xi,t_\varepsilon(\xi))=(s,t)
\]
near the characteristic curve through $(a,b)$. The jacobian of this parameterization is given by
\[
\det\begin{pmatrix}
1 & t'_\varepsilon
\\
0 & \partial t_\varepsilon/\partial \varepsilon
\end{pmatrix}=\frac{\partial t_\varepsilon}{\partial\varepsilon},
\]
which is positive because of the choice of initial condition for $t_\varepsilon$ and the fact that the curves $\gamma_\varepsilon(s)$ foliate a neighborhood of $(a,b)$. Any function $\varphi$ can be considered as a function of the variables $(\xi,\varepsilon)$ by making $\tilde{\varphi}(\xi,\varepsilon):=\varphi(\xi,t_\varepsilon(\xi))$. Changing variables, and assuming the support of $\varphi$ is contained in a sufficiently small neighborhood of $(a,b)$, we can express the integral \eqref{eq:areagraph} as
\[
\int_{I}\bigg\{\int_{a-r}^{a+r}\bigg(K\tilde{\varphi}+M\,\bigg(\frac{\partial\tilde{\varphi}}{\partial\xi}+2\tilde{\varphi} \tilde{u}_t\bigg)\bigg)\,\frac{\partial t_\varepsilon}{\partial\varepsilon}\,d\xi\bigg\}\,d\varepsilon,
\]
where $I$ is a small interval containing $0$. Instead of $\tilde{\varphi}$, we can consider the function $\tilde{\varphi} h /( t_{\varepsilon+h}-t_\varepsilon)$, where $h$ is a sufficiently small real parameter. We get that
\[
\frac{ \partial}{\partial\xi}\bigg(\frac{h\, \tilde{\varphi}}{ t_{\varepsilon+h}-t_\varepsilon}\bigg)=\frac{\partial\tilde{\varphi}}{\partial\xi}\cdot \frac{h}{t_{\varepsilon+h}-t_\varepsilon}-2\tilde{\varphi}\cdot \frac{\tilde{u}(\xi, \varepsilon+h)-\tilde{u}(\xi,\varepsilon)}{t_{\varepsilon+h}-t_\varepsilon}\cdot \frac{h}{t_{\varepsilon+h}-t_\varepsilon}
\]
tends to
\[
\frac{\partial\tilde{\varphi}/\partial\xi}{\partial t_\varepsilon/\partial\varepsilon}-\frac{2\tilde{\varphi}\tilde{u}_t}{\partial t_\varepsilon/\partial\varepsilon},
\]
when $h\rightarrow 0$. So using that $G_u$ is area-stationary we have that
\[
\int_{I}\bigg\{\int_{a-r}^{a+r} \frac{h}{t_{\varepsilon+h}-t_\varepsilon}\bigg(K\, \tilde{\varphi}+M\,\bigg(\frac{\partial\tilde{\varphi}}{\partial\xi}\cdot +2 \tilde{\varphi}\cdot \bigg(\tilde{u}_t-\frac{\tilde{u}(\xi, \varepsilon+h)-\tilde{u}(\xi,\varepsilon)}{t_{\varepsilon+h}-t_\varepsilon}\bigg)\bigg)\bigg)\,\frac{\partial t_\varepsilon}{\partial\varepsilon}\,d\xi\bigg\}\,d\varepsilon
\]
vanishes. Furthermore, letting $h\rightarrow 0$ we conclude
\[
\int_{I}\bigg\{\int_{a-r}^{a+r}\bigg(K\tilde{\varphi}+M\,\frac{\partial\tilde{\varphi}}{\partial\xi}\bigg)\,d\xi\bigg\}\,d\varepsilon=0.
\]
Let now $\eta:{\mathbb{R}}\to{\mathbb{R}}$ be a positive function with compact support in the interval $I$ and consider the family $\eta_\rho(x):=\rho^{-1}\eta(x/\rho)$. Inserting a test function of the form $\eta_\rho(\varepsilon)\psi(\xi)$, where $\psi$ is a $C^\infty$ function with compact support in $(a-r,a+r)$, making $\rho\to 0$, and using that $G_u$ is area-stationary we obtain
\[
\int_{a-r}^{a+r} \big(K(0,\xi)\,\psi(\xi)+M(0,\xi)\,\psi'(\xi)\big)\,d\xi=0
\]
for any $\psi\in C^\infty_0((a-r,a+r))$. By Lemma~\ref{lem:c1}, the function $M(0,\xi)$, which is the restriction of $\escpr{Z,Y}$ to the characteristic curve, is a $C^1$ function on the curve. By equation \eqref{eq:ZX}, the restriction of $\escpr{Z,X}$ to the characteristic curve is also $C^1$. This proves that horizontal curves are of class $C^2$.
\end{proof}
\begin{lemma}
\label{lem:c1}
Let $I\subset{\mathbb{R}}$ be an open interval, $k$, $m\in C^0(I)$, and $K\in C^1(I)$ be a primitive of $k$. Assume
\begin{equation}
\label{eq:inteq}
\int_I k\psi+m\psi'=0,
\end{equation}
for any $\psi\in C_0^\infty(I)$. Then the function $-K+m$ is constant on $I$. In particular, $m\in C^1(I)$.
\end{lemma}
\begin{proof}
Since $(K\psi)'=k\psi+K\psi'$, integrating by parts we see that \eqref{eq:inteq} is equivalent to
\[
\int_I \big(-K+m\big)\,\psi'=0,
\]
for any $\psi\in C^\infty_0(I)$. This implies that $-K+m$ is a constant function on $I$.
\end{proof}
\begin{remark}\label{remark:continuos approximation argument} Let us check that $\eqref{eq:1stvariation smooth test function}$ holds for any $w\in C^0_0(D)$ such that $w_x+2uw_t$ exists and is continuous.
Let us consider a sequence $w_j\in C_0^\infty(D)$, where $w_j=\rho_j \ast w$, and $\rho_j$ denote the standard mollifiers. We have that $w_j$ converges to $w$ and that $(w_j)_x+2u(w_j)_t$ converges to $w_x+2uw_t$ uniformly on compact subsets of $D$, for $j\rightarrow \infty$. We conclude
\[
0=\lim\limits_{j\rightarrow\infty} \int_D \big(K(w_j)+M\,((w_j)_x+2u(w_j)_t+2w_ju_t\big)\,dxdt= \int_D \big(Kw+M\,(w_x+2uw_t+2wu_t\big)\,dxdt,
\]
thus proving the claim.
\end{remark}
\begin{remark}
In case $M$ is the Heisenberg group ${\mathbb{H}}^1$, $G$ is the identity matrix and the expression for the sub-Riemannian area of the graph $G_u$ given in \eqref{eq:areagraph} reads
\[
A(G_u)=\int_D ((u_x+2uu_t)^2+1)^{1/2}\,dxdt,
\]
a well-known formula obtained in \cite{MR2223801}.
\end{remark}
\section{The mean curvature for $C^1$ surfaces}
\label{sec:mc}
Given a surface $\Sigma\subset M$ of class $C^1$ such that the vector fields $Z$ and $\nu_h$ are of class $C^1$ along the characteristic curves in $\Sigma\setminus\Sigma_0$, we define the \emph{mean curvature} of $\Sigma$ at $p\in \Sigma\setminus\Sigma_0$ by
\begin{equation}
\label{eq:Hstrongsense}
\big(\divv_\Sigma^h(\nu_{h})\big)(p):= -g(\nabla_Z \nu_h,Z)(p).
\end{equation}
This is the standard definition of mean curvature for $C^2$ surfaces, see e.g. \cite[(3.8)]{MR3044134} and the references there. The mean curvature is usually denoted by $H$. The mean curvature depends on the choice of $\nu_{h}$. In case $\Sigma$ is the boundary of a set $E$, we shall always choose $N$ as the inner normal and $\nu_{h}=N_h/|N_h|$.
Using the regularity Theorem~\ref{th:main} we get
\begin{proposition}
\label{prop:1stareaStrong}
Let $\Omega\subset M$ be a domain and $E\subset \Omega$ a set of prescribed mean curvature $f\in C^0(\Omega)$ with $C^1$ boundary $\Sigma$ with $H\in L^1_{loc}(\Sigma)$. Then the first variation of the functional \eqref{eq:prescribedfunctional} induced by a vector field $U\in C^1_0(\Omega)$ is given by
\begin{equation}
\label{eq:1stvariationprescribedfunctional}
\begin{split}
\int_{\Sigma}H\, g(U,N) \,d\Sigma-\int_\Sigma f \,g(U,N)\,d\Sigma.
\end{split}
\end{equation}
\end{proposition}
\begin{proof}
We first observe that formula
\[
\frac{d}{ds}\bigg|_{s=0} A(\Sigma_s)=-\int_{\Sigma}H\, g(U,N) \,d\Sigma
\]
can be proved as in \cite[Prop.~6.3]{MR3044134}, see also \cite[Remark~6.4]{MR3044134}. On the other hand, it is well-known that
\[
\frac{d}{ds}\bigg|_{s=0} \bigg(\int_{\phi_s(\Omega)} f \bigg)= -\int_\Sigma f \,g(U,N)\, d\Sigma,
\]
see e.g. \cite[17.8]{MR2976521}.
\end{proof}
Then we have that the mean curvature $H$ defined in \eqref{eq:Hstrongsense} coincides with the prescribed mean curvature $f$
\begin{corollary} Let $E\subset\Omega$ be a set of prescribed mean curvature $f\in C^0(\Omega)$ with $C^1$ boundary $\Sigma$ in a domain $\Omega\subset M$. Assume $H\in L^1_{loc}(\Sigma)$. Then $H(p)=f(p)$ for any $p\in\Sigma\setminus\Sigma_0$.
\end{corollary}
Finally we can improve the regularity of Theorem~\ref{th:main} assuming the mean curvature function is more regular. This result is specially useful if we assume that higher order horizontal derivatives of the function $f$ exist and are continuous.
\begin{proposition}
\label{prop:C^k}
Let $E\subset\Omega$ be a set of prescribed mean curvature $f\in C^0(\Omega)$ with $C^1$~boundary $\Sigma$ in a domain $\Omega\subset M$. Assume that $f$ is also $C^k$ in the $Z$-direction, $k\ge 1$. Then the characteristic curves of $\Sigma$ are of class $C^{k+2}$ in $\Sigma\setminus\Sigma_0$.
\end{proposition}
\begin{proof} Since we have $\nabla_Z Z= H \nu_h$, we can write
\[
\nabla_Z(\nabla_Z Z)= Z(H) \nu_h+H \nabla_Z \nu_h= Z(H)\nu_h-H^2 Z.
\]
Iterating the procedure we obtain the statement.
\end{proof}
In particular, this result holds when $f\in C_\mathbb{H}^1(\Omega)$, i.e., when $f$ has horizontal derivatives of class $(k-1)$, see \cite{MR1871966}.
An important particular case is that of critical points of perimeter, possibly under a volume constraint. Assuming $C^1$ regularity of the boundary, these sets are known to have constant prescribed mean curvature from the discussion in Section~\ref{sec:pmc}. From Proposition~\ref{prop:C^k}, we immediately obtain
\begin{proposition}
\label{prop:critical}
Let $E\subset\Omega$ be either a critical point of the sub-Riemannian perimeter or a critical point of the sub-Riemannian perimeter under a volume constraint. If $E$ has $C^1$ boundary, then the regular part of $\partial E$ is foliated by $C^\infty$ characteristic curves.
\end{proposition}
|
\section{Introduction}
Early-Type galaxies (ETGs) are embedded in hot gaseous haloes produced
mainly by stellar winds, and heated to X-ray temperatures by Type Ia
supernovae (SNIa) explosions and by the thermalization of stellar
motions \citep{fabbiano1989, o'sullivan2001, ciottietal1991,
david.etal1991}. The thermalization is due to the interaction
between the stellar and SNIa ejecta and the pre-existing hot ISM
\citep[e.g.,][]{mathews1989, parriott.bregman2008}. Recent high resolution
2D hydrodynamical simulations of hot gas flows (\citealt{negri.etal2014}; \citealt{negri.etal2014b}, hereafter N14) showed that the presence of ordered
rotation in
the stellar component can alter significantly the ISM evolution with
respect to that shown by fully velocity dispersion supported systems of
same total mass and mass distribution.
Firstly, it is found that the rotation field of the ISM in rotating galaxies is very
similar to that of the stars, with a consequent negligible heating
contribution from thermalization of the ordered motions. Secondly,
conservation of angular momentum in the ISM of rotating galaxies
results in the formation of a centrifugally supported cold equatorial disc, with the consequent reduction
of both the X-ray luminosity $\Lx$ and temperature $\Tx$ of the hot
ISM. These results compared well with observations, which show a
dependence of $\Lx$ and $\Tx$ on the galactic shape and internal
dynamics: $\Lx$ is observed to be high only in round and slowly
rotating galaxies, and is limited to lower values for flatter, fast
rotating ones \citep{eskridge.etal1995, pellegrini.etal1997, sarzi.etal2010,
pellegrini2012, li.etal2011b, sarzi.etal.2013}; $\Tx$ of slowly
rotating systems is consistent just with the thermalization of the
stellar random kinetic energy, estimated from $\sigma_{\mathrm e}$ (the stellar velocity dispersion averaged within one effective radius $R_{\rm e}$),
while fast rotating systems show $\Tx$ values below 0.4~keV, and not
scaling with $\sigma _{\mathrm e}$ (\citealt{sarzi.etal.2013};
see also \citealt{pellegrini2011, posacki.etal2013}).
A major outcome of the previous simulations is that cold material can
be accumulated in considerable amounts, during the lifetime of rotating
galaxies. The cold gas typically settles in the equatorial plane,
where it forms an extended disc (of 0.5--10 kpc radius), that can be as
massive as $\simeq 10^{10}$ M$_{\sun}$, in the largest galaxies. This
result brings in a few important questions: are these cold discs
observed? do they become, as seems natural, a site for star formation
(hereafter SF)? is SF observed? and what is the impact of SF in the disc on $\Lx$ and $\Tx$?
In recent years evidence
has been accumulating that ETGs host significant quantities of cold
gas, in the form of atomic and molecular hydrogen
\citep{morganti.etal2006, combes.etal2007, diseregoalighieri2007,
grossi.etal2009, young.etal2011}; approximately 50\% of massive ETGs
(of stellar mass $M_*\ga 10^{10}$M$_{\sun}$) contain $10^7$ to $10^9$
M$_{\sun}$ of HI and/or H$_2$ \citep{young.etal2014, serra.etal2012,
serra.etal2014, davis.etal2011, davis.etal2013}. A large, systematic
investigation of
the ATLAS$^\mathrm{3D}$ sample of ETGs with the Westerbork Synthesis Radio
Telescope found that $\simeq 40$\% of galaxies outside Virgo, and $\simeq
10$\% of galaxies inside it, are detected in HI, with $M_{\rm HI}\ga
10^7$M$_{\sun}$; the majority (2/3) of the detections consists of
settled configurations, where the cold gas is in discs or rings. Small
discs (size of a few kpc), confined within the stellar body, share the
same kinematics of the stars; large discs (up to $5\times 10^9$
M$_{\sun}$) extend to tens of kpc, and in half of the cases are
kinematically misaligned with the stars \citep{serra.etal2012}. In
particular, fast-rotating Virgo galaxies have kinematically aligned
gas, and the most massive ($M_*\ga 8\times 10^{10}$M$_{\sun}$)
fast-rotating ETGs always have kinematically aligned gas, independent
of environment; this alignment leads to hypothesize that the gas has
been internally generated \citep{davis.etal2011, davis.etal2013}. The HI
discs/rings around slowly-rotating ETGs, instead, are usually not
fully settled, which suggests an external origin (in mergers, or in
accretion from satellite galaxies, or from the intergalactic medium);
an external origin was considered likely also for those ETGs showing
stellar/gas kinematic misalignment. Interestingly, while HI is more
ubiquitous, molecular gas is detected only in fast rotators across the
entire ATLAS$^{\rm 3D}$ sample \citep{young.etal2011, young.etal2014,
davis.etal2011, davis.etal2013}.
Besides giving indication about its origin through its morphology and
kinematics, the observed cold gas seems also to provide material for
SF. Low level SF activity is present in $\sim $ one-third of ETGs
\citep{yi.etal2005, suh.etal2010, ko.etal2014}; \citet{sarzi.etal2006}
showed ongoing SF signatures in the optical spectra of at least $\sim
10$\% of nearby ETGs. In the ATLAS$^\mathrm{3D}$ sample, galaxies with HI
within $\sim$ 1 $R_\mathrm{e}$ exhibit ongoing SF in $\sim 70$\% of
the cases, $\sim 5$ times more frequently than galaxies without HI
\citep{serra.etal2012}. Interestingly, as for molecular gas,
some degree of SF and young stellar populations are detected only in
fast rotators, in the ATLAS$^\mathrm{3D}$ sample \citep{kuntschner.etal2010,
sarzi.etal.2013}. Integral-field spectroscopy showed that ETGs host
frequently a rotating stellar component younger and more metal rich
than the bulge \citep{krajnovic.etal2008}.
The presence of this component, and the occurrence of SF, imply an important role for the cold gas during the evolution of ETGs \citep{khochfar.etal2011, naab.etal2014, cappellari.etal2013}.
Finally, the cold gas content of ETGs is also an
important prediction of $\Lambda$CDM hydrodynamical simulations of
galaxy formation, that include cold gas evolution
\citep[e.g.,][]{martig.etal2013}, and
accretion via various processes during the secular evolution of galaxies
(\citealt{oser.etal2010}; see also \citealt{lagos.etal2014};
\citealt{dubois.etal2013}).
In conclusion, the cold gas has become a tool to gain
insight into recent (and less recent) galaxy evolution. In order to
correctly interpret the variety of observational results, and to use
them properly as constraints for different scenarios for the origin of
the structure and SF history of ETGs, it is crucial to
establish what is the relative importance of the various gas
production processes (internal and external to galaxies), gas
depletion ones (AGN and SF-driven outflows, environmental stripping),
and gas consumption ones (SF). Cold gas is usually thought to come
from accretion from the surrounding medium or satellites, as well as
from gas-rich mergers; in the cases of the giant central-dominant
galaxies in groups or clusters it can also come from cooling of hot
gas \citep{edge.etal2010, mcdonald.etal2011}. The numerical
investigation of N14 showed an additional {\it internal } contribution
of cold gas, coming from the evolution of the passive stellar
population, that can be substantial, or even too large with respect to
observed values. A natural sequel of the N14 work should address
the questions of whether this gas can possibly lead to SF, when SF
takes place, and what is the fate of the cold discs, whether they are
consumed or they are continuously replenished by cooling hot gas. In
this paper we add SF to the simulations of N14, and we explore the ISM
evolution including the removal of cold gas, and the injection of
mass, momentum and energy appropriate for the newly (and continuously)
forming stellar population. In this way we aim at establishing whether
i) the N14 results for the general trends of the hot gas properties
with galaxy shape and stellar kinematics still hold; ii) the formation
of stars can reduce the amount of cold gas in the simulations, thus
bringing it more in agreement with observed values; iii) a significant
channel for SF, previously neglected, should be taken into
consideration for rotating systems, and whether this can account for
the low level SF activity currently seen to be ongoing.
This paper is organized as follows. In Section 2 we describe the main
ingredients of the simulations, such as the galaxy models and the
input physics. In Section 3 we present and discuss the results of the
simulations. In Section 4 we summarize the conclusions.
\section{The simulations}
N14 performed a large set of 2D hydrodynamical simulations with the
ZEUS MP2 code to fully explore the large parameter space of realistic
(axisymmetric) galaxy models, characterized by different stellar mass,
intrinsic flattening, distribution of dark matter, and internal
kinematics. The galaxy flattening was either fully supported by ordered rotation,
originating the set of models that are isotropic rotators, or by
tangential anisotropy, originating the set of fully velocity
dispersion supported models. These two extreme configurations were built adopting the Satoh
decomposition, respectively with Satoh parameter $k=1$ and $k=0$.
The galaxy models were tailored to
reproduce the observed properties and scaling laws of early type
galaxies \citep[see also][]{posacki.etal2013}. In this work we perform
hydrodynamical simulations for a representative subset of rotating
models already investigated by N14 (Sect. 2.1), including SF in the
simulations (as described in Sect. 2.2 below), but keeping the code
(numerical set-up, grid properties) in all equal to that used by N14 (see N14 for details).
Also, a less extreme value of $k=0.1$ is explored.
\referee{A logarithmically spaced numerical mesh $(R,z)$ of
960$\times$480 gridpoints is employed,
with a resolution of 90 pc in the first 10 kpc from the centre.}
\renewcommand\arraystretch{1.4}
\begin{table*}
\caption{Main properties of the galaxy models.}
\begin{tabular}{ccccccccc}
\toprule
Name &$L_B$ &$R_{\rm e}$ &$M_*$ &$M_\mathrm{h}$ & $\sigma_{\rm e8}$ &$f_{\rm DM}$ & $c$ \\
&$(10^{11}$L$_{\sun})$& (kpc) &$(10^{11}$M$_{\sun})$&$(10^{11}$M$_{\sun})$ &(km s$^{-1}$)& & \\
(1) & (2) & (3) & (4) & (5) & (6) & (7) & (8) \\
\midrule
EO4$^{200}_{\rm{IS}}$ &0.26 &4.09 &1.25 &25 & 166 & 0.63 & 37 \\
EO7$^{200}_{\rm{IS}}$ &0.26 &4.09 &1.25 &25 & 124 & 0.66 & 37 \\
FO4$^{200}_{\rm{IS}}$ &0.26 &4.09 &1.25 &25 & 178 & 0.59 & 37 \\
FO7$^{200}_{\rm{IS}}$ &0.26 &4.09 &1.25 &25 & 150 & 0.57 & 37 \\
\midrule
EO4$^{250}_{\rm{IS}}$ &0.62 &7.04 &3.35 &67 & 207 & 0.62 & 28 \\
EO7$^{250}_{\rm{IS}}$ &0.62 &7.04 &3.35 &67 & 154 & 0.66 & 28 \\
\midrule
EO4$^{300}_{\rm{IS}}$ &1.32 &11.79 &7.80 &160 & 248 & 0.64 & 22 \\
EO7$^{300}_{\rm{IS}}$ &1.32 &11.79 &7.80 &160 & 185 & 0.68 & 22 \\
FO4$^{300}_{\rm{IS}}$ &1.32 &11.79 &7.80 &160 & 266 & 0.60 & 22 \\
FO7$^{300}_{\rm{IS}}$ &1.32 &11.79 &7.80 &160 & 224 & 0.59 & 22 \\
\bottomrule
\end{tabular}
\flushleft
\textit{Notes}. $(1)$ Model name: ``EO'' or ``FO'' indicate the procedure (EO-building or FO-building)
applied to the spherical (E0) model to obtain the shape indicated by the number (E4 or E7 shapes);
the superscript indicates $\sigma_{\rm e8}$ of the corresponding E0 model;
the subscript ``IS'' indicates the kinematical configuration of the isotropic rotator
(for example, FO4$^{200}_{\rm{IS}}$ is an E4 isotropic rotator, FO-built from a spherical model with $\sigma_{\rm e8}=200$ km s$^{-1}$).
$(2)$ Luminosities in the $B$ band.
$(3)$ Effective radius (for a FO view for FO-built models, and an EO view for
EO-built models).
$(4)$ Total stellar mass of the original (old) stellar population. The first four models are LM models, the next two are intermediate mass models,
the last four are HM models.
$(5)$ Total dark matter mass.
$(6)$ Luminosity-weighted average of the stellar velocity dispersion
within a circular aperture of radius $R_{\rm e/8}$;
for non-spherical models, $\sigma_{\rm e8}$ is the edge-on viewed value.
$(7)$ Dark matter fraction enclosed within a sphere of radius $R_{\rm e}$.
$(8)$ Concentration parameter of the dark matter profile. \\
Note: the properties listed above, for the EO4$^{250}$ and EO7$^{250}$ models with $k=0.1$, are not reported, since they are equal
to those of the EO4$^{250}_{\rm {IS}}$
and EO7$^{250}_{\rm {IS}}$ ones, except for $\sigma_{e8}$, which is respectively 223 km s$^{-1}$ for the EO4$^{250}_{k\rm{=0.1}}$ model,
and 184 km s$^{-1}$ for the EO7$^{250}_{k\rm{=0.1}}$ model.
\label{tab1}
\end{table*}
\renewcommand\arraystretch{1.}
\subsection{The galaxy models}
N14 built axisymmetric two-component galaxy models where the stellar
component has two different intrinsic flattening, corresponding to the
E4 and E7 shapes, while the dark matter halo is spherical. The
luminous matter is described by the deprojection \citep{mellier.etal1987} of the
\citet{devaucouleurs.1948} law, generalized for ellipsoidal
axisymmetric distributions; the dark matter profile is the
\citep{navarro.etal1997} one, with the dark mass $M_\mathrm{h}$ amounting at $\simeq 20$
times the total
stellar mass $M_*$. For any fixed galaxy mass and shape (E4 or E7), N14 built two models:
the first one, called ``FO-built'', when seen
face-on has the same $R_{\rm e}$ of the spherical E0
counterpart, thus its stellar mass distribution becomes more and more
concentrated than in the E0 model, as it gets flatter; the second one, called ``EO-built'',
when seen edge-on has the same circularized $R_{\rm e}$ of the E0
counterpart, which makes its stellar mass distribution to
expand with increasing flattening.
In this work we re-simulate a few flat (E4 and E7) rotating models
(isotropic rotators, with $k=1$) of N14 including SF in them. In order to
explore the effects of SF at the high and low ends of the galaxy mass
range explored by N14, we choose four flat models with luminosity-weighted stellar
velocity dispersion within $R_{\rm e/8}$ of $\sigma_{\rm e8} = 300$ km
s$^{-1}$ for the parent spherical model (and we call these high-mass models, ``HM'' models),
and four with $\sigma_{\rm e8} = 200$ km s$^{-1}$ for the E0 counterpart (low-mass models, ``LM'' models). In N14, the first
set was found to host inflows, with the creation of a massive,
centrifugally supported cold gaseous disc, while the second set was found in
a global wind, or close to the transition to it.
\referee{In a global wind the gas has very low density and positive velocity (directed outwards) through most of the galaxy
(e.g., \citealt{mathews.baker1971})\footnote{ In fact the phase of the gas flow (that can range from a wind to an inflow) is basically determined by the
different relative importance of SN heating in galaxies of different mass, as already
thoroughly discussed in \citealt{ciottietal1991}.}.}
In addition, we
re-simulated two intermediate-mass models (again E4 and E7), with
$\sigma_{\rm e8} = 250$ km s$^{-1}$ for the E0 counterpart. For these two models,
we also built a moderately rotating stellar kinematical configuration ($k=0.1$), not explored by N14.
The main structural properties of the ten re-simulated
models, identical to those presented in N14, are listed in Tab. 1;
the two new models with $k=0.1$ differ
from the intermediate-mass models only in the $\sigma_{\rm e8}$ value, which is given in the notes to the table.
\subsection{Star formation in the code}
We employed two different schemes for SF, a passive (pSF) one and an
active (aSF) one. In the pSF, only cold gas removal from the
numerical mesh is allowed; in the aSF, we also consider the injection
of mass, momentum and energy from the newly forming stellar
population, for simplicity limiting in this work to the evolution of stars more massive than
8~M$_{\sun}$ (ending with SNII explosions); note however that for reasonable stellar initial mass
functions (IMF) these massive stars are major contributors to the total mass return rate.
We describe below the
main inputs and sinks of mass and energy for the gas flow, due to both the original
stellar population (of total mass $M_*$) and the newly forming stars, and the
corresponding equations of hydrodynamics solved by the code.
\subsubsection{The mass injection and sink terms}\label{mass1}
The mass inputs are stellar winds and SNIa's ejecta produced during
the passive evolution of the original stellar population of the
galaxy (at a rate per unit volume respectively of $\dot\rho_*$ and
$\dot\rho_{\rm Ia}$, for which we adopt the standard recipes coming from
the stellar evolution theory; e.g., N14), and the Type II
supernovae ejecta produced by the newly born stellar population (at a
rate of $\dot\rho_{\rm II}$, calculated as detailed below).
The mass sink is due to SF subtracting gas from the grid, at an
adopted rate per unit volume of:
\begin{equation}
\dot\rho_{\rm SF} = \dfrac{\eta_{\rm SF} \rho}{t_{\rm SF}},\qquad
t_{\rm SF}=\mathrm{max}(t_\mathrm{cool},t_\mathrm{dyn}), \label{eq:sf}
\end{equation}
where $\rho$ is the gas density, and
$\eta_{\rm SF}$ is the SF efficiency, for which we adopt two values of $\eta_{\rm SF}=0.01$
and $0.1$. The rate $\dot\rho_{\rm SF}$ depends on the maximum between the cooling
timescale $t_\mathrm{cool}\equiv E/\cal{L}$ (where $E$ is the ISM internal
energy density, and $\cal{L}$ is the ISM bolometric luminosity per unit volume),
and the dynamical timescale $t_\mathrm{dyn} \equiv \sqrt{ 3\upi/32 G\rho} $
\citep[see also][]{ciotti.ostriker2007}.
When SF takes place in a computational cell, the ISM is removed and an
equal mass in stars $\Delta M_*$ appears at the same place. These new
stars are assumed to form with a Salpeter IMF, thus for a given
$\Delta M_*$, the number of stars having a mass greater than
8~M$_{\sun}$, that will explode as SNII, is $ N_{\rm II}\simeq 7\times
10^{-3} \Delta M_*($M$_{\sun})$. These new stars in turn inject mass into the
ISM. By integration of the mass difference between the mass of the
progenitor and that of the remnant, for the Salpeter IMF, one finds that the mass
injected by SNII's is $\simeq 0.2 \Delta M_*$. The final SNII mass
source term $\dot \rho_{\rm II}$ comes from considering that a given SF
episode generates SNII's that inject mass at a rate exponentially
declining on a timescale $\tau_{\rm II} = 2\times 10^7$ yr, and that at a
certain time during the evolution of a SF episode, another episode may
take place, forming younger SNII's that in turn eject mass into the
ISM. The whole process can be formalized as:
\begin{equation}
\dfrac{\diff \dot\rho_{\rm II}}{\diff t} = -\dfrac{\dot\rho_{\rm II}}{\tau_{\rm II}
} +
\dfrac{0.2 \dot\rho_{\rm SF}}{\tau_{\rm II}}, \label{rhoII}
\end{equation}
\citep{ciotti.ostriker2007}. The equation above is very useful in numerical works, as it allows to compute the mass return
of the new stars formed at each timestep
without storing the whole SF history at each gridpoint, but only the current value and the value of the mass return at the previous timestep.
For more complex recipes, built on the scheme above, see
\citet{calura.etal2014}.
\subsubsection{The energy injection and sink terms}
Energy is injected into the ISM by the thermalization of the kinetic energy of SNIa and SNII explosions, and
by the thermalization of random and streaming stellar motions for the stellar winds.
The rate of SNIa's explosions is the same
entering $\dot\rho_\mathrm{Ia}$ defined in Sect.~\ref{mass1}.
The energy input rate per unit volume from the original (old) stellar population is then:
\begin{equation}
{\dot E}_{\mathrm{old}} = {\dot E}_{\mathrm{Ia}} + \dfrac{\dot\rho_{\rm
Ia}+\dot\rho_*}{2} \left[
\norma{\vvphi {\ephi} -{\boldsymbol{u}}}^2 + \Tr ({\boldsymbol{\sigma}}^2)
\right],\label{eold}
\end{equation}
where ${\dot E}_{\mathrm{Ia}} = \dot\rho_\mathrm{Ia}\vartheta_\mathrm{SNIa} E_\mathrm{SN}/{\rm 1.4~M_{\sun}} $,
$\vartheta_\mathrm{SNIa}$ is the thermalization efficiency \referee{\citep[for which we adopt the value of 0.85, see ][]{thornton.etal1998, tang.wang2005}}, $E_\mathrm{SN}=10^{51}$ erg,
and 1.4~M$_{\sun}$ is the mass that is ejected by one SNIa event; $\vvphi$ is the stellar streaming velocity field,
${\boldsymbol{u}}$ is the velocity of the ambient gas,
$\boldsymbol{\sigma}^2$ is the stellar velocity dispersion tensor of the stars.
The energy input rate ${\dot E}_\mathrm{new}$ from the newly born stellar population is derived as follows.
A mass of newly formed stars $\Delta M_*$ injects energy through SNII with an efficiency
\begin{equation}
\varepsilon_\mathrm{II} = \dfrac{N_\mathrm{II} E_\mathrm{SN}}{\Delta M_* c^2}
\simeq 3.9\times 10^{-6}.
\end{equation}
Consistently with eq.~\ref{rhoII}, the SNII energy injection rate ${\dot E}_{\rm II}$
due to the thermalization of the ejecta is given by:
\begin{equation}
\dfrac{d {\dot E}_{\rm II}}{dt} = -\dfrac{\dot E_{\rm II}}{\tau_{\rm II} } +
\dfrac{\varepsilon_\mathrm{II} c^2 \dot\rho_{\rm SF}
\vartheta_\mathrm{SNII}}{\tau_{\rm II}}, \label{eII}
\end{equation}
where we take the fiducial value of $\vartheta_\mathrm{SNII}=\vartheta_\mathrm{SNIa}/5=0.17$
to account for the lower thermalization efficiency of SNII exploding in a cold and dense medium.
We assume that the
new stars inherit the kinematical configuration of the original stellar
component in the place where they are born (i.e. the velocity dispersion and rotation of the new stars are the same as those of
the original stellar distribution in the same place).
\referee{This allows us to treat the heating terms described in eq. (7) below by using the same properties of the old stars. Note however
that this choice is not unreasonable, because the bulk of star formation takes place in the cold and rotationally supported disc,
whose rotational properties are very similar to those of the stars.}
Thus, the total heating rate per unit volume due to Type II SNe is
\begin{equation}
\dot E_\mathrm{new} = {\dot E}_{\rm II}
+ \dfrac{\dot\rho_{\rm II}}{2} \left[ \norma{\vvphi \ephi -{\boldsymbol{u}}}^2 +
\Tr ({\boldsymbol{\sigma}}^2)
\right].
\end{equation}
The total energy injection due to the old and new stellar populations is then
$\dot E = \dot E_\mathrm{old} + \dot E_\mathrm{new}$, that is:
\begin{equation}
\dot E = {\dot E}_{\mathrm{Ia}} +
{\dot E}_{\mathrm{II}} + \dfrac{\dot\rho_{\mathrm{Ia}}+\dot\rho_* + \dot\rho
_{\mathrm{II}}}{2} \left[
\norma{\vvphi {\ephi} -{\boldsymbol{u}}}^2 + \Tr ({\boldsymbol{\sigma}}^2)
\right] . \label{edot}
\end{equation}
Finally, \referee{an energy and momentum sink associated with SF are present;
they are respectively written as}:
\begin{equation}
{\dot E}_{\rm SF} = \dfrac{\eta_{\rm SF} E}{t_{\rm SF} }, \quad\quad\quad
{\dot{{\boldsymbol{m}}}_\mathrm{SF}} = \dfrac{\eta_{\rm SF}
{\boldsymbol{m}}}{t_{\rm SF} }, \label{eq:star_form_edot}
\end{equation}
\referee{where $E$ and $\boldsymbol{m}$ are the internal energy and momentum density of
the ISM.}
\subsubsection{The hydrodynamical equations}
The hydrodynamical equations are the same as in N14, with the addition
of all the source and sink terms related with the SF process described above:
\begin{gather*}
\dtpartial{\rho} + \diver (\rho {\boldsymbol{u}}) = \dot\rho_\mathrm{Ia} +
\dot\rho_* +
\dot\rho_{\rm II} - \dot\rho_{\rm SF},\label{eq:uff} \\[2ex]
\begin{split}
\rho \dtpartial{\boldsymbol{u}} + \rho
\convective{\boldsymbol{u}}{\boldsymbol{u}} = &-\grad p
-\rho\grad\Phi_\mathrm{tot} +\\
&+ (\dot\rho_\mathrm{Ia} + \dot\rho_* +\dot\rho_{\rm II} ) (\vvphi \ephi -
{\boldsymbol{u}}) ,
\end{split}\\[2ex]
\dtpartial{E} + \diver (E{\boldsymbol{u}}) = -p \diver {\boldsymbol{u}} -
{\mathcal{L}}
+ \dot E -\dot E_\mathrm{SF}, \label{eq:hydroeqs}
\end{gather*}
where $\rho$, ${\boldsymbol{u}}$, $E$, $p$, $\Phi_{\rm tot}$, and
${\cal L}$ are respectively the ISM mass density, velocity, internal energy
density, pressure, total gravitational potential, and bolometric luminosity
per unit volume. The gas is assumed to be an ideal
monoatomic fully ionized plasma, so that $p = (\gamma -1)E$, where $\gamma =
5/3$ is the adiabatic index. The chemical composition is fixed to solar
($\mu\simeq 0.62$), and the gas self-gravity is neglected.
In the pSF scheme, the source terms $\dot\rho_{\rm II}$ and ${\dot
E}_{\mathrm{II}}$
are zero, thus they either do not enter the hydrodynamical
equations above and they do not contribute to $\dot E$. In the more realistic
aSF, $\dot\rho_{\rm II}$ and ${\dot E}_{\mathrm{II}}$ are non-zero.
Most galaxy models are simulated with pSF and aSF; for each of the two schemes, we adopted
two values of the SF efficiency in eqs.~\ref{eq:sf} and~\ref{eq:star_form_edot},
$\eta_{\rm SF}=10^{-1}$ and $\eta_{\rm SF}=10^{-2}$ (see Sect. 3 below).
\begin{figure*}
\centering
\includegraphics[width=0.95\linewidth,keepaspectratio]{fig1}
\caption{Top panels: ISM X-ray luminosity $\Lx$ in the 0.3--8 keV band at 13 Gyr
as a function of $\sigma_{\rm e8}$, for the selection of ten models from N14 re-simulated here
(Tab.~\ref{tab1}), plus two models with $\sigma_{\rm e8}=250$ km s$^{-1}$ for the E0 counterpart,
and $k=0.1$ (shown with thicker ellipses).
The three panels (from left to right) refer to the same models without SF (from N14), with passive SF, and with active SF
(with $\eta_{\rm SF}=0.01$).
Bottom panels: the same for the X-ray emission weighted temperature $\Tx$ in the 0.3--8 keV band at 13 Gyr.
All $L_X$ and $\Tx$ values are given in Tabs. \ref{tab2}-\ref{tab5}. See
Sect.~\ref{comp} for more details.}
\label{Lx}
\end{figure*}
\begin{figure}
\centering
\includegraphics[width=\linewidth,keepaspectratio]{fig2_new}
\caption{$\Lx$ as a function of $\Tx$ for aSF models listed in Tables~\ref{tab1}-\ref{tab5} (open ellipses, as in Fig.~\ref{Lx}), compared with \textit{Chandra} data for the hot gas of ETGs taken from \citet{boroson.etal2011} and \citet{kim.fabbiano2015} (grey squares with error bars). From the latter two samples
we selected only ETGs of morphological type later than E3, for a proper comparison with our models.}
\label{fig2_new}
\end{figure}
We tested that the code conserves the total mass and energy; the total
mass at any time $t$ is $M_\mathrm{gas} +M_\mathrm{esc} = M_\mathrm{inj}
+M_\mathrm{inj}^{\rm II}-M_*^{\rm new}$, where $M_\mathrm{gas}$ is the gas mass within the
simulation box, $M_\mathrm{esc}$ is the cumulative escaped mass out of the
simulation box, until that time; $M_\mathrm{inj}$, $M_\mathrm{inj}^{\rm II}$, and
$M_*^{\rm new}$ are the cumulative masses respectively injected by the
passively evolving stellar population, injected by SNII events due to
the newly born stellar population, and of the new stars. The values
of these masses at the end of the simulations are given in Tabs.
\ref{tab2}-\ref{tab5}.
Finally, the radiative cooling is implemented by adopting a modified
version of the cooling law of \citet{sazonov.etal2005},
with
a lower limit for the ISM temperature of $T = 10^4$ K.
The total X-ray emission in the 0.3–8 keV
band ($\Lx$), and the emission weighted temperature in the same band ($\Tx$),
are calculated as volume integrals over the whole computational grid,
using as weight the emissivity in the
0.3–8 keV band of a hot, collisionally ionized plasma (see N14 for more details).
\section{Results}
Each simulation starts with the galaxy empty of gas, when the age of
the original stellar population is 2 Gyr, and the ISM evolution is
followed for the next 11 Gyr. Tables \ref{tab2}-\ref{tab5} list for all models
the main
quantities of interest at the end of the simulations; the
tables also list the same quantities for the corresponding models
without SF taken from N14 (hereafter ``noSF'' models). First
(Sects.~\ref{comp}) we present the main results focussing on the two
sets of re-simulated models with largely differing values for the galaxy mass
(i.e., the two sets with E0 counterpart of $\sigma_{\rm e8} = 200$ and
300 km s$^{-1}$, the LM and HM models). In Sect.~\ref{interm}
we concentrate on the intermediate mass models (run only for the aSF), where this time
$k$ is equal to 1 or lower ($k=0.1$). In Sects.~\ref{mass}
and~\ref{hist} we
discuss the consumption of cold gas mass and the formation of new stars.
Finally, we investigate the relationship
between the adopted recipe for SF (eq.~\ref{eq:sf}) and the
Kennicutt-Schmidt \referee{relation} (Sect.~\ref{KS}).
\begin{figure*}
\includegraphics[angle=90,width=0.83\linewidth,keepaspectratio]{fig2}
\caption{Meridional sections of the heating over cooling time ratio for the LM EO7$^{200}_\mathrm{IS}$ model,
at a selection of representative times (indicated as galaxy ages in the top right of each panel).
The heating time is $t_\mathrm{heat}=E/\dot E$ (where $\dot E$ is the source term given in eq.~\ref{edot}); the cooling time is defined in Sect. 2.2.1.
From left to right, each column refers to the noSF, pSF and aSF (with $\eta_{\rm SF}= 0.1$) models, respectively.
Arrows describe the meridional velocity field; their lenght
is proportional to the modulus of the gas velocity in the ($z,R)$ plane, according to a scale
shown in the bottom left corner. See Sect.~\ref{comp} for more details.}
\label{EO7}
\end{figure*}
\begin{figure*}
\includegraphics[width=0.85\linewidth,keepaspectratio]
{fig3}
\caption{Meridional sections of the heating over cooling time ratio for the HM
EO7$^{300}_\mathrm{IS}$ model. From top to bottom, each row refers to the noSF, pSF and aSF
(with $\eta_\mathrm{SF} = 0.1$) models,
respectively. Arrows describe the velocity field, as in Fig.~\ref{EO7}.
Note the large cold disc, that does not completely disappear with SF. See Sect.~\ref{comp} for more details.}
\label{s300}
\end{figure*}
\subsection{Comparison between noSF and SF models}\label{comp}
In a short summary, for noSF models, N14 found that rotation in LM
galaxies favours the establishment of global winds, with the
consequent reduction of $\Lx$; in medium-to-high mass galaxies the
conservation of angular momentum lowers the hot gas density in the
central galactic region, leading again to a reduction of $\Lx$, and
also of $\Tx$ (because the external and colder regions weight more in
the computation of $\Tx$). In LM galaxies, instead, $\Tx$ can become
higher if rotation triggers a wind, due to the decrease of the ISM
density, and the additional heating due to the high meridional
velocities of the escaping material (see eq.~\ref{edot}).
\begin{figure*}
\subfloat{\includegraphics[width=0.8\linewidth,height=0.35\linewidth]{LxTx_NFWi_200_IS_E7_EO}}\\
\subfloat{\includegraphics[width=0.8\linewidth,height=0.35\linewidth]{LxTx_NFWi_200_IS_E7_FO}}\\
\subfloat{\includegraphics[width=0.8\linewidth,height=0.35\linewidth]{LxTx_NFWi_300_IS_E7_EO}}
\caption{Time evolution of the X-ray luminosity $\Lx$ and X-ray emission
weighted temperature $\Tx$ for the EO7$^{200}_\mathrm{IS}$ (top panels),
FO7$^{200}_\mathrm{IS}$ (middle panels), EO7$^{300}_\mathrm{IS}$ (bottom panels)
models. The black, green and red lines refer to the noSF model, and to the passive and active star formation schemes, respectively, with $\eta_{\rm SF} =
0.1$. The FO7$^{300}_\mathrm{IS}$ model has an evolution almost
identical to that of EO7$^{300}_\mathrm{IS}$ shown here. See Sect.~\ref{comp} for more details.}
\label{LTt}
\end{figure*}
The first important result here is that when SF is added (both in the pSF and aSF modalities), the hot ISM
evolution remains substantially similar to that found by N14 without
SF, for models of the same mass.
As a consequence of the insensitivity of the general behaviour to the addition
of SF, the values of $\Lx$ and $\Tx$ for the models with SF, and
their trends with the main galactic structural properties,
are on average very similar to those found by N14. Therefore, the conclusions of N14 about
the importance of shape and rotation in determining $\Lx$ and $\Tx$,
also as a function of galaxy mass, are confirmed.
Figure~\ref{Lx} demonstrates this
result by plotting the final $\Lx$ and $\Tx$ values for the same rotating galaxy models evolved
with noSF and with SF (with $\eta_{\rm SF}=0.01$): the
distribution of the points in the noSF and SF panels
is very similar (and the analogous figure with the $\eta_{\rm SF}=0.1$ models produces the same conclusions).
In particular, $\Lx$ of HM models is only marginally sensitive to the presence of SF,
while in LM models SF can introduce variations of $\Lx$; however, these keep within
the (large) range of $\Lx$ values already found in N14 without SF. Thus, SF adds another cause of spread for $\Lx$,
at lower galaxy masses.
\referee{Figure~\ref{fig2_new} shows the resulting $\Lx$ vs. $\Tx$ for the aSF models, compared with the same quantities derived
recently from \textit{Chandra} data for the hot gas of ETGs (\citealt{boroson.etal2011} and \citealt{kim.fabbiano2015}).
From the latter two samples
we have selected only ETGs of morphological type later than E3, for a proper comparison with our models.
Overall, the simulation results agree well with the observed X-ray properties of real ETGs (except perhaps for the most X-ray luminous models
that may be slightly hotter than observed).}
The reason for variations in $\Lx$ linked with SF in LM models is
given by anticipations or delays in the most conspicuous features of
their typical flow behaviour: \referee{N14 found (and the present simulations confirm) that the ISM in rotating models experiences
periodic cooling episodes, where the gas injected by the stellar population accumulates until the radiative losses become catastrophic.
In these episodes, the ISM quickly cools, emitting a large part of its internal energy as radiation; peaks in $\Lx$ and throats in $\Tx$
are then produced, as apparent in Fig. 5. The tenuous hot atmosphere left after a major cooling is then replenished by the new mass
injection from the stellar population. In correspondence of the formation of these significant amounts of cold material (with short cooling
times) star formation is enhanced, and peaks in pace with $\Lx$, as discussed in Sect. 3.4.}
On the contrary, very
little variations of the flow due to SF are seen in HM
galaxies. Figures~\ref{EO7}, \ref{s300}, and \ref{LTt} give particular
examples of the general similarity in the hydrodynamical evolution of
noSF and SF models, by plotting in parallel the ISM evolution for the
noSF, pSF and aSF cases; at the same time, the figures point out how
this similarity is very close for HM models (see Fig.~\ref{s300}, for
the HM model EO7$^{300}_{\rm IS}$), while some differences can be
present in LM ones (see Fig.~\ref{EO7}, for the LM model
EO7$^{200}_{\rm IS}$). Notably, the hot ISM evolution in the HM
models is practically identical with and without SF (see also
Fig.~\ref{LTt}): a massive cold disc forms early, and afterwards, even
when it is mostly consumed by SF (see Sect.~\ref{mass} below), around
the disc the hot ISM evolution proceeds almost unaltered by the
energy and mass injection from SNII's. Less massive models, instead,
are sensitive to the inclusion of SF, similarly to what found in
previous studies where they showed to be very sensitive to any change
in the galaxy properties (e.g., the mass distribution, the stellar
population inputs, the stellar kinematics; \citealt{ciottietal1991},
N14). As shown by Figs.~\ref{EO7} and \ref{LTt}, for the
EO7$^{200}_{\rm IS}$ galaxy, at around 4 Gyr the noSF and the pSF
models have already past their major cooling episode (see the peak in
$\Lx$ in Fig.~\ref{LTt}) followed by the onset of a wind, and the
equatorial outflow becomes stronger and stronger thereafter (see how
the purple regions become more and more extended in Fig.~\ref{EO7},
starting from 4.6 Gyr,
and the sharp and steady decline in $\Lx$ in Fig.~\ref{LTt}). The aSF
model, instead, has its major cooling catastrophe much later, as
shown by the prominent dark green region in Fig.~\ref{EO7} at $\simeq
8$ Gyr, \referee{which emits a conspicuous amount of radiation in the X-ray band, thus prompting} the corresponding peak in $\Lx$ in Fig.~\ref{LTt}.
This delay in the major cooling episode, preceding the onset of a wind,
is due to the newly formed stars that inject mass, and most
importantly heat, in the ISM; thanks to this heat the galaxy can
remain hot gas rich until $\simeq 8$ Gyr, when finally the major
cooling takes place. pSF, where mass is not injected instead,
corresponds to an evolution very similar to that without SF.
A complementary illustration of the different behaviour between the HM
and the LM galaxies is given by Fig.~\ref{LTt}, that shows the $\Lx$
and $\Tx$ evolution for the same models of Figs.~\ref{EO7}
and~\ref{s300}, plus another set of three models representative of the
LM class, the FO7$^{200}_{\rm IS}$ ones. In LM models, a number of
cooling episodes (peaks in $\Lx$) typically take place, possibly terminated by a major
one that is followed by the onset of a wind and the clearing of the
ISM from the galaxy (as in N14). The galaxy mass distribution in the FO7$_{\rm IS}^{\rm 200}$
models (middle panels of Fig.~\ref{LTt}), is more concentrated
than in the EO7$^{200}_{\rm IS}$ ones, thus the last major peak in $\Lx$ is delayed
with respect to what happens for the EO7$^{200}_{\rm IS}$ models. Thus,
in the FO7$_{\rm IS}^{\rm 200}$ galaxy, aSF produces a delay in the
major cooling episode that is even longer than for the EO7$_{\rm
IS}^{\rm 200}$ aSF model, to the point that it does not take place
within the present epoch. pSF instead produces its anticipation with
respect to what shown in the noSF case (Fig.~\ref{LTt}), due
to pSF subtracting gas and leaving a lower density region, radiating
less and more easy to push out of the galaxy.
A final, general result is also that the evolution of a model with SF is
more and more different from that of its corresponding noSF model, when
increasing $\eta_{\rm SF}$.
For illustration purposes, Fig.~\ref{newst} shows a meridional section
of the density distribution of the newly formed stars at the end of
the simulations, for the models in Fig.~\ref{LTt}. One can notice the
more extended and massive disc in HM models, where it is also quite
independent of the pSF or aSF scheme. On the contrary, in LM models,
the disc is much less extended, and it can become larger (for the
EO7$^{200}_\mathrm{IS}$ model) or smaller (for the
FO7$^{300}_\mathrm{IS}$ model) when switching from the pSF to the aSF
scheme; this is another evidence of the high sensitivity of the flow
evolution to any change in the input parameters, at low galaxy masses.
\begin{figure*}
\includegraphics[width=0.75\linewidth,keepaspectratio]{fig5}
\caption{Meridional sections of the density distribution of the
new stars formed up to 13 Gyr (in units of M$_{\sun}$ pc$^{-3}$), for the same models
of Figs.~\ref{LTt} and~\ref{Mt}:
EO7$^{200}_\mathrm{IS}$ (left panels), FO7$^{200}_\mathrm{IS}$ (middle panels),
and EO7$^{300}_\mathrm{IS}$ models (right panels). Top and bottom panels refer respectively
to pSF and aSF simulations, with $\eta_{\rm SF}= 0.1$. }
\label{newst}
\end{figure*}
\begin{figure*}
\subfloat{\includegraphics[width=0.65\linewidth,height=0.55\linewidth]{fig6}}
\caption{Time evolution of the X-ray luminosity $\Lx$ and X-ray emission
weighted temperature $\Tx$ for the EO4$^{250}_\mathrm{IS}$ and EO7$^{250}_\mathrm{IS}$ models
(top panels), and for the EO4$^{250}_{k=0.1}$ and EO7$^{250}_{k=0.1}$ models (bottom panels).
For the aSF models, $\eta_{\rm SF} =0.1$. See Sect.~\ref{interm} for more details.}
\label{s250LT}
\end{figure*}
\begin{figure*}
\subfloat{\includegraphics[width=0.8\linewidth,height=0.35\linewidth]{paper3_mass_NFWi_200_IS_E7_EO}}\\
\subfloat{\includegraphics[width=0.8\linewidth,height=0.35\linewidth]{paper3_mass_NFWi_200_IS_E7_FO}}\\
\subfloat{\includegraphics[width=0.8\linewidth,height=0.35\linewidth]{paper3_mass_NFWi_300_IS_E7_EO}}
\caption{Left panels: time evolution of the cold gas mass $M_\mathrm{c}$ (solid lines), and of the cumulative mass in newly formed stars
$M_*^{\rm new}$ (dashed lines). Right panels: time evolution of the SFR rates. The models are the same of Fig.~\ref{LTt}.
The black line refers to the noSF model,
green and red lines refer to pSF and aSF, respectively, with $\eta_{\rm SF} = 0.1$. Note how $M_\mathrm{c}$ is much reduced
in the SF models (green and red solid lines) with respect to the noSF models (black lines), and how at the same time $M_*^{\rm new}$
(dashed lines) reaches close to the noSF $M_\mathrm{c}$ values. See Sect.~\ref{mass} for more details.}
\label{Mt}
\end{figure*}
\subsection{Intermediate mass ETGs}\label{interm}
Here we present the results for two representative models, of E4 and
E7 shapes, with intermediate galaxy mass ($\sigma_{\rm e8}=250$ km
s$^{-1}$ for the E0 counterpart, see Tab.~\ref{tab1}), that have been run with
the same kinematic configuration of the LM and HM models presented
previously ($k=1$, isotropic rotator), and also with a lower rotation ($k=0.1$).
For
these, only aSF has been considered, with $\eta_{\rm SF} = 0.01$ and
$\eta_{\rm SF} = 0.1$. The final values for the most relevant quantities
of these eight models are presented in Tabs. \ref{tab4} and \ref{tab5}.
The E4 models with $k=1$ show the same smooth evolution of
the HM models, i.e., no or very small oscillations in $\Lx$ and $\Tx$
(see the dashed lines in Fig.~\ref{s250LT}, similar to
those in Fig.~\ref{LTt}); in fact, the
evolution of the hot gas is similarly not affected by SF. The E7 models with $k=1$,
instead, show oscillations in $\Lx$ and $\Tx$,
similar to those of the EO7$^{200}_{\rm IS}$ and FO7$^{200}_{\rm IS}$ models, but fewer and less
pronounced (see the solid lines in Fig.~\ref{s250LT}, to be compared
with those in Fig.~\ref{LTt}).
In the aSF model the final $\Lx$ is larger than for the
noSF case, as found for the FO7$^{200}_{\rm IS}$ models (Sect. 3.1).
In models with rotational velocities reduced by a factor of ten
($k=0.1$), $\Lx$ and $\Tx$ increase largely, as expected from the
results of N14 showing a larger $\Lx$ and $\Tx$ in non-rotating
galaxies ($k=0$). Again the E4 models show no oscillations in $\Lx$
and $\Tx$, while the E7 models show numerous small oscillations in
$\Lx$ and $\Tx$, that increase in amplitude with time increasing
(Fig.~\ref{s250LT}). The cold gas disc is much less extended: its
radius is $\la 1$ kpc for $k=0.1$, and $\la 10$ kpc for $k=1$
(slightly larger for flatter shape, and for lower $\eta_{\rm SF}$).
\subsection{Mass exchange between the cold disc and the new stars}\label{mass}
The second main result of the new simulations is that most of the cold
gas disc is consumed by star formation, both for pSF and aSF. This result, of general
validity, is shown for the particular models of Fig.~\ref{LTt} in
Fig.~\ref{Mt}, where the cold gas mass $M_\mathrm{c}$ of the noSF models (black solid line)
is at the end very close to that of the new stars $M_*^{\rm new}$
(dashed lines). For example, the EO7$_\mathrm{IS}^{\rm 200}$ model in the noSF case has a cold disc extending out to a 2.5 kpc radius, of $M_\mathrm{c}=1.6\times
10^9$M$_{\sun}$, while the corresponding pSF model at the end has no
cold disc, and a mass in new stars of $M_*^{\rm new}\simeq 1.7\times
10^9$M$_{\sun}$ (a bit more gas has cooled in the pSF case than in the
noSF case). Similarly, this same galaxy model,
with aSF and $\eta_{\rm SF}=0.1$, at the end has a cold disc of just
$M_\mathrm{c} \simeq 5.7\times 10^7$M$_{\sun}$, and $M_*^{\rm new}\simeq
3.8\times 10^9$M$_{\sun}$.
This $M_*^{\rm new}$ is larger ($\sim$ double)
than for the pSF case, a fact that is partly accounted for by
the injected mass from the new stars in the aSF model\footnote{We
recall that, in the aSF scheme, the mass injected back into the ISM
by the new stars can itself condense and produce new stars, that in turn
contribute to the computation of $M_*^{\rm new}$.}, and mostly by the
delay in the final cooling episode after which a galactic wind is
established (Fig.~\ref{LTt}); this delay keeps the galaxy gas-rich and
starforming for a longer time (which is proved also by the lower
escaped mass $M_\mathrm{esc}$ in the aSF than in the pSF case, see Tab.~\ref{tab3}).
In general, the LM models with SF show a final $M_\mathrm{c}$ that can range
from 0 to $10^9$M$_{\sun}$, but is typically $\la 10^8$M$_{\sun}$
(Fig.~\ref{Mcold} and Tab.~\ref{tab3}). The final mass in the newly formed
stellar disc is of the order of $M_*^{\rm new}\simeq $a few$\times 10^9$~M$_{\sun}$; thus,
the $M_*^{\rm new}$ values reach almost those of the $M_\mathrm{c}$ of the
corresponding
noSF models (Tab.~\ref{tab3}; Fig.~\ref{Mcold}).
temperature below 2$\times 10^4$K.
\begin{figure*}
\subfloat{\includegraphics[width=0.73\linewidth,keepaspectratio]{fig7}}
\caption{Top panels: final values of the cold gas mass,
for the models in Tabs. \ref{tab2}-\ref{tab5} without SF (left panel), and
with aSF, for the two different $\eta_{\rm SF}$ values adopted here.
Middle and bottom panels: final values of the stellar mass in newly formed stars $M_*^{\rm new}$, of the SFR
normalized to the original $M_*$,
and of the mean formation time of the new stars $\langle t\rangle _*^{\rm new}$,
calculated from the beginning of the simulation,
for the same models with aSF in the top panels. All values are given in Tabs.
\ref{tab2}-\ref{tab5}. See Sects.~\ref{mass}
and~\ref{hist} for more details.}
\label{Mcold}
\end{figure*}
Also in the HM models, similarly to what found for the LM models, $M_*^{\rm new}$
at the end of the simulations
is of the order of the $M_\mathrm{c}$ values of the noSF models.
A larger massive disc remains in the HM models, though, even when including SF
(e.g., Fig.~\ref{Mt} and~\ref{Mcold}). Their final $M_\mathrm{c}$ is $\simeq
10^9$ M$_{\sun}$, for
$\eta _{\rm SF}=0.1$ ($\simeq 1/50$ of $M_\mathrm{c}$ without SF), and ($8-12)\times
10^9$ M$_{\sun}$, if $\eta
_{\rm SF}=0.01$ ($M_{c}\simeq 1/5$ of $M_\mathrm{c}$ without SF), independently
of pSF or aSF.
As found for the LM galaxies, in the aSF models
$M_*^{\rm new}$ is larger than in the pSF ones, and can be even larger than
$M_\mathrm{c}$ of the noSF models (Figs.~\ref{Mt} and~\ref{Mcold}). This time, though, the
difference is lower, of $\simeq 20$\%, and corresponds to secondary SF
sustained by the material injected by the new stars. The
final mass in the newly formed stellar disc is of the order of $M_*^{\rm new}\simeq $a
few$\times 10^{10}$ M$_{\sun}$ (Tab.~\ref{tab2}). However, it must be noted
that the structural properties of the HM models may not be very
realistic: ETGs of this mass with the isotropic rotator kinematics are
quite unlikely, since they are not observed to rotate this much
\citep{emsellem.etal2011}. Real massive galaxies will have less rotation
than adopted here with $k=1$, and then they will have less cold
gas and less new stars.
More realistic massive and rotating ETGs are those of Sect.~\ref{interm}, with aSF. For
these, if $k=1$, the final cold gas mass is again largely reduced when
introducing SF, by a factor of $\simeq 50$ (E4) and $\simeq 100$ (E7),
when $\eta_{\rm SF} = 0.1$. At 13 Gyr, $M_\mathrm{c}$ is few$\times
10^9$M$_{\sun}$ for $\eta_{\rm SF}=0.01$, and few$\times
10^8$M$_{\sun}$ for $\eta_{\rm SF}=0.1$; $M_*^{\rm new}$ is $\simeq
2\times 10^{10}$M$_{\sun}$, for both $\eta_{\rm SF}$ values.
If $k=0.1$, the final $M_\mathrm{c}$ is reduced even more than for the $k=1$ models
(Fig.~\ref{Mcold}), for both the E4 and E7 shapes:
$M_\mathrm{c}$ is a few $\times 10^8$M$_{\sun}$ ($\eta_{\rm SF} = 0.01$), and $\la 10^8$M$_{\sun}$ ($\eta_{\rm SF} = 0.1$).
Note that the final cold mass $M_\mathrm{c}$, without SF, is
similarly large for the $k=1$ and $k=0.1$ models (Tabs. \ref{tab4} and
\ref{tab5}; Fig.~\ref{Mcold}), but
it is instead lower for the $k=0.1$ models than for the $k=1$ ones, when adding SF;
one could conclude that in the $k=0.1$ models SF is more efficient, because it takes place
in a cold disc which is smaller and denser, or that the evolution is different, if less gas
overall cooled (the heating was more efficient). The second hypothesis is supported by the finding
that, at the end of the simulations, $M_*^{\rm new}$ is similar for $k=1$ and $k=0.1$
(e.g., $2.4$ and $2.3\times 10^{10}$M$_{\sun}$ for the E4 models, respectively with $k=1$ and $k=0.1$,
for $\eta_{\rm SF}=0.1$), and similarly larger
than $M_\mathrm{c}$ of the corresponding noSF models (e.g.,
$2.0$ and $1.9\times 10^{10}$M$_{\sun}$, for the E4 models respectively with
$k=1$ and $k=0.1$).
In conclusion, SF seems to be an important mechanism to solve the worrysome feature of massive cold
gas discs in rotating ETGs without SF.
In addition, we found the following trends, at the end of the simulations (see also
the top two rows of panels of Fig.~\ref{Mcold}):
$M_\mathrm{c}$ increases with $M_*$, and has a large spread for LM models, that increases for larger $\eta_{\rm SF}$;
$M_\mathrm{c}/M_*$ is roughly constant with $M_*$, with a spread reaching down to values much lower
than this constant for LM models; at any $M_*$, $M_\mathrm{c}$ decreases when $\eta_{\rm SF}$ increases.
$M_*^{\rm new}$ and $M_*^{\rm new}/M_*$ increase with $M_*$, with a spread for LM models, and are
roughly independent of $\eta_{\rm SF}$. This latter result means that more massive ETGs have been overall
more efficient in forming stars from recycled material, considering their past $\simeq 11$ Gyr.
\subsection{The SF history}\label{hist}
SF is not only responsible for the final mass budget of $M_\mathrm{c}$ and $M_*^{\rm new}$, but also
for the rate of SF (SFR), its evolution, and the (fiducial) age of $M_*^{\rm
new}$;
that is, for the SF history that we discuss here.
Star formation shows a different evolution in LM and HM galaxies, in
pace with their different ISM evolution \referee{(see Fig.~\ref{Mt})}. In LM galaxies, where major
cooling episodes take place recursively, the SFR peaks
suddenly in correspondence of these cooling events \referee{due to the increased density of cold gas}, and declines right
after, going eventually to practically zero \referee{either} if \referee{the cold gas reservoir is completely consumed or} the cooling episode is
followed by a global wind (Sect.~\ref{comp}). Thus, in general LM
galaxies have a larger and peaked SFR in their past (reaching 1-2
M$_{\sun}$ yr$^{-1}$), when the rate of stellar mass losses due to
the original stellar population was much larger, and a low SFR ($\la
0.1$M$_{\sun}$ yr$^{-1}$) at the present epoch. Typical values at 13
Gyr can be as low as zero, and at most as high as 0.45 M$_{\sun}$
yr$^{-1}$ (Tab.~\ref{tab3}). These values compare well with the current rates
estimated for the ATLAS$^{\rm 3D}$ sample, that range from $\approx 0.01$ and 3
M$_{\sun}$ yr$^{-1}$,
with a median value of 0.15
M$_{\sun}$ yr$^{-1}$ for ETGs of $\sigma_{\rm e8}$ comparable to
those of the LM models run here \citep{davis.etal2014}. HM galaxies,
instead, show a more regular and steady production of cold gas, and so
is their SFR (Fig.~\ref{Mt}); at 13 Gyr, their SFR is 2-3 M$_{\sun}$
yr$^{-1}$ (Tab.~\ref{tab2}; but we recall these models are not realistic). In
the intermediate mass models (Sect.~\ref{interm}), the SFR is larger
for the E4 models without oscillations in $\Lx$ than for the E7 ones.
Overall, the SFR at 13 Gyr is similar for $k=1$ and $k=0.1$: for
$k=1$, the SFR ranges from 0.42 to 1.1 M$_{\sun}$ yr$^{-1}$ (i.e., it
is larger than for the LM models of same $k=1$); when $k=0.1$, the SFR
ranges from 0.45 to 1.0 M$_{\sun}$ yr$^{-1}$.
Figure~\ref{Mcold} summarizes the {\it present epoch} SFR behaviour, as a function of
$M_*$, for the aSF models.
To compare models of different mass, the plotted quantity is the SFR normalized to $M_*$.
This quantity is
$\sim (1-4)\times 10^{-3}$~Gyr$^{-1}$, quite independent of $\eta_{\rm SF}$, for
the intermediate mass and HM models; it can be largely varying for the
LM models (from $4\times 10^{-4}$ to $6\times 10^{-3}$~Gyr$^{-1}$, for
$\eta_{\rm SF}=0.1$, and from $8\times 10^{-5}$ to $\sim 10^{-3}$~Gyr$^{-1}$, for $\eta_{\rm SF}=0.01$). On average, the present epoch
SFR/$M_*$ increases with $M_*$, a result similar to that
of the previous Sect.~\ref{mass}, where we found that the integration over time of the SFR,
that is $M_*^{\rm new}$,
and $M_*^{\rm new}/M_*$, increase with $M_*$.
Tables \ref{tab2} and \ref{tab3} also list the mean formation time $\langle t\rangle _*^{\rm new}$
of the new stars since the beginning of
the simulation ($t_0=2$ Gyr), and the mean star formation rate (defined as $M_*^{\rm new}/\langle t
\rangle _*^{\rm new}$),
both at 13 Gyr.
$\langle t\rangle _*^{\rm new}$ as a function of time is defined as:
\begin{equation}
\langle t\rangle _*^{\rm new} (t)=\dfrac{1}{M_*^{\rm new}(t)} \int _{t_0} ^t (t^{\prime}-t_0) \, {\rm SFR}(t^{\prime}) \diff t^{\prime},
\end{equation}
where the SFR$(t)$ is the volume integrated, instantaneous star formation rate.
The values of $\langle t\rangle _*^{\rm new}$ at the {\it present epoch} are shown in Fig.~\ref{Mcold}
for the aSF cases. They range from 2.4 to 6.6 Gyr, for the LM models,
with a tendency to be lower for the aSF than for the pSF, and for the larger $\eta_\mathrm{SF}$.
The HM and intermediate mass models have $\langle t\rangle _*^{\rm new}$ in a narrower range (from 4.2 to 5.5 Gyr),
independent of the aSF or pSF, and are again lower for the larger $\eta_{\rm SF}$ (Fig.~\ref{Mcold}).
Therefore, the new stars
in LM models have a larger spread of formation times,
and correspondingly a larger range of ages at the current epoch,
than in larger mass models.
For the aSF, for $\eta_\mathrm{SF}=0.01$, the upper envelope of the more recent
formation times (largest $\langle t\rangle _*^{\rm new}$ values) is $\simeq 5.5$ Gyr,
roughly independent of $M_*$, while
the lower envelope (lowest $\langle t\rangle _*^{\rm new}$ values) increases with $M_*$.
A similar trend is shown by the HM and intermediate mass models,
for $\eta_{\rm SF}=0.1$,
but shifted towards lower $\langle t\rangle _*^{\rm new}$ values ($\la 4.5$ Gyr);
the spread of $\langle t\rangle _*^{\rm new}$ for LM models keeps instead large,
extending to even lower $\langle t\rangle _*^{\rm new}$ values of $\simeq 3$ Gyr.
The decrease of $\langle t\rangle _*^{\rm new}$ for larger $\eta_{\rm SF}$, i.e., for a more efficient SF,
obtained here from hydrodynamical simulations provides support to the similar conclusion
drawn in galaxy formation models focussed on chemical evolution: these
assume that the efficiency of star formation is an increasing function of mass, which for them
has -- among others -- the consequence that smaller galaxies continue to form stars for longer periods, while more
massive galaxies host an older stellar population \citep[e.g.,][]{matteucci1994}.
It is interesting that $\langle t\rangle _*^{\rm new}$ decreases slightly, or remains similar,
when switching from pSF to aSF; in fact, one could expect a more extended
SF, due to the injection of material from the new stars, for the aSF.
Evidently, the net results of the injection of both mass
and energy is to produce the bulk of
SF at roughly the same time, for the pSF and aSF schemes; only, the final mass in new stars is larger for aSF.
The mean age of the newly formed stars in Gyr is 13$-t_0-\langle t\rangle _*^{\rm new}$.
This age is in any case larger than $\sim 5.5$ Gyr, for the aSF.
\begin{figure*}
\includegraphics[width=0.65\linewidth,keepaspectratio]{fig9}
\caption{Projected SFR versus mass surface density of
the gas (at $T<2\times 10^4$K, for the blue symbols, and at $T>2\times 10^4$K,
for the red ones), for two aSF models: EO7$^{200}_{\rm IS}$ on the left, and
EO7$^{250}_{\rm IS}$ on the right.
Full circles refer to $\eta_\mathrm{SF}=0.01$ and crosses to $\eta_\mathrm{SF}=0.1$.
The lower $\eta_{\rm SF}=0.01$ coefficient for the SF
recipe (eq.~\ref{eq:sf}) reproduces better the empirical
Kennicutt-Schmidt scaling, that is shown by the solid line (with a
slope of $N=1.4$, and normalized as in \citet{davis.etal2014} for their ATLAS$^{\rm 3D}$ sample
of molecular gas-rich ETGs). The blue
dotted and dashed lines show the best fits for the cold gas only
(coefficients in the frames in the lower right corners). See
Sect.~\ref{KS} for more details.}
\label{KSfig}
\end{figure*}
\subsection{The SF recipe and the Kennicutt-Schmidt relation}\label{KS}
From eq.~\ref{eq:sf} it follows that SF is favoured in regions of high density, and relatively short $t_\mathrm{SF}$.
These conditions are naturally present in low temperature regions; accordingly, the simulations show that
most of the SF takes place where the gas
temperature is $\la 2\times 10^4$K, i.e., in the cold rotating disc.
Quite obviously, our results depend somehow on the adopted prescription
for the SF, while a realistic, physically consistent,
description of SF is not available. Yet, it is interesting to ask at what level
our implementation of SF recovers some basic observational features of SF.
More than half a century ago \citet{schmidt1959} conjectured that the rate
of SF should vary as a power law of the gas density, and later
\citet{kennicutt1998} suggested the parameterization in terms of mass surface densities $\Sigma_{\rm SFR} =
A\Sigma_\mathrm{cold\, gas}^N$, valid for
starforming galaxies, with $N\simeq 1.4-1.5$ \citep{kennicutt.evans2012}.
We investigate here what scaling (if any) the
adopted SF recipe of eq.~\ref{eq:sf} produces in terms of $\Sigma_{\rm SFR}$
vs. $\Sigma_{\rm gas}$, when implemented in the simulations.
Figure~\ref{KSfig} shows for a few representative models at the end of
the simulation the relationship between $\Sigma_{\rm SFR}$ and
$\Sigma_{\rm gas}$, calculated separately for the hotter (T$> 2\times
10^4$K) and colder (T$\leq 2\times 10^4$K) phases of the gas;
$\Sigma_{\rm SFR}$ is the face-on projection of $\dot\rho_\mathrm{SF}$. The
figure shows for comparison a recent scaling for the Kennicutt-Schmidt
\referee{relation}, derived by \citet{davis.etal2014} for the ATLAS$^{\rm 3D}$ sample of
molecular gas-rich ETGs (this has a slope of $N=1.4$).
Figure~\ref{KSfig} tells that SF in the gas takes place basically in
two regimes, and by far most of it is due to cold gas. The behaviour
of the cold gas shows two remarkable features: the first is that a
scaling similar to that of the Kennicutt-Schmidt \referee{relation} is followed quite
closely by the blue (cold gas) points; the second is that even the
normalization is reasonably well reproduced by our scheme of ``active
SF'', when $\eta_{\rm SF}=0.01$ (models with $\eta_{\rm SF}=0.1$
correspond instead to a normalization that is too large). In
particular, the cold gas in the log$\Sigma_{\rm SFR}$--log$\Sigma_{\rm
gas}$ plane follows a trend best-fitted by a line of slope of $\simeq
1.4-1.5$, a fact that can be explained as follows. SF in the cold gas
is most likely regulated by the dynamical timescale $t_\mathrm{dyn}$, being
$t_\mathrm{cool}$ very short at such temperatures and densities; thus, one
roughly expects that $\dot\rho_{\rm SF} \propto \rho^{1.5}$. A slope
for the $\Sigma_{\rm SFR}$--$\Sigma_{\rm gas}$ relation close to the
empirical one then follows when neglecting the difference between
volume and surface density, which is not unreasonable considering that
the cold disc resides mainly in a thin region above and below the
equatorial plane, extending vertically for just a couple of gridpoints
($< 200$ pc; see also \citealt{kennicutt.evans2012}). For the hotter gas,
the red points in Fig.~\ref{KSfig} are invariably fitted by a steeper
line, with slope $\simeq 2$; this is explained by a longer cooling
time than the dynamical time, and then by $\dot\rho_{\rm SF} \propto
\rho^{2}$.
Overall, Fig.~\ref{KSfig} proves that the recipe of eq.~\ref{eq:sf} is
very reasonable for SF, and, together with the chosen range of
normalization for $\eta_{\rm SF}$, it provides a scaling close to that
observed for normal to starburst galaxies. These considerations add
strength to the overall results obtained in this study.
\section{Conclusions}
In this work we have explored the effects of SF on the evolution of
rotating hot gas flows in early-type galaxies; our previous (N14) high resolution 2D hydrodynamical
simulations of such flows, run for axisymmetric two-component
models, were successful in producing the observed $\Lx$ and $\Tx$
of flat and rotating ETGs, but also revealed the formation of very massive
cooled gas discs. To study the SF effects, we performed
hydrodynamical simulations using the same numerical code of N14, where
SF with a Salpeter IMF is inserted, following a simple recipe
depending on the gas density, and the cooling and dynamical timescales of the ISM
(eq.~\ref{eq:sf}), with two possible values for the efficiency of SF
($\eta_{\rm SF}=0.01$ and 0.1). We considered the new stellar generations
to be passive (a pure sink of gas), or active (the new stars
contribute mass and energy to the gas flow, via stellar winds and
SNIIs). The new simulations have been run for
a subsample of the N14 models, of high, intermediate and low galaxy masses, of E4 and E7 shapes,
with the ordered stellar rotation described by the isotropic rotator, or
with the rotational velocities reduced by a factor of ten.
It is found that subsequent generations of stars are formed from the
cold gas that accumulates in the equatorial
plane, and that most of the extended and massive cold disc found by N14 is consumed
by this process. In particular, the main conclusions are summarized as follows.
$\bullet $ Remarkably, both for the passive and the active SF
implementation, we confirm the results of the previous investigation
without SF, concerning the trends of $\Lx$ and $\Tx$ of the hot ISM with
galactic rotation. As found by N14, $\Lx$ is lowered by rotation: in
low mass ETGs this happens because rotation favours the establishment of global winds,
and in medium-to-high mass ETGs because rotation lowers the hot gas density
in the central galactic region. The average temperature $\Tx$ is also
lower in intermediate/high mass ETGs, when rotation is important, due
to the lower contribution of the central regions (usually hotter and
denser in non-rotating systems), while it can be higher in low mass ETGs, if rotation triggers
a wind, due to the thermalization of meridional winds.
$\bullet $ The robustness of the N14 results even after the addition of SF is explained differently for HM and LM galaxies.
In HM galaxies the evolution of the hot ISM is only marginally sensitive
to the presence of SF, and $\Lx$ and $\Tx$ evolve smoothly as without SF: a
massive cold disc forms early, it is mostly consumed by SF, and around
it the hot ISM evolution proceeds almost unaltered by SF (even in
presence of energy and mass injection by SNII’s). In less massive
models, instead, the evolution of $\Lx$ is not smooth, but shows many peaks
corresponding to major cooling episodes, as in N14. The time-occurrence of the
peaks depends on SF, thus SF can induce variations of $\Lx$ at a
certain chosen epoch; however, these keep within the already large range of $\Lx$ values
typical of these masses, even without SF, being the gas flow very sensitive
to many factors \citep[as already found previously, e.g.,][N14]{ciottietal1991}.
$\bullet $ For what concerns the global gas mass budget, at the end of the simulations the cold gas mass left in the equatorial disc
$M_\mathrm{c}$ on average increases with $M_*$, with a larger spread for LM models; at fixed $M_*$,
$M_\mathrm{c}$ is lower for lower stellar ordered rotation, and for larger $\eta_{\rm SF}$. In addition, the ratio
$M_\mathrm{c}/M_*$ is roughly constant with $M_*$, with a spread reaching down to values much lower
than this constant for LM models.
In any case, $M_\mathrm{c}\la 2\times 10^9$~M$_{\sun}$ for galaxies with $M_*\leq 4\times 10^{11}$~M$_{\sun}$.
Typical values for $M_\mathrm{c}$ are $\la $a few $\times
10^7$~M$_{\sun}$ for LM galaxies (with the maximum rotational level of
$k=1$), with exceptions of $M_\mathrm{c}\simeq 10^9$~M$_{\sun}$; and a few
$\times 10^7$~M$_{\sun}$ (if $k=0.1$) to $3\times 10^9$~M$_{\sun}$ (if
$k=1$) for ETGs of intermediate mass. These values compare well with
those recently observed \citep{young.etal2014, serra.etal2012, serra.etal2014}.
In particular, they can explain the observation that
massive, fast-rotating ETGs often have kinematically aligned gas, independent
of environment \citep{davis.etal2011, davis.etal2013}.
$\bullet $ The mass in newly formed stars $M_*^{\rm new}$, and the ratio $M_*^{\rm new}/M_*$,
increase with $M_*$, again with a large spread for LM models, and are
roughly independent of $\eta_{\rm SF}$. This result means that more massive (rotating) ETGs have been overall
more efficient in forming stars via recycling of their stellar mass losses,
during the past $\simeq 10$ Gyr.
The mass in secondary generations of stars is $(1-6)\times
10^9$~M$_{\sun}$ (if $k=1$) for LM models, and $\simeq 2\times
10^{10}$~M$_{\sun}$ (for both $k=1$ and $k=0.1$) for
intermediate mass models. These should reside mostly in a disc, as most fast rotator ETGs possess,
and be related to the birth of a younger, more metal rich disky stellar component that is indeed
observed \citep{krajnovic.etal2008, cappellari.etal2013}.
They should not be recognized as a ``young'' stellar population in an absolute sense, though,
since most of $M_*^{\rm new}$ formed a few Gyr ago, as can be evaluated from the
evolution of their SFR.
$\bullet $ The time evolution of the SFR depends on the mass of the galaxy.
LM galaxies ($M_*=1.25\times 10^{11}$~M$_{\sun}$) have a larger and
peaked SFR in their far past (reaching 1-2~M$_{\sun}$ yr$^{-1}$),
when the rate of stellar mass losses due to the original stellar
population was much larger, and when major cooling episodes were frequent. In
fact, the average age of the
new stars of LM models ranges from 5 to 8 Gyr (for the aSF scheme). At the present epoch, their
SFR is low ($\la 0.1$~M$_{\sun}$ yr$^{-1}$ typically), with the full
range of values going from zero to $\simeq 0.7$~M$_{\sun}$
yr$^{-1}$. These results agree nicely with the low degree of SF and young
stellar populations that is detected only in fast rotators, in the
ATLAS$^{\rm 3D}$ sample \citep{kuntschner.etal2010, sarzi.etal.2013}; also, they
compare well with the current estimated rates, whose median value is
$\approx 0.15$~M$_{\sun}$ yr$^{-1}$ \citep{davis.etal2014}.
More massive galaxies, instead,
show a more regular and steady production of cold gas, and so is their
SFR; the average age of their new stars
ranges from 5.5 to 7.0 Gyr (in the aSF scheme).
At 13 Gyr the SFR is $\simeq (0.4-1)$~M$_{\sun}$ yr$^{-1}$,
larger than for LM models, and somewhat larger than observed
\citep[e.g.,][]{mcDermid2015}; this could be fixed by
assuming
a lower $\eta_{\rm SF}$, or a lower ordered stellar rotation,
at larger galaxy masses (Fig.~\ref{Mcold}). Thus, our models may still be
rotating too much and producing too much cold gas, at galaxy
masses $>2\times 10^{11}$~M$_{\sun}$.
The current SFR/$M_*$, instead, remains quite constant from LM
to intermediate mass models, with a larger spread for LM models (as already found for other properties).
Finally, the SF recipe adopted for this
work proved to be a reasonable one, given that it reproduces the slope of the
Kennicutt-Schmidt \referee{relation}, and even the normalization if $\eta_{\rm SF}\simeq 0.01$.
Overall, we can suggest an origin (mostly) in the SF from cooling hot gas,
for the presence of cold gas phases
kinematically aligned with the stars, and for the low-level degree of SF, all features detected only in fast
rotators in the ATLAS$^{\rm 3D}$ sample.
In the perspective of the galaxy evolution, we finally recall that
the precise knowledge of the amount of gas flowing towards the galactic centre is of
great importance for a proper study of feedback effects from the
central black hole in rotating galaxies \citep[e.g.,][]{novak.etal2011, gan.etal2014}. A major
implication of the present work is the fact that, if the gas produced
by stellar evolution forms new stars, the amount of gas in principle available for
fuelling the central supermassive black hole is much lowered. Of course, such fuelling would be
impossible in presence of rotation and absence of viscosity.
Disc
instabilities, though, could break axial symmetry, and allow for
the gas infall onto the black hole, possibly escaping SF.
Phenomenologically, the effect of gravitational instabilities in the disc
can be modelled as an effective gravitational viscosity \citep[e.g.,][]{bertin.lodato.2001, hopkins.quataert2011, rafikov2015}, that favours accretion of cold
gas towards the centre. Indeed, viscosity (that could be due for example to MRI)
is another ingredient not present in the current simulations, but that could affect the disc
evolution. As the code is axisymmetric, we cannot follow the
complex physics of (non-axisymmetric) disc instabilities that could be
present.
Although self-gravity of the gas is not considered in our simulations, yet
for some models at a selection of times we computed the fiducial
value of the radial profile of the stability
parameter $Q(R)= c_\mathrm{s} k_R/\upi G\Sigma_\mathrm{c}$, where $k_R$ is the local
epicyclic frequency of the galaxy, and $c_s$ and $\Sigma_\mathrm{c}$ are
respectively the vertically mass-averaged values of the ISM sound
velocity, and the surface density of the cold rotating gas. The $Q$ values were computed
for a layer $\Delta z\simeq 200$ pc thick above and below the galaxy equatorial
plane; these $Q$ turned out to be quite independent of the
adopted thickness value $\Delta z$, being dominated by the cold disc.
These fiducial $Q$ values maintained
invariably larger than unity (indicating stability) when SF is
allowed, while $Q <1$ on a central $\simeq$kpc-size region, in absence of SF. From this preliminary
and qualitative analysis, it follows that SF {\it and} $Q$-instability are both present in
the gaseous rotating discs. The relevant question is then if and what
of the two processes is dominant in the disc. In a future work we plan to extend
further the present investigation by implementing in the code
the cooperating effects of SF and mass discharge
from the disc to the centre. In this new study, we will also consider
the neglected late-time mass return from the newly born stellar
population, that is the mass injected by stars of mass
$<8$~M$_{\sun }$; this input could be modelled with the method described
in \citet{calura.etal2014}.
\section*{Acknowledgements}
We acknowledge Giuseppe Lodato and Jerry Ostriker for useful discussions, and Silvia Posacki for
providing the galaxy models of N14.
L.C. and S.P. were supported by the MIUR grant PRIN 2010-2011, project
`The Chemical and Dynamical Evolution of the Milky Way and Local Group
Galaxies', prot. 2010LY5N2T.
|
\section{Introduction}
In the first chapter we compile the most important results with regard to the effective Lagrangian in a constant electromagnetic field.
Our objective is to find the Green's function of a spin-$\tfrac{1}{2}$ particle in an external constant magnetic field
that points in the $\hat{z}$ direction. This can be achieved with Schwinger's proper-time technique.
With the result we can compute the entire effective Lagrangian as a function of the constant $(E,H)$ field. In this way we obtain the
famous Heisenberg-Euler effective Lagrangian. We will then set up a relation between the effective Lagrangian and the trace
of the energy-momentum tensor for constant magnetic and electric fields. Thereafter we give up the requirement that the fields be constant
and allow for arbitrary varying fields. This is done in connection with the effective action for Yang-Mills fields.
Rather than attempting to compute $\mathscr{L}^{eff}$, we will make an ansatz motivated by the requirement that $\mathscr{L}^{eff}$
give the correct trace anomaly for the energy-momentum tensor. In this way we are able to construct the leading-log effective Lagrangian.
Similar considerations are used to investigate the effective Lagrangian in QED. Finally we briefly study Adler's leading-log model in QCD
and state his result concerning the static potential between a quark-antiquark pair for long and short distances. Although the calculations are
highly non-trivial, the results of the linearly rising potential for large quark separation and the Coulombic $r^{-1}$ potential for small
distances are very encouraging.
\section{Compendium of Useful Formulae\cite{Dittrich1985} }
We start with the Green's function of a spin-$\tfrac{1}{2}$ particle in an external electromagnetic field:
\begin{equation}
\left[ \gamma^\mu\left(\frac{1}{i}\partial_\mu - eA_\mu\right) + m\right]G_+(x,x';A) = \delta(x-x').
\end{equation}
If we pick a special gauge field so that $F_{\mu\nu}$ is constant, we obtain the closed-form solution
\begin{align*}
G_+(x,x';A) = \phi(x,x') &\int_0^\infty \frac{1}{s^2}\left[m-\frac{1}{2}\gamma^\mu\left\lbrace f(s)+ eF\right\rbrace_{\mu\nu}(x-x')^\nu\right]\\
&\times e^{-im^2s-L(s)+\frac{i}{4}(x-x')f(s)(x-x')}e^{\frac{i}{2}\sigma Fs}ds\quad ,
\intertext{with}
f(s) &= eF\coth(eFs)\\
L(s) &= \frac{1}{2}\text{tr} \ln \left[ \frac{\sinh(eFs)}{eFs} \right]\ ,
\end{align*}
and $\phi(x,x') = e^{ie\int_{x'}^x A^\mu(\xi)d\xi_\mu }$ with a straight path between $x$ and $x'$.
Our central subject of interest is the vacuum amplitude in the presence of an external field which,
in the framework of a one-loop approximation for the effective Lagrangian, can be written as
\begin{align}
\langle 0_+\vert0_-\rangle^A &= e^{iW^{(1)}\left[A\right]} = e^{i\int\mathscr{L}^{(1)}(x)d^4x}\quad ,\\
\intertext{with}
iW^{(1)}\left[A\right] &= -\text{Tr}\ln\left(\frac{1}{1-e\gamma AG_+}\right)=-\text{Tr}\ln\left(\frac{G_+\left[A\right]}{G_+\left[0\right]}\right)\ .
\end{align}
Here $G_+=G_+[0]$ is the electron propagator in the field-free case, connected with $G_+[A]$ by
\begin{equation}
G_+[A] = G_+(1-e\gamma AG_+)^{-1}\quad .
\end{equation}
Furthermore, $\text{Tr}$ indicates the trace both in spinor and configuration space.
The one-loop effective action $W^{(1)}$, i.e., the effective Lagrangian $\mathscr{L}^{(1)}$, is the formal expression for the effect which an arbitrary number of
\textquotedblleft external photon lines\textquotedblright\ can have on a single Fermion loop.
The functional derivative with respect to the potential $A_\mu(x)$ is given by
\begin{equation}
i\frac{\delta W^{(1)}[A]}{\delta A_\mu(x)} = -e\text{tr}\left[\gamma^\mu G_+(x,x;A)\right]\quad .
\end{equation}
This equation is fulfilled by the ansatz
\begin{equation}
iW^{(1)} := i\int\mathscr{L}^{(1)} d^4x = -\frac{1}{2}\int_0^\infty \frac{e^{-ism^2}}{s}\text{Tr}\left[e^{is(\gamma\cdot \Pi)^2}\right]ds\ ,
\end{equation}
where the proper-time representation of $G_+[A]$ is given by
\begin{equation}
G_+[A]\cdot \frac{\gamma\cdot \Pi - m}{(\gamma\cdot \Pi)^2 - m^2} = (m-\gamma\cdot \Pi)i\int_0^\infty e^{-is\left[m^2-(\gamma\cdot\Pi)^2\right]}ds\quad .
\end{equation}
We can then write for the unrenormalized Lagrangian
\begin{equation}
\mathscr{L}^{(1)}(x) = \frac{i}{2}\text{tr} \int_0^\infty \frac{e^{-im^2s}}{s}\langle x\vert e^{is(\gamma\cdot\Pi)^2} \vert x\rangle ds\ ,
\end{equation}
where the trace refers only to the spinor index.
With this expression for $\mathscr{L}^{(1)}(x)$ we can show that
\begin{equation}
i\frac{\partial \mathscr{L}^{(1)} }{\partial m}= \text{tr}\ G_+(x,x;A)\ .
\end{equation}
Without further proof we also find for the trace of the energy-momentum tensor
\begin{equation}
\langle T_\mu ^\mu(x)\rangle = -im\ \text{tr}\ G_+(x,x;A)\ .
\end{equation}
This leads us to the equation
\begin{align*}
\langle T_\mu ^\mu(x)\rangle &= -im\text{tr}\langle x\vert (-\gamma\cdot\Pi+m)i\int_0^\infty e^{-i\left(m^2-(\gamma\cdot\Pi)^2\right)s}ds \vert x\rangle\\
&= m^2 \langle x\vert \text{tr} \int_0^\infty e^{-i\left(m^2-(\gamma\cdot\Pi)^2\right)s} ds\vert x\rangle\ .
\end{align*}
With the former expression for $\mathscr{L}^{(1)}$ we obtain the useful equality
\begin{equation}
\langle T_\mu ^\mu(x)\rangle = m\frac{\partial \mathscr{L}^{(1)} (x)}{\partial m} = \frac{\partial \mathscr{L}^{(1)} (x) }{\partial(\ln m)}\quad .
\end{equation}
For a purely constant magnetic field the renormalized one-loop effective Lagrangian is known to be
\begin{equation}
\mathscr{L}^{(1)}(H) = -\frac{1}{8\pi^2}\int_0^\infty \frac{e^{-m^2s}}{s^2}\left[ (eHs)\coth(eHs)-\frac{1}{3}(eHs)^2-1\right]ds\ .
\end{equation}
The integral can be explicitly calculated by dimensional or $\zeta$-function regularization. In the next chapter we will make the explicit expression
for $\mathscr{L}^{(1)}(H)$ the starting point for our detailed discussion of the trace anomaly of the energy-momentum tensor in QED.
\section{The trace anomaly of the energy-momentum tensor from the one-loop effective Lagrangian in QED}
We already mentioned the close connection between the effective Lagrangian $\mathscr{L}^{(1)}$ and the trace of the energy-momentum tensor:
\begin{equation}
\langle T_\mu ^\mu(x)\rangle = m\frac{\partial \mathscr{L}^{(1)} (x) }{\partial m}\ .
\end{equation}
For constant fields we have the expression
\begin{align*}
\langle T_\mu ^\mu(x)\rangle \left(\mathcal{F},\mathcal{G}\right) &= \frac{m^2}{16\pi^2}4\int_0^\infty \frac{e^{-m^2s}}{s^2}\left[ e^2s^2\mathcal{G}
\frac{ \mathfrak{Re}\cosh\left(es\sqrt{2}(\mathcal{F}+i\mathcal{G})^{1/2}\right) }{ \mathfrak{Im}\cosh\left(es\sqrt{2}(\mathcal{F}+i\mathcal{G})^{1/2}\right) } \right.\\
&\left.-1 -\frac{2}{3}e^2s^2\mathcal{F}\right]ds\quad ,
\intertext{where}
\mathcal{F} &= \frac{1}{4}F_{\mu\nu}F^{\mu\nu} = \frac{1}{2}\left( H^2-E^2\right)\ ,\\
\mathcal{G} &= \frac{1}{4}F^{\mu\nu}F_{\mu\nu} = \vec{E}\cdot \vec{H} \quad ,
\intertext{with}
F_{\mu\nu} &= \frac{1}{2}\epsilon_{\mu\nu\kappa\lambda}F^{\kappa\lambda},\quad \epsilon_{0123}=1\ .
\end{align*}
The closed-form expression $\mathscr{L}^{(1)}$ for an external constant $H$-field only is given by
\begin{align*}
\mathscr{L}^{(1)}(H) &= -\frac{1}{32\pi^2}\left[(2m^4-4m^2(eH)+\frac{4}{3}(eH)^2)\left[1+\ln\left(\frac{m^2}{2eH}\right)\right] \right.\\
&\left. +4m^2(eH)-3m^4-(4eH)^2\zeta'\left(-1,\frac{m^2}{2eH}\right)\right]\quad .
\end{align*}
The result of the mass-differentiation turns out to be
\begin{align*}
\langle T_\mu^\mu\rangle(H) &= -\frac{1}{12\pi^2}(eH)^2-\frac{m^4}{4\pi^2}\ln \left( \frac{m^2}{2eH} \right)+\frac{m^2}{4\pi^2}(eH)\ln\left( \frac{m^2}{2eH} \right)\\
& + \frac{m^4}{4\pi^2}+\frac{(eH)m^2}{2\pi^2}\left[\ln\Gamma\left(\frac{m^2}{2eH}\right)-\frac{1}{2}\ln2\pi\right]\quad .
\end{align*}
This, by the way, is also the result of the calculation of the integral
\begin{equation}
\langle T_\mu^\mu\rangle(H) = -i\frac{eHm^2}{4\pi^2}\int_0^\infty \frac{e^{-2ihz} }{z^2}\left[z\cot z-1+\frac{1}{3}z^2\right]dz,\quad h=\frac{m^2}{2eH}\ .
\end{equation}
Now observe that for $h\ll 1$ we can approximate $\ln\Gamma(h)\approx -\ln h$, such that
\begin{equation}
\lim_{m\rightarrow 0} \langle T_\mu^\mu\rangle(H) = -\frac{1}{12\pi^2}e^2H^2\ ,
\end{equation}
which, when written covariantly, yields
\begin{equation}
m\rightarrow 0\ :\quad \langle T_\mu^\mu\rangle = -\frac{1}{24\pi^2}e^2F_{\mu\nu}F^{\mu\nu} = -\frac{2\alpha}{3\pi}\frac{1}{4}F_{\mu\nu}F^{\mu\nu}\ .
\end{equation}
We can also obtain the next-to-leading term,
\begin{equation}
\langle T_\mu^\mu\rangle = -\beta(\alpha)\frac{1}{4}F_{\mu\nu}F^{\mu\nu},\quad \beta(\alpha)=\frac{2}{3}\left(\frac{\alpha}{\pi}\right)+\frac{1}{2}\left(\frac{\alpha}{\pi}\right)^2\ ,
\end{equation}
by incorporating results for the two-loop calculation $\mathscr{L}^{(2)}$.
For large field strengths, $\frac{eH}{m^2}\gg 1$, the dominant term is
\begin{equation}
\langle T_\mu^\mu\rangle(H) = -\frac{\alpha}{3\pi}H^2\ ,
\end{equation}
while for small field strength, $\frac{eH}{m^2}\ll 1$, we obtain, using Stirling's (Moivre's) formula for the logarithm of the $\Gamma$-function,
\begin{equation}
\langle T_\mu^\mu\rangle(H)=4\left(-\frac{2\alpha^2}{45}\frac{H^4}{m^4}+\frac{64}{315}\pi\alpha^3\frac{H^6}{m^8}+\dots\right)\ .
\end{equation}
Interestingly, the first term in this expansion agrees with Schwinger's\cite{Schwinger1951} from the Heisenberg-Euler Lagrangian:
\begin{equation}
T_{\mu\nu} = T_{\mu\nu}^{Maxwell}\left(1-\frac{16}{45}\alpha^2\frac{1}{m^4}\mathcal{F}\right) - \delta_{\mu\nu}\frac{2}{45}\alpha^2\frac{1}{m^4}\left(4\mathcal{F}^2
+\mathcal{F}\mathcal{G}^2\right)\ .
\end{equation}
In our present case we use $\hbar=c=1$, and for $\vec{E}=\vec{0}$ we have $\mathcal{F}=\tfrac{1}{2}(H^2-E^2)=\tfrac{1}{2}H^2$ and $\mathcal{G}=\vec{E}\cdot \vec{H}=0$.
Since we are interested in the trace of $T_{\mu\nu}$, we obtain (in Schwinger's notation)
\begin{align*}
\langle T_{\mu\mu}\rangle &= -4\left(\frac{8\alpha^2}{45m^4}\mathcal{F}^2+\frac{14\alpha^2}{45m^4}\mathcal{G}^2\right)\ ,
\intertext{which for $\mathcal{G} = 0$ indeed yields}
\langle T_{\mu\mu}\rangle(H) &= -4\left(\frac{2\alpha^2}{45}\frac{H^4}{m^4}\right)\ .
\end{align*}
Let us prove Schwinger's formula. He starts with
\begin{equation}
T_{\mu\nu} = \delta_{\mu\nu}\mathscr{L} - 2\frac{\partial\mathscr{L}}{\partial F_{\mu\lambda}}F_{\nu\lambda}\ .
\end{equation}
(Note that in Schwinger's formula the factor $2$ is missing!)
We need the following derivatives:
\begin{align*}
\frac{\partial\mathscr{L} (\mathcal{F},\mathcal{G}) }{\partial F_{\mu\lambda}} &= \frac{\partial\mathscr{L}}{\partial\mathcal{F}}\frac{\partial\mathcal{F}}{\partial F_{\mu\lambda}}
+\frac{\partial\mathscr{L}}{\partial\mathcal{G}}\frac{\partial\mathcal{G}}{\partial F_{\mu\lambda}},\\
\frac{\partial\mathcal{F}}{\partial F_{\mu\lambda}} &=\frac{\partial}{\partial F_{\mu\lambda}}\left(\frac{1}{4}F^2_{\rho\sigma}\right)=\frac{1}{2} F_{\mu\lambda}\\
\frac{\partial\mathcal{G}}{\partial F_{\mu\lambda}} &= \frac{\partial}{\partial F_{\mu\lambda}}\left(\frac{1}{4}F_{\rho\sigma}F^{*}_{\rho\sigma}\right)
= \frac{1}{4}\frac{\partial}{\partial F_{\mu\lambda}}\left( F_{\rho\sigma}\frac{i}{2}\epsilon_{\rho\sigma\tau\omega}F_{\tau\omega}\right) \\
&= \frac{1}{2} F_{\lambda\mu}^{*}\quad .
\end{align*}
Hence we can write
\begin{align*}
\frac{\partial\mathscr{L} (\mathcal{F},\mathcal{G}) }{\partial F_{\mu\lambda}} &= \frac{\partial\mathscr{L}}{\partial\mathcal{F}}\frac{1}{2}F_{\mu\lambda} + \frac{\partial\mathscr{L}}{\partial\mathcal{G}}\frac{1}{2}F_{\lambda\mu}^{*}\ ,
\intertext{so that}
-2\frac{\partial\mathscr{L}}{\partial F_{\mu\lambda}}F_{\nu\lambda} &= -\frac{\partial\mathscr{L}}{\partial\mathcal{F}}F_{\mu\lambda}F_{\nu\lambda} - \frac{\partial\mathscr{L}}{\partial\mathcal{G}} \underbrace{F_{\lambda\mu}^{*}F_{\nu\lambda}}_{=\mathcal{G}\delta_{\mu\nu}}\\
&= -F_{\mu\lambda}F_{\nu\lambda}\frac{\partial\mathscr{L}}{\partial\mathcal{F}} - \delta_{\mu\nu}\mathcal{G}\frac{\partial\mathscr{L}}{\partial\mathcal{G}}+\left(\delta_{\mu\nu}\mathcal{F}\frac{\partial\mathscr{L}}{\partial\mathcal{F}}
-\delta_{\mu\nu}\mathcal{F}\frac{\partial\mathscr{L}}{\partial\mathcal{F}}\right)\ .
\end{align*}
Putting everything together we obtain
\begin{align*}
T_{\mu\nu} &= -F_{\mu\lambda}F_{\nu\lambda}\frac{\partial\mathscr{L}}{\partial\mathcal{F}}+\delta_{\mu\nu}\mathcal{F}\frac{\partial\mathscr{L}}{\partial\mathcal{F}} + \delta_{\mu\lambda}\mathscr{L}
- \delta_{\mu\nu}\mathcal{F}\frac{\partial\mathscr{L}}{\partial\mathcal{F}} - \delta_{\mu\nu}\mathcal{G}\frac{\partial\mathscr{L}}{\partial\mathcal{G}}\\
&= - \underbrace{\left( F_{\mu\lambda}F_{\nu\lambda}-\delta_{\mu\nu}\frac{1}{4}F_{\lambda\kappa}^2\right)}_{=T_{\mu\nu}^M}\frac{\partial\mathscr{L}}{\partial\mathcal{F}}
+\delta_{\mu\nu}\left(\mathscr{L} - \mathcal{F}\frac{\partial\mathscr{L}}{\partial\mathcal{F}}-\mathcal{G}\frac{\partial\mathscr{L}}{\partial\mathcal{G}}\right)\ ,
\end{align*}
which is a gauge-invariant expression.
Now, from the Heisenberg-Euler effective Lagrangian we are given
\begin{equation}
\mathscr{L} = -\mathcal{F}+C\left[4\mathcal{F}^2+7\mathcal{G}^2\right],\ C=\frac{2\alpha^2}{45m^4},\ \hbar=c=1\ .
\end{equation}
From this expression we obtain the derivatives
\begin{align*}
\frac{\partial\mathscr{L}}{\partial\mathcal{F}} &= -1+8C\mathcal{F}\quad ,& \mathcal{F}\frac{\partial\mathscr{L}}{\partial\mathcal{F}} &= -\mathcal{F}+8C\mathcal{F}^2\\
\frac{\partial\mathscr{L}}{\partial\mathcal{G}} &= 14C\mathcal{G}\quad ,& \mathcal{G}\frac{\partial\mathscr{L}}{\partial\mathcal{G}} &= 14C\mathcal{G}^2\ .
\end{align*}
Finally we end up with
\begin{align*}
T_{\mu\nu} &= T_{\mu\nu}^M \left( 1- \frac{16\alpha^2}{45m^4}\mathcal{F}\right)+\delta_{\mu\nu}\left(\mathscr{L}+\mathcal{F}-8C\mathcal{F}^2-14C\mathcal{G}^2\right),\left( \mathscr{L} = -\mathcal{F} +C(4\mathcal{F}^2+7\mathcal{G}^2)\right)\\
&= T_{\mu\nu}^M\left(1-\frac{16\alpha^2}{45m^4}\mathcal{F}\right) -\delta_{\mu\nu}\frac{2\alpha^2}{45m^4}\left(4\mathcal{F}^2+7\mathcal{G}^2\right)\qquad \square
\end{align*}
Let us put things together. Besides $\langle T_\mu^\mu\rangle(H)$, we can easily produce the corresponding result for a constant electric field by substituting
$H\rightarrow -iE$. The result is
\begin{align*}
\langle T_\mu^\mu\rangle(E) &= \frac{e^2E^2}{12\pi^2}-\frac{m^4}{4\pi^2}\left(i\frac{\pi}{2}+\ln\frac{m^2}{2eE}\right) -i\frac{eEm^2}{4\pi^2}\left(i\frac{\pi}{2}+\ln\frac{m^2}{2eE}\right)\\
&+\frac{m^4}{4\pi^2} -i \frac{eEm^2}{2\pi^2}\left[\ln\Gamma\left(\frac{im^2}{2eE}\right) - \frac{1}{2}\ln2\pi\right]\quad .
\end{align*}
If we split this equation up into its real and imaginary part we obtain
\begin{align*}
\mathfrak{Re}\langle T_\mu^\mu\rangle(E) &= \frac{e^2E^2}{12\pi^2}-\frac{m^4}{4\pi^2}\ln\frac{m^2}{2eE}+\frac{m^2eE}{8\pi}+\frac{m^4}{4\pi^2}
+\frac{eEm^2}{2\pi^2}\mathfrak{Im}\ln\Gamma\left(\frac{im^2}{2eE}\right)\\
\mathfrak{Im}\langle T_\mu^\mu\rangle(E) &= -\frac{m^4}{8\pi}-\frac{m^2}{4\pi^2}eE\ln\frac{m^2}{2eE}-\frac{eEm^2}{2\pi^2}\left[\mathfrak{Re}\ln\Gamma\left(\frac{im^2}{2eE}\right)
-\frac{1}{2}\ln 2\pi\right]\ .
\end{align*}
The last expression can be simplified with the aid of
\begin{equation}
\mathfrak{Re}\ln\Gamma(i\alpha) =\ln\vert\Gamma(i\alpha)\vert = \frac{1}{2}\ln\vert\Gamma(i\alpha)\vert^2 = -\frac{1}{2}\ln\left( \frac{\alpha\sinh(\pi\alpha)}{\pi}\right)\ .
\end{equation}
The result is
\begin{equation}
\mathfrak{Im}\langle T_\mu^\mu\rangle(E) = -\frac{m^4}{8\pi}+\frac{eEm^2}{4\pi^2}\ln\left[2\sinh\frac{\pi m^2}{2eE}\right]\ .
\end{equation}
Let us study these expressions in the limiting case $m\rightarrow 0$. To do this we employ the asymptotic formulae (for $z\ll 1$):
\begin{align*}
\ln\Gamma(z) &\approx \ln z\\
\mathfrak{Im}\ln\Gamma(iz) &\approx Cz,\quad \left( C\approx 0.577216\right)\\
\ln\sinh z &\approx \ln z\quad .
\intertext{We then obtain}
\lim_{m\rightarrow 0} \langle T_\mu^\mu\rangle(H) &= -\frac{1}{12\pi^2}e^2H^2\\
\lim_{m\rightarrow 0} \mathfrak{Re}\langle T_\mu^\mu\rangle(E) &= \frac{1}{12\pi^2}e^2E^2\\
\lim_{m\rightarrow 0} \mathfrak{Im}\langle T_\mu^\mu\rangle(E) &= 0\ .
\end{align*}
These three results are contained in
\begin{equation}
\lim_{m\rightarrow 0} \langle T_\mu^\mu\rangle = -\frac{1}{24\pi^2}e^2F_{\mu\nu}F^{\mu\nu}\ .
\end{equation}
We thus obtain a confirmation of the more general formula (in one-loop approximation)
\begin{equation}
\langle T_\mu^\mu (x) \rangle = -m\langle \bar{\psi}(x)\psi(x)\rangle -\frac{1}{24\pi^2}e^2F_{\mu\nu}(x)F^{\mu\nu}(x)\ ,
\end{equation}
where $\lim_{m\rightarrow 0} \left(m\langle \bar{\psi}(x)\psi(x)\rangle\right) = 0$.
Let us have a final look at $\mathfrak{Im} \langle T_\mu^\mu\rangle(E)$, and write it in units of $E^2_{cr} :=\tfrac{m^4}{4\pi\alpha}$, and the electric field
in units of $E_{cr} := \frac{m^2}{e}$. Thus we obtain
\begin{align*}
\mathfrak{Im} \langle T_\mu^\mu\rangle(E) &= -\frac{\alpha}{2}+\frac{\alpha}{\pi}E\ln\left[2\sinh\frac{\pi}{2E}\right]\\
&= -\frac{\alpha}{2}+\frac{\alpha}{\pi}E\ln\left[e^{\frac{\pi}{2E}}-e^{-\frac{\pi}{2E}}\right]\\
&= -\frac{\alpha}{2}+\frac{\alpha}{\pi}E\left[\ln e^{\frac{\pi}{2E}}+\ln\left(1-e^{-\frac{\pi}{E}}\right) \right]\ .
\end{align*}
Here we use $\ln(1-x)=-\sum_{n=1}^\infty \frac{x^n}{n},\ -1\leq x < 1$, which yields
\begin{equation}
\mathfrak{Im} \langle T_\mu^\mu\rangle(E) = -\frac{\alpha}{2}+\frac{\alpha}{2}-\frac{\alpha}{\pi}E\sum_{n=1}^\infty\frac{1}{n}e^{-\frac{\pi}{E}n} =
-\frac{\alpha}{\pi}E\sum_{n=1}^\infty\frac{1}{n}e^{-\frac{\pi}{E}n}\ .
\end{equation}
For small values we find approximately ($E \ll 1$):
\begin{equation}
\mathfrak{Im}\ \langle T_\mu^\mu\rangle(E)\approx -\frac{\alpha}{\pi}Ee^{-\frac{\pi}{E}}\ ,
\end{equation}
which goes to zero for $E\rightarrow 0$.
Our result can also be obtained by using the well-known formula (c.f. $e^+-e^-$ pair production)
\begin{equation}
\mathfrak{Im}\ \mathscr{L}^{(1)}(E) = \frac{\alpha}{2\pi^2}E^2\sum_{n=1}^\infty \frac{1}{n^2}e^{-\frac{\pi m^2}{eE}n}\quad .
\end{equation}
We only need to write
\begin{align*}
\mathfrak{Im}\ \langle T_\mu^\mu\rangle(E) &= m\frac{\partial}{\partial m}\mathfrak{Im}\ \mathscr{L}^{(1)}(E)
= m\frac{\alpha}{2\pi^2}E^2\sum_{n=1}^\infty \frac{1}{n^2}\left(\frac{-2m\pi n}{eE}\right)e^{-\frac{\pi m^2}{eE}n}\\
&= -\frac{m^2 eE}{4\pi^2}\sum_{n=1}^\infty \frac{1}{n}e^{-\frac{\pi m^2}{eE}n}\quad ,
\intertext{or, in our units,}
\mathfrak{Im} \langle T_\mu^\mu\rangle(E) &= -\frac{\alpha}{\pi}E\sum_{n=1}^\infty \frac{1}{n}e^{-\frac{\pi}{E}n}\qquad \square
\end{align*}
Up to now, we have always restricted our calculations to the case of constant electric or magnetic fields. It can be shown, however, that the leading terms for strong fields,
i.e., those of order $H^2\ln H$ or $E^2\ln E$, are the same if the fields are not constant. This will be demonstrated in the next chapter, where we extend our discussion to the effective
Yang-Mills field theory.
\section{The Effective Action for Yang-Mills Theory}\label{sec:EffActYangMills}
The effective action of quantum chromodynamics (QCD) for covariant constant color fields has been extensively treated in the literature by many researchers.
But they rarely pose the question in how far their results are physically reasonable and applicable. If one assumes that the confinement hypothesis is correct,
then no constant color fields can exist. Thus it would be physically senseless to study the effective Lagrangian (or the effective potential) in an
exact covariant constant color field. If, however, we regard a color field that in an expanded, but limited space, can be considered
to be approximately covariant constant, then one could suppose that in the space in question the effective Lagrangian can be approximated by the effective
Lagrangian of a covariant constant field. Thus we try to extrapolate from the case of an unlimited, expanded covariant color field to the case of a color field
that is in an expanded but limited space, approximately covariant constant. Upon looking more closely, it turns out that this procedure is physically unsatisfactory,
because one first calculates the effective Lagrangian for the covariant constant field configuration, which is not even theoretically feasible - this is forbidden by the
confinement hypothesis - and then tries to extrapolate to a physical situation. For such an extrapolation from a nonphysical to a physical situation, one cannot expect that the
result is in any way physically acceptable. Thus the results obtained for the effective potentials with covariant constant color fields should not be used to describe
the nature of the QCD vacuum, but rather a transitional phase in the search for the true QCD vacuum.
After this prelude we will return to the role of the energy-momentum tensor in Yang-Mills theory. Rather than attempt the difficult task of computing $\mathscr{L}^{eff}$ as
done in \cite{Dittrich1983}, we will instead make an ansatz. Our ansatz will be motivated by the requirements that $\mathscr{L}^{eff}(x)$ gives the correct trace anomaly for the
energy-momentum tensor, and depends only on the algebraic invariant $F^2 := F^a_{\mu\nu}F^{a\mu\nu}$\cite{Pagels1978}.
So we require
\begin{align*}
\Theta^{\mu\nu} &= 2\frac{\mathscr{L} (F^2)}{\partial\eta_{\mu\nu}} - \eta^{\mu\nu}\mathscr{L}(F^2)\\
\Theta_\mu^\mu &= \frac{\beta(\bar{g}(t) )}{2\bar{g}^3(t)}F^a_{\mu\nu}F^{a\mu\nu}\ ,
\end{align*}
so that $\Theta_\mu^\mu$ has the usual form of the trace anomaly. Now we want to prove that these requirements are satisfied by the ansatz
\begin{equation}
\mathscr{L}^{eff} := -\frac{1}{4}\frac{1}{\bar{g}^2(t)}F^2\ ,\quad t:=\frac{1}{4}\ln\frac{F^2}{\mu^2}\quad .
\end{equation}
So we have to calculate
\begin{align*}
\Theta_{\mu\nu} &= 2\frac{\partial\mathscr{L}^{eff}}{\partial \eta^{\mu\nu}} - \eta_{\mu\nu}\mathscr{L}^{eff}\\
&= \left( 2\frac{1}{\bar{g}^2(t)}\frac{\partial}{\partial \eta^{\mu\nu}}\left(-\frac{1}{4}F^2\right) - \eta_{\mu\nu}\mathscr{L}\right)
- \frac{2}{4}\left(\frac{\partial}{\partial\eta^{\mu\nu}}\frac{1}{\bar{g}^2(t)}\right)F^2\\
&= \frac{1}{\bar{g}^2(t)}\left[ \frac{1}{4}\eta_{\mu\nu}F^2-F_\mu^\alpha F_{\nu\alpha}\right] - \frac{1}{2}\left(\frac{\partial t}{\partial\eta^{\mu\nu}}\frac{d}{dt}\frac{1}{\bar{g}^2(t)}\right)F^2\\
&= \frac{1}{\bar{g}^2(t)}\left[ F_\mu^\alpha F_{\alpha\nu} + \frac{1}{4}\eta_{\mu\nu}F^2\right] + \frac{1}{4}\left(\frac{d}{dt}\frac{1}{\bar{g}^2(t)}\right)F^a_{\mu\beta}F^{a\beta}_\nu\ .
\end{align*}
The last term on the right-hand side uses the result
\begin{equation}
\frac{d}{dt}\frac{1}{\bar{g}^2(t)}= - \frac{2}{\bar{g}^3(t)}\frac{d\bar{g}}{dt}(t) = -2\frac{\beta(\bar{g}(t))}{\bar{g}^3(t)}\quad ,
\end{equation}
where we employed
\begin{equation}
t = \int_{g}^{\bar{g}(t)} \frac{dg'}{\beta(g')}\quad ,
\end{equation}
which, when taken the derivative $\tfrac{d}{dt}$ of, gives
\begin{equation}
1 = \frac{1}{\beta(\bar{g}(t))}\frac{d}{dt}\bar{g}(t),\text{ i.e., } \frac{d}{dt}\bar{g}(t) = \beta(\bar{g}(t))\quad .
\end{equation}
So we proved the relation
\begin{align*}
\Theta_{\mu\nu} &= \frac{1}{\bar{g}^2(t)}\left[F_\mu^\alpha F_{\alpha\nu} + \frac{1}{4}\eta_{\mu\nu}F^2\right] - \frac{\beta(\bar{g}(t))}{2\bar{g}^3(t)}F^a_{\mu\beta}F^{a\beta}_\nu\ ,
\intertext{and from here the trace}
\Theta_\mu^\mu &= \frac{\beta(\bar{g}(t))}{2\bar{g}^3(t)}F^a_{\mu\beta}F^{a\mu\beta}\ ,\quad t=\frac{1}{4}\ln\frac{F^2}{\mu^4}\quad .\quad \square
\end{align*}
We also can verify that our ansatz $\mathscr{L}^{eff} = -\frac{1}{4}\frac{F^2}{\bar{g}^2(t)},\ t=\frac{1}{4}\ln\frac{F^2}{\mu^4}$ satisfies the
renormalization group equation
\begin{equation}
\left[\mu\frac{\partial}{\partial\mu}+\beta(g)\frac{\partial}{\partial g}\right]\left(-\frac{1}{4}F^2\right)\frac{1}{\bar{g}^2(t)} = 0\quad .
\end{equation}
Proof:
\begin{align*}
\mu\frac{\partial}{\partial\mu}\frac{1}{\bar{g}^2(t)} &= \mu\frac{\partial t}{\partial\mu}\frac{d}{dt}\frac{1}{\bar{g}^2(t)}
= \mu\frac{1}{4}\frac{\partial}{\partial\mu}\left( \ln F^2 -4\ln\mu\right)(-2)\frac{\beta(\bar{g}(t))}{\bar{g}^3(t)}\\
&= 2 \frac{\beta(\bar{g}(t))}{\bar{g}^3(t)}\quad .
\end{align*}
Taking the derivative $\tfrac{\partial}{\partial g}$ of $t=\int_{g}^{\bar{g}(t)} \tfrac{dg'}{\beta(g')}$ we obtain
\begin{equation}
\frac{\partial}{\partial g}t = 0 = \frac{\partial\bar{g}(t)}{\partial g}\frac{1}{\beta(\bar{g}(t))}-\frac{1}{\beta(g)}\ \text{ or }\ \frac{\partial\bar{g}(t)}{\partial g} =
\frac{\beta(\bar{g}(t))}{\beta(g)}
\end{equation}
and from here
\begin{equation}
\beta(g)\frac{\partial}{\partial g }\frac{1}{\bar{g}^2(t)} = \beta(g)(-2)\bar{g}^{-3}(t)\frac{\partial\bar{g}(t)}{\partial g}
= -2\frac{1}{\bar{g}^3(t)}\beta(g)\frac{\beta(\bar{g}(t))}{\beta(g)} = -2\frac{\beta(\bar{g}(t))}{\bar{g}^3(t)}\ .
\end{equation}
So we proved the equation
\begin{equation}
\left[\mu\frac{\partial}{\partial\mu}+\beta(g)\frac{\partial}{\partial g}\right]\frac{1}{\bar{g}^2(t)}=0\ .
\end{equation}
The effective Lagrangian is specified once we know $\beta(g)$. For weak coupling we have
\begin{align*}
g\beta(g) &= -\frac{1}{2}b_0g^4+b_1g^6+\dots\\
\beta(g) &= -\frac{1}{2}b_0g^3+b_1 g^5+\dots = -\frac{1}{2}b_0 g^3\left( 1- 2\frac{b_1}{b_0}g^2\right)+\dots \quad .
\end{align*}
This we substitute into
\begin{align*}
t &= \int_{g}^{\bar{g}(t)} \frac{dg'}{\beta(g')} = -\frac{2}{b_0}\int_g^{\bar{g}(t)} \frac{1}{g'^3(1-2\frac{b_1}{b_0}g'^2+\dots )}dg'\\
&= -\frac{2}{b_0}\int_{g}^{\bar{g}(t)}\left(\frac{1}{g'^3}+2\frac{b_1}{b_0}\frac{1}{g'}+\dots\right) dg'\\
&= -\frac{2}{b_0}\left[ -\frac{1}{2g'^2}+2\frac{b_1}{b_0}\ln g'+\dots\right]^{\bar{g}(t)}_{g}
\intertext{ to obtain}
t &= \frac{1}{b_0}\frac{1}{\bar{g}^2(t)}-\frac{2}{b_0}2\frac{b_1}{b_0}\ln\bar{g}(t)+\dots \underbrace{-\frac{1}{b_0}\frac{1}{g^2}+\frac{2}{b_0}2\frac{b_1}{b_0}\ln g+\dots}_{=\text{const.}\cdot t}\quad .
\end{align*}
Hence for $g(t)\ll 1$ we find the approximations
\begin{align*}
t &= \frac{1}{b_0}\frac{1}{\bar{g}^2_{(1)}(t)} &: \bar{g}_{(1)}(t) &= \sqrt{\frac{1}{b_0t}}\\
t &= \frac{1}{b_0}\frac{1}{\bar{g}^2_{(2)}(t)}\underbrace{-\frac{2}{b_0}2\frac{b_1}{b_0}\ln\sqrt{\frac{1}{b_0t}}}_{=\frac{1}{b_0}2\frac{b_1}{b_0}\ln t+\text{const.}}
&: \frac{1}{\bar{g}_{(2)}^2(t)} &=b_0t-2\frac{b_1}{b_0}\ln t\ .
\end{align*}
Consequently for large fields $F^2$, the effective Lagrangian is controlled by perturbation theory (asymptotic freedom) and is given by
\begin{equation}
\mathscr{L}^{eff}_{\text{leading log}} = -\frac{1}{16}b_0F^2\ln\frac{F^2}{\mu^4} + \mathcal{O}\left(F^2\ln\ln F^2\right)\quad .
\end{equation}
By the way, in QCD we have
\begin{align*}
\beta(g) &= \mu\frac{\partial g}{\partial\mu}\quad ,
\intertext{where}
\beta(g) &= -\frac{g^3}{16\pi^2}\left(\frac{11}{3}N-\frac{2}{3}N_f\right)+\dots\\
g\beta(g) &= -\frac{1}{2}g^4\frac{1}{8\pi^2}\left(\frac{11}{3}N-\frac{2}{3}N_f\right)+\mathcal{O}(g^6)\\
&= -\frac{1}{2}b_0g^4 + \mathcal{O}(g^6)\quad ,
\intertext{so that}
b_0 &= \frac{1}{8\pi^2}\left(\frac{11}{3}N-\frac{2}{3}N_f\right)\quad .
\end{align*}
Finally, let us write the leading-log effective Lagrangian in a form that will also be useful in QED:
\begin{align*}
\mathscr{L}^{eff} &= -\frac{1}{4}\frac{1}{\bar{g}^2(t)}F^2\ ,\\
t &:= \frac{1}{4}\ln\frac{F^2}{\mu^4}=\frac{1}{4}\ln\frac{2(B^2-E^2)}{\mu^4}\\
F^2 &:= F^a_{\mu\nu}F^{a\mu\nu} = -2\left(\vec{E}^a\cdot \vec{E}^a - \vec{B}^a\cdot\vec{B}^a\right) \equiv -2\left( E^2-B^2\right)\ .
\end{align*}
Then $\mathscr{L}^{eff}$ is given by
\begin{align*}
\mathscr{L}^{eff} &= -\frac{1}{4}b_0tF^2+\dots = -\frac{1}{4}b_0\frac{1}{4}\left[\ln\frac{F^2}{\mu^4}\right]F^2+\dots \\
&= -\frac{1}{16}b_0F^2\ln\frac{F^2}{\mu^4}+\dots \\
\mathscr{L}_{eff}^{(1)} &= \frac{1}{8}b_0(E^2-B^2)\ln\frac{2(B^2-E^2)}{\mu^4}+\dots\quad .
\end{align*}
When we rescale $g^2$ into the fields we obtain
\begin{equation}
\mathscr{L}^{eff} = \frac{1}{2}\left( \vec{E}^2-\vec{B}^2\right) + \frac{1}{2}\left( \vec{E}^2-\vec{B}^2\right)\frac{g^2}{48\pi^2}N\frac{11}{2}\ln\frac{g^2(\vec{B}^2-\vec{E}^2)}{\mu^4}
+\mathcal{O}(F^2)\ .
\end{equation}
This, by the way, is the same result that one obtains for covariant constant color fields.
Special cases are given by
\begin{align*}
\vec{E}=\vec{0}\quad :\quad \mathscr{L}^{eff}(B) &= -\frac{1}{2}B^2-\frac{1}{2}B^2\frac{g^2}{48\pi^2}N\frac{11}{2}\ln\frac{g^2B^2}{\mu^4}\\
&= -\frac{1}{2}B^2\left(1+\frac{g^2}{48\pi^2}11N\ln\frac{gB}{\mu^2}\right)\\
B\rightarrow\frac{1}{i}E\quad : \quad \mathscr{L}^{eff}(E) &= \frac{1}{2}E^2\left(1+\frac{g^2}{48\pi^2}11N\ln\frac{g(-iE)}{\mu^2}\right)\\
&= \frac{1}{2}E^2\left( 1+\frac{g^2}{48\pi^2}11N\left(\ln\frac{gE}{\mu^2} - \frac{i\pi}{2}\right)\right)\ .
\end{align*}
Since $V^{eff}(B) = -\mathscr{L}^{eff}(B)$ we obtain for the effective potential in QCD for $SU(N=3)$:
\begin{equation}
V^{eff}(B) = \frac{1}{2}B^2\left[ 1+\frac{g^2}{4\pi^2}\left(11\cdot 3 - 2N_f\right)\ln\frac{gB}{\mu^2}\right]\quad .
\end{equation}
Since $-\tfrac{1}{4}F^2=\tfrac{1}{2}(E^2-B^2)$, we have for the color magnetic fields only $B^2 = \tfrac{1}{2}F^2$:
\begin{align*}
V^{eff}(B) &= \frac{1}{4}F^2\left[ 1+\frac{g^2}{4}b_0\ln\frac{(gF)^2}{2\mu^4}\right]\\
b_0 &= \frac{1}{8\pi^2}\left(\frac{11}{3}\cdot 3 -\frac{2}{3}N_f\right)\ , \quad N=3\quad .
\end{align*}
To find the minimum we take the derivative $\tfrac{\partial V^{eff}(F)}{\partial F^2}$ and set it equal to zero. The result is
\begin{align*}
\ln\frac{e(gF)^2}{2\mu^4} &= -\frac{4}{b_0g^2}
\intertext{or}
\langle 0\vert (gF)^2\vert 0 \rangle &= \frac{2\mu^4}{e}e^{-\frac{4}{b_0g^2}}\quad ,
\intertext{which gives the dimensionless number}
\frac{\langle 0\vert (gF)^2\vert 0 \rangle}{\mu^4} &= \frac{2}{e}\underbrace{e^{-\frac{4}{b_0g^2}} }_{< 1}=0.7357e^{-\frac{4}{b_0g^2}}\quad .
\end{align*}
From here we find the expression for $V^{eff}_{min}(F)$ to be:
\begin{align*}
V^{eff}_{min}(F) &= \frac{1}{4}\langle F^2\rangle\left[1+\frac{g^2}{4}b_0\ln\frac{\langle (gF)^2\rangle}{2\mu^4}\right]\\
&= \frac{1}{4}\langle F^2\rangle \left[1+\frac{g^2}{4}b_0\left(\ln\frac{1}{e} + \ln e^{-\frac{4}{b_0g^2}}\right) \right]\\
&= -\frac{b_0}{16}\langle 0\vert (gF)^2\vert 0 \rangle \quad \left(=0.7124\cdot 10^{-2}\langle 0\vert (gF)^2\vert 0 \rangle \text{ for 3 massless flavours}\right)\\
&= -\frac{1}{128\pi^2}\left(11-\frac{2}{3}N_f\right)\langle 0\vert (gF)^2\vert 0 \rangle\\
&\underset{g^2=4\pi\alpha}{=} -\frac{1}{128\pi^2}\left(\frac{11}{3}N-\frac{2}{3}N_f\right)\langle 0\vert 4\pi\alpha F^2\vert 0 \rangle\\
&= -\frac{1}{32}\left(\frac{11}{3}N-\frac{2}{3}N_f\right)\langle 0 \vert \frac{\alpha}{\pi}F^a_{\mu\nu}F^{a\mu\nu}\vert 0\rangle + \mathcal{O}(\alpha^2)\quad .
\end{align*}
This result is consistent with the trace
\begin{equation}
\langle 0\vert T_\mu^\mu\vert 0 \rangle = \frac{1}{8}\left(\frac{11}{3}N-\frac{2}{3}N_f\right)\langle 0 \vert \frac{\alpha}{\pi}F^a_{\mu\nu}F^{a\mu\nu}\vert 0\rangle\ .
\end{equation}
From Lorentz invariance $T_{\mu\nu}(x)=\text{const.}\cdot g_{\mu\nu}$ and $T_{00} = \epsilon$; therefore $T_{\mu\nu}=\epsilon g_{\mu\nu}$, and we obtain
$T_\mu^\mu = 4\epsilon$ so that the lowering of the energy caused by non-perturbative fluctuations of the color field yields in the vacuum state
\begin{equation}
\epsilon = -\frac{1}{32}\left(\frac{11}{3}N-\frac{2}{3}N_f\right)\langle 0 \vert \frac{\alpha}{\pi}F^a_{\mu\nu}F^{a\mu\nu}\vert 0\rangle+\mathcal{O}(\alpha^2)\ .
\end{equation}
For $N=3,N_f=3$ we have $(\tfrac{11}{3}N-\tfrac{2}{3}N_f) = 9$.
\section{The Effective Lagrangian in QED}
We want to investigate the modification of Coulomb's law for long and short distances. First, we will ask for the effective Lagrangian
for weak, but otherwise arbitrary, fields.
In the weak-field limit, $e^2F_{\mu\nu}F^{\mu\nu}/m^4$ becomes small due to the smallness of $\alpha$. This leads to the expression
(wf = weak-field):
\begin{equation}
W_{wf}^{(1)}[A] = \int\mathscr{L}^{(1)}_{wf}d^4x = \frac{1}{2}\int A^\mu(x)\Pi_{\mu\nu}(x,y)A^\nu(y)d^4xd^4y
\end{equation}
for the weak-field limit of the one-loop effective action, where $\Pi_{\mu\nu}$ is nothing but the well-known order-$e^2$ polarization tensor of QED.
In momentum space it is given by
\begin{align*}
\Pi_{\mu\nu}(k) &= \left( g_{\mu\nu} k^2-k_\mu k_\nu\right)\Pi(k^2)\ ,\\
\Pi(k^2) &= -\frac{\alpha}{3\pi}k^2\int_{4m^2}^\infty \frac{1}{t}\rho(t)\frac{1}{k^2+t-i\epsilon}dt\ ,\\
\rho(t) &= \left(1+\frac{2m^2}{t}\right)\left(1-\frac{4m^2}{t}\right)^{\frac{1}{2}}\quad .
\end{align*}
As a consequence of the particular tensor structure of $\Pi_{\mu\nu}(k)$, $W_{wf}^{(1)}$ can be written in terms of $F_{\mu\nu}$ only and therefore is
gauge invariant.
Adding the classical Maxwellian term, we obtain for the real field effective Lagrangian
\begin{equation}
\mathscr{L}^{eff}_{wf} = -\frac{1}{4}F_{\mu\nu}(x)\left[1+\frac{\alpha}{3\pi}\square\int_{4m^2}^\infty \frac{1}{t}\frac{\rho(t)}{t-\square}dt\right]F^{\mu\nu}(x)
\end{equation}
where, as usual, $\square = -\partial_t^2+\vec{\nabla}^2$.
The equations of motion resulting from this effective Lagrangian for the weak-field case are linear, because $\mathscr{L}^{eff}_{wf}$ is quadratic in the fields.
Next, let us apply $\mathscr{L}^{eff}_{wf}$ to the Coulomb problem.
Specializing to the static case begins with the variation
\begin{equation}
\frac{\delta}{\delta A^0(\vec{x})}\int d^3x'\left[ \frac{1}{2} \vec{\nabla} A^0\cdot \left( 1+\frac{\alpha}{3\pi}\vec{\nabla}^2\int_{4m^2}^\infty \frac{1}{t}\frac{\rho(t)}{t-\vec{\nabla}^2}dt\right)
\vec{\nabla} A^0-A^0J_0\right] = 0\ .
\end{equation}
For $J_0(\vec{x})$ we assume two point charges with the separation $r$:
\begin{equation}
J_0(\vec{x})=Q\left[\delta\left(\vec{x}-\vec{x}_1\right) - \delta\left(\vec{x}-\vec{x}_2\right)\right]\ ,\quad |\vec{x}_1-\vec{x}_2| = r\ .
\end{equation}
The variation then gives the equation of motion
\begin{align*}
D\vec{\nabla}^2 A^0(\vec{x}) &= -J_0(\vec{x})\quad ,
\intertext{where}
D &:= 1+\frac{\alpha}{3\pi}\vec{\nabla}^2\int_{4m^2}^\infty \frac{1}{t}\frac{\rho(t)}{t-\vec{\nabla}^2}dt\quad .
\end{align*}
Making use of the position space representation of the resolvent $(t-\vec{\nabla}^2)^{-1}$, one can easily calculate the potential energy
$V=-\int\mathscr{L}_{wf}^{eff}(A^0)d^3\vec{x}$ associated with the interaction of two point charges. One finds
\begin{equation}
V(r) = -\frac{Q^2}{4\pi}\left[\frac{1}{r}+\frac{\alpha}{3\pi}\int_{4m^2}^\infty \frac{\rho(t)}{t}\frac{e^{-\sqrt{t}r}}{r}dt\right]+\mathcal{O}(\alpha^2)\ .
\end{equation}
The second term in the brackets is the well-known Uehling correction to the classical Coulomb potential. $V(r)$ was derived in the weak-field limit and thus should be
valid at large distances. Because the equation of motion is linear, $V(r)$ takes the form of a superposition of Yukawa potentials.
The quantum mechanical correction to a many-particle static potential,
\begin{align*}
A^0(\vec{x}) &= \sum_i \frac{Q_i}{4\pi|\vec{x}-\vec{x}_i|}\ ,\quad J_0(\vec{x}) = \sum_iQ_i\delta(\vec{x}-\vec{x}_i)\\
V_{static} &= \frac{1}{2}\sum_{i\neq j}\frac{Q_iQ_j}{4\pi|\vec{x}_i-\vec{x}_j|}\ , \quad r_{ij} = |\vec{x}_i-\vec{x}_j|\ ,\\
\intertext{becomes}
V_{static} &= \frac{1}{2}\sum_{i\neq j}\frac{Q_iQ_j}{4\pi}\left[\frac{1}{r_{ij}}+\frac{\alpha}{3\pi}\int_{4m^2}^\infty \frac{\rho(t)}{t}\frac{e^{-\sqrt{t}r_{ij}}}{r_{ij}}dt
\right]+\mathcal{O}(\alpha^2)\ .
\end{align*}
Now that we have established the Lagrangian $\mathscr{L}^{eff}_{wf}$ and the (one-loop) correction to the Coulomb potential for weak fields, we want to set up the
generalized Maxwell equations for strong fields. The following calculations are justified by noting that the ansatz
\begin{equation}
\mathscr{L}_{eff} = -\frac{1}{4e^2(F^2)}F_{\mu\nu}F^{\mu\nu}
\end{equation}
leads to the correct trace anomaly of the energy-momentum tensor (c.f. \ref{sec:EffActYangMills})
One starts from the free Maxwell Lagrangian
\begin{equation}
\mathscr{L} = -\frac{1}{4}F_{\mu\nu}(x)F^{\mu\nu}(x)
\end{equation}
and scales the electromagnetic coupling $e$ out of the fields
\begin{align*}
A_\mu &\rightarrow \frac{1}{e}A_\mu\quad ,
\intertext{giving}
\mathscr{L} &= -\frac{1}{4e^2}F_{\mu\nu}F^{\mu\nu}\ .
\end{align*}
Note that in the complete interacting QED Lagrangian this is the only term containing $e$, because the vertex now simply
reads $\bar{\psi}\gamma\cdot A\psi$ instead of $e\bar{\psi}\gamma\cdot A\psi$.
The next step is to renormalization-group improve $\mathscr{L}$ by replacing $e$ with the running coupling constant $e(\mu)$ to first
order in $\alpha$.
To achieve this we make use of the scale variation of the gauge coupling constant\cite{Ramond1997}
\begin{equation}
\mu\frac{\partial e}{\partial\mu} = \beta(e) = \frac{1}{12\pi^2}e^3\ .
\end{equation}
This equation is solved by
\begin{equation}
\frac{1}{e^2(\mu)}-\frac{1}{e^2(\mu_0)} = -\frac{1}{6\pi^2}\ln\frac{\mu}{\mu_0}\ ,
\end{equation}
where $\mu_0$ denotes an arbitrary scale.
To prove this let us rewrite the last equation in the form
\begin{align*}
\frac{1}{e^2(\mu)} &= \frac{1}{e^2(\mu_0)} -\frac{1}{6\pi^2}\ln\frac{\mu}{\mu_0}\\
\text{or}\qquad e(\mu) &= \left(\frac{1}{e^2(\mu_0)} -\frac{1}{6\pi^2}\ln\frac{\mu}{\mu_0}\right)^{-\frac{1}{2}}\ .
\intertext{Hence}
\frac{\partial e(\mu)}{\partial \mu} &= -\frac{1}{2} \left(\frac{1}{e^2(\mu_0)} -\frac{1}{6\pi^2}\ln\frac{\mu}{\mu_0}\right)^{-\frac{3}{2}}
\left(-\frac{1}{6\pi^2}\right)\frac{1}{\mu}\ ,
\intertext{which yields}
\mu\frac{\partial e(\mu)}{\partial\mu} &= \frac{1}{12}e^3\qquad \square
\end{align*}
Let us rewrite our solution slightly:
\begin{align*}
\frac{1}{e^2(\mu_0)}\left(\frac{e^2(\mu_0)}{e^2(\mu)}-1\right) &= -\frac{1}{6\pi^2}\ln\frac{\mu}{\mu_0}\\
\text{or}\qquad\frac{e^2(\mu_0)}{e^2(\mu)} &= 1-\frac{e^2(\mu_0)}{6\pi^2}\ln\frac{\mu}{\mu_0}\ .
\end{align*}
So the scaling equation for $e^2(\mu) := 4\pi\alpha(\mu)$ is given by
\begin{equation}
e^2(\mu) = \frac{e^2(\mu_0)}{1-\frac{e^2(\mu_0)}{6\pi^2}\ln\frac{\mu}{\mu_0}}\ .
\end{equation}
This equation has a singularity which follows from
\begin{align*}
1-\frac{e^2(\mu_0)}{6\pi^2}\ln\frac{\mu}{\mu_0} &= 0\\
\text{or}\qquad \frac{e^2(\mu_0)}{6\pi^2}\ln\frac{\mu}{\mu_0} &= 1\ ,
\end{align*}
which is solved by $\mu = \mu_0 e^{\frac{6\pi^2}{e^2(\mu_0)} }$ and is known as Landau singularity.
Our scaling equation underlines the fact that the electric charge grows weaker and weaker at large distances (i.e., small scales),
which means that the identification of the free Lagrangian ($e=0$) in terms of physical photons is perfectly justified.
In our application, where the fields are sufficiently strong so that fermionic masses are negligible, the length or mass scale is set by the
magnitude $F^2 = F_{\mu\nu}F^{\mu\nu}$. Therefore we replace in the scaling equation $\mu^4$ by $F^2$ to obtain
\begin{equation}
e^2(F^2) = \frac{e^2(\mu_0)}{1-\frac{e^2(\mu_0)}{24\pi^2}\ln\frac{F^2}{\mu_0^4}}\ ,
\end{equation}
with an arbitrary integration constant and arbitrary reference mass $\mu_0$. After replacing $e^2$ by the field-dependent running coupling
constant $e^2(F^2)$ we obtain
\begin{align*}
\mathscr{L}_{eff} &= -\frac{1}{4e^2(F^2)}F_{\mu\nu}F^{\mu\nu}\\
&= -\frac{1}{4e^2(\mu_0)}F_{\mu\nu}F^{\mu\nu}\left[ 1- \frac{e^2(\mu_0)}{24\pi^2}\ln\frac{F^2}{\mu_0^4}\right]\ .
\end{align*}
The one-loop part is (we scale back $e^2(\mu_0) := e^2$ into the fields)
\begin{align*}
\mathscr{L}^{(1)} &= \frac{1}{4}F_{\mu\nu}F^{\mu\nu}\frac{e^2}{24\pi^2}\ln\frac{e^2F^2}{\mu_0^4}\\
&\overset{F^2=-2(\vec{E}^2-\vec{B}^2)}{=} \frac{1}{2}\left(\vec{B}^2-\vec{E}^2\right)\frac{e^2}{24\pi^2}\ln\frac{e^2\left(\vec{B}^2-\vec{E}^2\right)}{\mu_0^4}
+\mathcal{O}(F^2)\\
\vec{E}=\vec{0}:\qquad \mathscr{L}^{(1)}(B) &= \frac{1}{2}B^2\frac{e^2}{24\pi^2}\ln\frac{(eB)^2}{\mu_0^4} = \frac{\alpha B^2}{6\pi}\ln\frac{eB}{\mu_0^2}\\
B\rightarrow\frac{1}{i}E:\qquad \mathscr{L}^{(1)}(E) &= -\frac{\alpha E^2}{6\pi}\ln\frac{e(-iE)}{\mu_0^2}\\
&= -\frac{\alpha E^2}{6\pi}\ln\frac{eE\cdot e^{-i\frac{\pi}{2}} }{\mu_0^2} = -\frac{\alpha E^2}{6\pi}\left[\ln\frac{eE}{\mu_0^2}-\frac{\pi}{2}i\right]\\
\mu_0 \equiv m:\qquad\mathscr{L}^{(1)}(E) &= -\frac{\alpha E^2}{6\pi} \left(\ln\frac{eE}{m^2}-\frac{\pi}{2}i\right)\quad ,
\end{align*}
and this is our old formula for $\mathscr{L}^{(1)}\left(\frac{eE}{m^2}\rightarrow\infty\right)$ which was formally derived for constant fields only.
However, at no point in the above \textquotedblleft derivation\textquotedblright\ of $\mathscr{L}^{(1)}$ did we have to demand the fields to be
constant; thus we may assume that $\mathscr{L}^{(1)}$ is correct to order $F^2\ln F$ for arbitrary varying fields.
Now that we have established the Lagrangian for strong but otherwise arbitrary fields, we can set up the generalized Maxwell equations and try to solve them
for a given source distribution. In general, they are of the form
\begin{equation}
\frac{\delta}{\delta A_\mu(x)}\int\left[\mathscr{L}^{(0)}+\mathscr{L}^{(1)}-J_\mu A^\mu \right]d^4x' = 0\ ,
\end{equation}
with $\mathscr{L}^{(1)}$ given by its real part. Furthermore $J_\mu (x) = J_0(\vec{x})\delta_\mu$ and $\vec{E}(\vec{x}) = -\vec{\nabla} A^0(\vec{x})\ (A^0=\phi)$.
This leads us to evaluating
\begin{equation}
\frac{\delta}{\delta A^0(\vec{x})}\int \left(\frac{1}{2}|\vec{\nabla} A^0|^2\left[1-\frac{e^2}{12\pi^2}\ln\frac{e|\vec{\nabla} A^0|}{m^2}\right]-J^0A^0\right) d^3\vec{x}' = 0\ ,
\end{equation}
which is equivalent to
\begin{equation}
V_{static} = -\text{ext.}_{\phi} \int\left( \frac{1}{2}(\vec{\nabla} \phi)^2\left[ 1 -\frac{\alpha}{3\pi}\ln\frac{e|\vec{\nabla}\phi|}{m^2}\right] -\phi J_0\right)d^3x\ .
\end{equation}
The variation of our $\phi$ then gives the local nonlinear differential equation for $\phi \equiv A^0$:
\begin{align*}
\partial_k \frac{\partial \mathscr{L}}{\partial(\partial_k\phi)} &= \frac{\partial \mathscr{L}}{\partial \phi}\quad ,
\intertext{where}
\mathscr{L}(\phi,\partial_i\phi) &= \frac{1}{2}\partial_k\phi\partial_k\phi\left[1-\frac{\alpha}{6\pi}\ln\frac{e^2(\partial_k\phi)(\partial_k\phi)}{m^4}\right]-\phi J_0\\
\frac{\partial\mathscr{L}}{\partial\phi} &= -J_0\\
\frac{\partial\mathscr{L}}{\partial(\partial_k\phi)} &= \underbrace{\partial_k\phi}_{=-E_k}\left[1-\frac{\alpha}{6\pi}\ln\frac{e^2(\partial_k\phi)(\partial_k\phi)}{m^4}\right]
+\left( -\frac{\alpha}{6\pi}\partial_k\phi\right)\ .
\intertext{Therefore}
\partial_k\frac{\partial\mathscr{L}}{\partial(\partial_k\phi)} &= \partial_k\left((-E_k)\left[1-\frac{\alpha}{3\pi}\ln\frac{eE}{m^2}\right]+\left( \frac{\alpha}{6\pi}E_k\right)\right)
=-J_0\\
\text{and thus}\ &\vec{\nabla}\cdot\left[\left(1-\underbrace{\frac{\alpha}{6\pi}}_{\ll \frac{\alpha}{3\pi}\ln\frac{eE}{m^2}} - \frac{\alpha}{3\pi}\ln\frac{eE}{m^2}\right)\vec{E}\right] = J_0\ .
\end{align*}
Alltogether we have
\begin{align*}
\vec{\nabla}\cdot \vec{D} &= J_0 ,& \vec{D} &:=\epsilon(E)\vec{E}\\
\epsilon(E) &= 1-\frac{\alpha}{3\pi}\ln\frac{eE}{m^2} & \vec{E} &= -\vec{\nabla}\phi
\end{align*}
These are well-known classical equations from electrostatics of polarizable media. Looking back at the microscopic origin of $\epsilon(E)$,
we see that the effect of the vacuum fluctuations of the electron field is such that the vacuum responds to an external electric field as
if it were some sort of crystal which possesses a field-dependent dielectric constant. Obviously, Maxwell's equations become nonlinear due to the
logarithm in $\epsilon(E)$.
To summarize, we can say that in deriving $\vec{\nabla}\cdot \vec{D} = J_0$ with $\vec{D}=\epsilon(E)\vec{E}$, we have solved the problem of finding the nonlinear generalization of Maxwell's
equations - the non-linearities being caused by the electrons, which are hidden from direct observation but which influence the dynamics of the $A_\mu$ field
for a strong and static, but otherwise arbitrary, electrical field.
To gain some insight into the effect produced by the nonlinear medium $\epsilon(r)$, let us look at a specific example. We consider the case where $J_0$
contains only a single isolated charge $Q$ at $\vec{x}=\vec{0}$ (together with a compensating spherical shell of charge $-Q$ at infinity):
\begin{equation}
J_0(\vec{x}) = Q\delta(\vec{x})\ .
\end{equation}
Making the spherically symmetric ansatz
\begin{equation}
\vec{D}=\frac{Q}{4\pi r^2}\hat{r},\quad \vec{E} = \frac{Q(r)}{4\pi r^2}\hat{r},\quad r = |\vec{r}|,
\end{equation}
the equation
\begin{equation}
\vec{\nabla}\cdot \vec{D} = Q\delta(\vec{x})
\end{equation}
is solved, provided that the function $Q(r)$ is a solution to the transcendental equation
\begin{equation}
Q = Q(r)\epsilon\left(\frac{eQ(r)}{4\pi r^2}\right)\ .
\end{equation}
The physical interpretation of $Q(r)$ is that it is the charge lying within a sphere of radius $r$ centered at $\vec{x}=\vec{0}$. The value of $Q(r)$
is always larger than $Q$ because the vacuum polarization effects screen the charge. If we let $r\rightarrow\infty,\ Q(r)$ approaches the (macroscopically)
observed charge $Q$. We thus got an implicit equation for the modification of Coulomb's law by the electron fluctuations:
\begin{equation}
E(r) = \frac{Q(r)}{4\pi r^2}\ .
\end{equation}
One can show that when
\begin{equation}
\frac{eQ(r)}{4\pi}\gg 1\quad ,
\end{equation}
the approximation of neglecting the nonlocal one-loop contribution in the effective action becomes self-consistent. To see why this should be so, we note
that because of the leading local but nonlinear correction to $\epsilon$,
\begin{equation}
-\frac{\alpha}{3\pi}\ln X_1\ ,\quad X_1 = \frac{eQ(r)}{4\pi r^2m^2}\ ,
\end{equation}
while for the leading nonlocal correction
\begin{equation}
-\frac{\alpha}{3\pi}\ln X_2\ ,\quad X_2 = \frac{\left(\vec{\nabla}\ln A^0\right)^2}{m^2}\sim\frac{1}{r^2m^2}\ ,
\end{equation}
so that $\tfrac{eQ(r)}{4\pi}\gg 1$ is just the condition for $X_1 \gg X_2$, i.e., nonlinear local effects win out at short distances.
\section{Adler's Leading-Log Model in QCD}
Here, the leading approximation to the effective action is obtained by replacing $g^2$ in the classical Lagrangian
\begin{equation}
\mathscr{L} = \frac{1}{2g^2}\left(\vec{E}^2-\vec{B}^2\right)\ ,
\end{equation}
by the running coupling constant
\begin{align*}
\mathscr{L}^{eff} &= \frac{1}{2g^2_{running}}\left(\vec{E}^2-\vec{B}^2\right)\ ,
\intertext{where}
g^2_{running}\left(\frac{\vec{E}^2-\vec{B^2}}{\mu^4}\right) &= \frac{g^2(\mu^2)}{1+\frac{1}{4}b_0g^2(\mu^2)\ln\left(\frac{\vec{E}^2-\vec{B^2}}{\mu^4}\right)}\ ,
\intertext{so that}
\mathscr{L}^{eff} &= \frac{1}{2g^2}\left(\vec{E}^2-\vec{B}^2\right)\left[1+\frac{1}{4}b_0g^2(\mu^2)\ln\left(\frac{\vec{E}^2-\vec{B^2}}{\mu^4}\right)\right]\ .
\end{align*}
Here $\mu$ is an arbitrary subtraction point, $g^2=g^2(\mu^2)$, and $b_0$ is a certain constant one gets from calculating the one-loop radiative corrections,
\begin{equation}
b_0 = \frac{1}{8\pi^2}\frac{11}{3}C_{ad} > 0\ .
\end{equation}
Our formula for $\mathscr{L}^{eff}$ is applicable if $g^2_{running}$ is small:
\begin{align*}
\left(\vec{E}^2-\vec{B}^2\right) &\gg \mu^4\\
\vec{E},\vec{B} &\approx \vec{0}\ :\ g^2\text{ small, }g^2<0\ .
\end{align*}
The variational equation will be
\begin{equation}
\frac{\delta W^{eff}}{\delta(E,B)} = 0\quad ,
\end{equation}
and the static potential follows from
\begin{equation}
V_{static} = -W^{eff}_{extremum} + \text{self energies}\ .
\end{equation}
A reminder:
\begin{align*}
\mathscr{L} &= -\frac{1}{4}F_{\mu\nu}F^{\mu\nu}+j_\mu A^\mu\\
\vec{j}=\vec{0}\quad W &= \int\left[ \frac{1}{2}\left(\vec{E}^2-\vec{B}^2\right) - \rho\varphi\right] d^3x,\quad W=\frac{\text{action}}{T}\\
\frac{\delta W}{\delta B}=0\Rightarrow \vec{B}=\vec{0}\quad W &= \int\left[ \frac{1}{2}(\vec{\nabla}\varphi)^2-\rho\varphi\right] d^3x\\
\delta W &= \int\left[ \vec{\nabla}\varphi\cdot \vec{\nabla}\delta\varphi - \rho\delta\varphi\right] d^3x\\
&= \int \left(-\vec{\nabla}^2\varphi - \rho\right)\delta\varphi d^3x = 0\quad .
\end{align*}
Thus $\varphi_{extr.}$ satisfies $\vec{\nabla}^2\varphi_{extr.} = -\rho$.
\begin{align*}
W_{extr.} := W[\varphi_{extr.}] &= \int \left[\frac{1}{2}\vec{\nabla}\varphi\cdot \vec{\nabla}\varphi-\rho\varphi\right]d^3x\\
&= \int \left[-\frac{1}{2}\varphi\vec{\nabla}^2\varphi-\rho\varphi\right]d^3x\\
&= \int \left[-\frac{1}{2}\varphi(-\rho)-\rho\varphi\right]d^3x\\
&= -\frac{1}{2}\int \rho\varphi d^3x = -V_{static}+\text{self energies}.
\end{align*}
Before we continue, let us rewrite $\mathscr{L}^{eff}$ in a more compact form (here we follow Adler and use $F$ instead of $F^2$ so far).
Using $F:=\vec{E}^2-\vec{B}^2,\quad \vec{E} = -\vec{\nabla}\varphi,\quad \vec{B}=\vec{\nabla}\times\vec{A}$ we have
\begin{align*}
\mathscr{L}^{eff}(F) &= \frac{1}{8}b_0F\left[\frac{4}{g^2(\mu^2)b_0}+\ln\frac{F}{\mu^4}\right]\\
&= \frac{1}{8}b_0F\left[ -\ln e^{-\frac{4}{g^2(\mu^2)b_0} } +\ln\frac{F}{\mu^4} \right]\\
&= \frac{1}{8}b_0F\ln\frac{F}{\mu^4e^{-4/g^2(\mu^2)b_0} }\\
\mathscr{L}^{eff} &= \frac{1}{8}b_0\ln\frac{F}{e\kappa^2}\ ,
\intertext{where $\kappa^2$ is the constant}
\kappa^2 &:= \frac{\mu^4}{e}e^{-\frac{4}{b_0g^2(\mu^2)} }\ .
\end{align*}
This constant, $\kappa$, is a combination of $\mu$ and of $g^2(\mu^2)$. However, $\kappa^2$ is renormalization-group invariant (to one-loop order), so that
$\kappa$ is a physical parameter, whereas $\mu$ is an unphysical parameter. We recall:
\begin{equation}
\beta(g) = \mu\frac{\partial g}{\partial\mu} = -\frac{1}{2}g^3b_0\quad .
\end{equation}
Then we obtain
\begin{align*}
\frac{d}{d\mu}\kappa^2 &= \frac{1}{e}\left[ 4\mu^3e^{-\frac{4}{b_0g^2}}+\mu^4\frac{d}{d\mu}\left(-\frac{4}{b_0g^2}\right)e^{-\frac{4}{b_0g^2}}\right]\\
&= \frac{1}{e}e^{-\frac{4}{b_0g^2}}\left[4\mu^3-\frac{4\mu^4}{b_0}(-2)g^{-3}\frac{d g}{d\mu}\right]\\
&= \frac{1}{e}e^{-\frac{4}{b_0g^2}}4\mu^3\left[1+\frac{2}{b_0g^3} \underbrace{\mu\frac{\partial g}{\partial\mu}}_{-\frac{b_0}{2}g^3} \right] = 0\quad .
\end{align*}
The graph of $\mathscr{L}^{eff}$ is shown in fig. \ref{fig:EffLagrangian}.
\begin{figure}
\centering
\includegraphics[width=0.6\linewidth]{EffectiveLagrangianImage.eps}
\caption{Schematic graph of $\mathscr{L}^{eff}$. The dashed line represents classical action.}
\label{fig:EffLagrangian}
\end{figure}
The minimum of $\mathscr{L}^{eff}$ is given by
\begin{align*}
\left( F\ln\frac{F}{e\kappa^2}\right)' &= \left( F\ln F - F\ln e\kappa^2\right)' = \ln F+1-\ln e\kappa^2\\
&= \ln F -\ln\kappa^2 = 0\\
\Rightarrow F_{min} &= \kappa^2 = \frac{\mu^4}{e}e^{-\frac{4}{b_0g^2(\mu^2)} },\quad b_0>0\quad .
\end{align*}
The renormalization group argument says that $\mathscr{L}^{eff}(F)$ is a good approximation in a region for strong fields and a region close to the origin for very weak fields.
Around the minimum, the approximation is not reliable. But what is reliable is the fact that the minimum is away from the origin where the interesting structure of the model comes from.
Now we have
\begin{equation}
W^{eff} = \int\left[ \mathscr{L}^{eff} - \rho\varphi\right] d^3x\ ,
\end{equation}
with
\begin{equation}
\rho(\vec{x}) = Q\left[\delta^{(3)}(\vec{x}-\vec{x}_1) - \delta^{(3)}(\vec{x}-\vec{x}_2)\right]\ .
\end{equation}
The variational equations come from
\begin{align*}
& W^{eff} &= \int\left[ \mathscr{L}^{eff}(\vec{\nabla}\varphi, \vec{A})-\rho\varphi\right] d^3x & \\
\mathscr{L} \equiv \mathscr{L}^{eff}:\quad \vec{\nabla}\cdot \frac{\partial\mathscr{L}}{\partial(\vec{\nabla}\varphi)}-\frac{\partial\mathscr{L}}{\partial\varphi} &= 0 &
\frac{1}{2} F = \frac{1}{2}\left(\vec{E}^2-\vec{B}^2\right) &= \frac{1}{2}\left[(\vec{\nabla}\varphi)^2-\left(\vec{\nabla}\times\vec{A}\right)^2\right]\\
\text{or}\quad\vec{\nabla}\cdot\underbrace{\frac{\partial\mathscr{L}}{\partial\frac{1}{2}F}}_{=:\epsilon}\underbrace{\frac{\partial \frac{1}{2}F}{\partial(\vec{\nabla}\varphi)}}_{=-\vec{E}}+\rho &= 0 &
\frac{\partial\frac{1}{2}F}{\partial(\vec{\nabla}\varphi)} = \vec{\nabla}\varphi &= -\vec{E}\\
\Longrightarrow \vec{\nabla}\cdot\epsilon\vec{E} &= \rho & &
\end{align*}
\begin{align*}
\frac{\delta\mathscr{L}(x)}{\delta A_i(z)} &= \int \frac{\partial\mathscr{L}(x)}{\partial B_j(x)}\frac{\delta B_j(x)}{\delta A_i(z)}dx\\
&= \int \frac{\partial\mathscr{L}(x)}{\partial\frac{1}{2}F(x)}\frac{\partial\frac{1}{2}F}{\partial B_j(x)}\frac{\delta}{\delta A_i(z)}\epsilon_{jmn}\partial_m A_n(x)dx\\
&= \int (-\epsilon B_j)(x)\epsilon_{jmn}\delta_{in}\partial_m\delta(x-z)dx\\
&= \int \partial_m(\epsilon B_j)(x)\epsilon_{jmi}\delta(x-z)dx\\
&= \epsilon_{ijm}\partial_m(\epsilon B_j)(z) = -\left[\vec{\nabla}\times(\epsilon\vec{B})\right]_i = 0\\
\Rightarrow\quad \vec{\nabla}\times(\epsilon\vec{B}) &= \vec{0}
\end{align*}
\begin{align}
\epsilon(F) := \frac{\partial\mathscr{L}^{eff}}{\partial(\frac{1}{2}F)} &= \frac{1}{4}b_0\left(F\ln\frac{F}{e\kappa^2}\right)'\notag\\
\Rightarrow\quad \epsilon(F) &= \frac{1}{4}b_0\ln\frac{F}{\kappa^2}\ .\label{eq:PermettivityFunction}
\end{align}
\begin{figure}
\centering
\includegraphics[width=0.5\linewidth]{PermettivityFunctionImage.eps}
\caption{Scaled permettivity function $\epsilon$ as given in Eq. \eqref{eq:PermettivityFunction}.}
\end{figure}
Now the equation
\begin{equation}
\vec{\nabla}\times(\epsilon\vec{B}) = \vec{0}
\end{equation}
is solved by $\epsilon\vec{B}=\vec{0}$.
Thus there are two branches we have to consider here:
\begin{align}
\vec{B} &= \vec{0},\quad \text{ or }\\
\epsilon &= 0,\quad \text{i.e., at }F=\vec{E}^2-\vec{B}^2=\kappa^2\ .
\end{align}
Near the source charges, where the fields are strong, asymptotic freedom tells us that the solution will look like the Abelian case. This means the electric field be big
and the magnetic field should be zero, or small:
\begin{align*}
\vec{B}=\vec{0}\ :\quad F=\vec{E}^2 &= (\vec{\nabla}\varphi)^2\ ,\ \vec{E}=-\vec{\nabla}\varphi\ .
\intertext{Now define}
\vec{D} &=\epsilon(\vec{E}^2)\vec{E} = \epsilon((\vec{\nabla}\varphi)^2)\vec{E}\quad ,
\intertext{together with}
\vec{\nabla}\cdot\vec{D} &= \rho\\
\vec{\nabla}\times\vec{E} &= \vec{0}\ .
\end{align*}
Thus, we now have a non-linear dielectric problem.
As S. Adler\cite{Adler1981} has shown, the leading-log model gives us a qualitatively correct, and semiquantitatively accurate account of the $q\bar{q}$ force. He shows in a highly non-trivial
calculation that as the distance between the quarks $R=|\vec{x}_1-\vec{x}_2|\rightarrow 0,\infty$, one obtains
\begin{align*}
V_{static} (R) &\rightarrow \kappa QR + \mathcal{O}\left(\sqrt{\kappa}\ln(\sqrt{\kappa}R)\right),\quad R\rightarrow\infty\\
V_{static}(R) &\rightarrow -\frac{Q^2}{4\pi R\frac{1}{2}b_0}\left[\frac{1}{\ln\left(\frac{1}{\Lambda_p^2R^2}\right)}+\mathcal{O}\left(\frac{\log\log}{\log^2},\frac{1}{\log^3}\right)\right],\
R\rightarrow 0\ ,
\end{align*}
with $\Lambda_p = 2.52\sqrt{\kappa}$ for the parameter values $Q=\sqrt{\frac{4}{3}},b_0 = \frac{9}{8\pi^2}$ appropriate to $SU(3)$ with $3$ light quark flavors.
|
\section{Introduction \label{introduction}}
It has been shown that the global star-formation rate in the Universe gradually increases
from z$\sim$10 to z$\sim$2-3 and then steeply decreases till the present epoch,
$z=0$ \citep[see e.g.][and references therein]{Dunlop2011}. Because metals are
produced by stars, the determination of metal abundance in the gas provides complementary information
about star-formation history \citep{Rafelski2012}. This can be done using Damped Lyman-$\alpha$
absorption systems (DLAs) that represent the main reservoir of
neutral gas at high redshift \citep{Prochaska2009, Noterdaeme2009} and are likely to be located
in galaxies or in their close environment \citep[e.g.][]{Krogager2012}.
DLAs arise mostly in
the warm neutral medium \citep[e.g.][]{Petitjean2000, Kanekar2014} and have a multicomponent velocity
structure, with metal absorption lines spread typically over 100-500 km\,s$^{-1}$
\citep{Ledoux1998}. In a small fraction of DLAs, the line of sight intercepts
cold gas, as traced by molecular hydrogen \citep[e.g.][]{Noterdaeme2008,Noterdaeme2011,Balashev2014}
and/or 21-cm absorption \citep[e.g.][]{Srianand2012}.
Important progress has been
made towards understanding the properties of the gas (through e.g. deriving physical
conditions \citealt{Srianand2005,Noterdaeme2007, Jorgenson2009} and physical extent \citealt{Balashev2011})
and the incidence of cold gas in DLAs has been related to other properties (such
as the metallicity \citep{Petitjean2006} or the dust content \citep{Ledoux2003, Noterdaeme2008}).
However, due to the strong saturation of H\,{\sc i} Lyman series lines, it remains impossible to directly determine the H~{\sc i} column density associated with the individual cold gas components traced by H$_2$ absorption.
Even for metals, whose absorption lines are not saturated, it is very difficult to determine what fraction originates from the cold phase.
Difficulties arise as well with the 21-cm absorption that do not always exactly coincide
with H$_2$ absorption \citep{Srianand2013}
although it could be due to the different structures of the optical and radio
emitting regions of the background quasars.
Out of all the metals, chlorine shows a unique behavior in the presence of H$_2$.
Because the ionization potential of chlorine (12.97 eV) is less than that of atomic hydrogen,
chlorine is easily ionized in the diffuse neutral medium. However, this species reacts exothermically with H$_2$ at a very high rate converting rapidly Cl$^+$ into HCl$^+$. The latter subsequently releases neutral chlorine through several channels \citep{Jura1974,Neufeld2009}. This process is so efficient that chlorine
is completely neutral in presence of a small amount of H$_2$.
In our Galaxy the fact that chlorine abundance anti-correlates with the average number density along the line of sight \citep{Harris1984, Jenkins1986} has been interpreted as chlorine depletion. However, models predict as well as observations indicate, that gas with moderate dust content presents negligible depletion of chlorine \citep[e.g.][]{Neufeld2009, Savage1996, Jenkins2009}.
Observationally, a tight relation is indeed found between Cl~{\sc i} and H$_2$ in the local
ISM \citep{Jura1974, Sonnentrucker2006, Moomey2012}. In this letter, we present the
first study of this relation at high redshift and over a wide range of column
densities.
\section{Data sample and measurements}
Since the first detection by \citet{Levshakov1985}, about two dozen H$_2$ absorption
systems have been detected at high redshifts in QSO spectra.
The detection limit of the strongest Cl\,{\sc i} absorption line (1347\AA, f=0.0153 \citep{Schectman1993}) in high quality spectrum (${\rm S/N}\sim 50$, ${\rm R}\sim 50\,000$) corresponds to $N(\mbox{\ClI}) \sim 10^{12}$\,cm$^{-2}$. The solar abundance of chlorine is $10^{-6.5}$ that of hydrogen \citep{Asplund2009} and given previous measurements of N(\ClI)/2N(H$_2$) \citep[e.g.][]{Moomey2012} we conservatively limit our study to systems with $N({\rm H}_2) \gtrsim 10^{17}$~cm$^{-2}$.
Redshifts, H\,{\sc i} and H$_2$ column densities, and metallicities were mainly taken from the literature
and are based on VLT/UVES, Keck/HIRES or HST/STIS data.
We refitted H$_2$ absorption systems towards Q\,2123$-$0050 and Q\,J2340$-$0053
to take into account the positions of the detected Cl~{\sc i} components.
We also detect a new H$_2$ absorption system in the $z=3.09$ DLA towards J\,2100$-$0641 in
which \citet{Jorgenson2010} have reported the presence of neutral carbon. C\,{\sc i} is indeed known
to be an excellent indicator of the presence of molecules \citep[e.g.][]{Srianand2005}.
We used the MAKEE package (T. Burles) to reduce archival data from this quasar obtained in 2005, 2006 and 2007 under
programs U17H (PI: Prochaska), G400H (PI: Ellison) and U149Hr (PI: Wolfe). We have found strong H$_2$ absorption
lines from rotational levels up to $J=5$ (see Fig.~\ref{J2100_H2}) at $z=3.091485$ with a total column
density of $\log N($H$_2) = 18.76 \pm 0.03$.
For all systems we retrieved data from the VLT/UVES or the Keck/HIRES archives.
We reduced the data and fitted the lines using profile fitting.
Neutral chlorine is detected in nine DLAs (Fig~\ref{new}). Four detections were already reported
in the literature: Q\,1232+082 \citep{Balashev2011}, Q\,0812$-$3208 \citep{Prochaska2003},
Q\,1237$+$0647 \citep{Noterdaeme2010} and Q\,2140-0321 (Noterdaeme et al. submitted). The
remaining five are new detections.
We measured upper-limits of N(\ClI) for the remaining nine systems.
We used mainly the $1347$\AA\ \ClI\ line.
Whenever possible, we also used \ClI\ lines at $1088$\AA, $1188$\AA, $1084$\AA, $1094$\AA\
and $1085$\AA, with oscillator strength from respectively \citet{Schectman1993}, \citet{Morton2003}, \citet{Morton2003}, \citet{Sonnentrucker2006} and \citet{Oliveira2006}.
Table~\ref{table_results} summarizes the results of \ClI\ measurements. We have kept all components with log~$N$(H$_2$)~$>$~17.
We did not use two known H$_2$ absorption systems towards Q\,0013$-$0029 and J\,091826.16$+$163609.0 since H$_2$ column densities in these systems are not well defined.
\begin{figure}[t]
\begin{center}
\includegraphics[clip=,width=0.95\hsize]{J2100_H2.eps}
\caption{Voigt profile fits to the newly detected H$_2$ absorption lines from rotational levels J=0 to J=5 at $z=3.091485$
towards J\,2100$-$0641. The column densities are indicated (in log(cm$^{-2}$)) in the
top right corner of each panel.}
\label{J2100_H2}
\end{center}
\end{figure}
\begin{figure}[t]
\begin{center}
\includegraphics[clip=,width=0.95\hsize]{Cl_lines_3x3.eps}
\caption{Voigt profile fits to \ClI\ $\lambda$1347 absorption lines associated with high-$z$ strong H$_2$ absorption systems. }
\label{new}
\end{center}
\end{figure}
\input table.tex
\section{Results}
\begin{figure}
\centering
\includegraphics[bb=62 178 490 570,clip=,width=0.95\hsize]{ClH2.ps}
\caption{Column densities of Cl\,{\sc i} versus that of H$_2$. The red and blue points indicate the measurements at high redshift (this work)
and in our Galaxy \citep{Moomey2012}, respectively. The straight (dashed, dotted) lines show the respective least-squares bisector fits to the data. \label{resn}}
\end{figure}
Fig.~\ref{resn} shows the Cl\,{\sc i} column density, $N$(\ClI), versus $N$(H$_2$) and
compares our high-$z$ measurements to those obtained in the local ISM
using the Copernicus satellite \citep{Moomey2012}. As can be seen our high-$z$ measurements extend the
relation to H$_2$ column densities ten times smaller than those measured in the local ISM.
Cl\,{\sc i} and H$_2$ are found to be very well correlated ($r=0.95$) over the entire
$N$(H$_2$) range.
It is striking that measurements at high and low redshifts are indistinguishable in the overlapping regime ($\log N($H$_2) \sim $19-20.2).
The correlation is seen over about three orders of magnitude in column density with a dispersion of
about 0.2~dex only. A least-squares bisector linear fit provides a slope of 0.83 and 0.87 for the high-$z$ and
$z=0$ data, respectively, with an almost equal normalization ($\log N(\ClI) \approx 13.7$ at $\log N($H$_2)=20$).
We note that the upper limits on $N$(\ClI) lie mostly at the low $N($H$_2)$ end and are least constraining
since they are compatible with the values expected from the above relation.
For this reason, we will not consider them further in the discussion but still include them in the
figures for completeness.
The slopes are less than one, meaning that the \ClI/H$_2$ ratio slightly
decreases with increasing $N($H$_2)$. This is unlikely to be due to conversion of Cl into H$_2$Cl$^+$ and/or HCl, since chlorine chemistry models \citep{Neufeld2009} as well as measurements (e.g. towards Sgr B2(S), \citealt{Lis2010}) show that in diffuse molecular clouds only $\sim1$\% of chlorine is in the molecular form.
A $<1$ slope could in principle be due to dust depletion. However, there is no trend for increasing Cl depletion with increasing H$_2$ or \ClI\ column densities in Galactic clouds (see \citealt{Moomey2012}).
In addition, for high redshift measurements, elemental abundance pattern \citep{Noterdaeme2008} as well as direct measurements \citep[e.g.][]{Noterdaeme2010} indicate $A_v < 0.2$ when modeling of chlorine chemistry \citep{Neufeld2009} shows that for such low extinction (A v < 1) all chlorine is in the gas phase.
A possibility is thus that the molecular fraction in the gas probed by \ClI\ is slightly increasing with
increasing $N$(H$_2)$. It can be expected since H$_2$ self-shielding increase while \ClI\ is already completely in the neutral form. Finally, the similarity between our Galaxy and high-$z$ measurements at $\log N($H$_2)\ge 19$ might indicate that the chemical and physical conditions in the cold gas can be similar, otherwise fine tuning would be required between the different factors that impact on the $N$(\ClI) to $N$(H$_2$) ratio (e.g. number density, metallicity, dust content and UV flux).
Before continuing further, we note that in H$_2$-bearing gas, chlorine is found exclusively in the neutral form
(i.e. $N($Cl$) = N(\ClI)$) \citep[e.g.][]{Jura1974}.
Since we expect that all chlorine is in gas-phase\footnote{ We note that the presence of \ClII\ in the outer envelope of the H$_2$ cloud is not excluded, but it does not influence our derivation}, the abundance of chlorine, [Cl/H], in H$_2$-bearing gas can be expressed as
\begin{equation}
{\rm [Cl/H]} = {\rm [\mbox{\ClI}/H_2]} + \log{f},
\label{Metalfl1}
\end{equation}
where
\begin{equation}
{\rm [\mbox{\ClI}/H_2]} = \log\left( \frac{N({\mbox{\ClI})}}{2N({\rm H}_2)} \right) - \log\left(\frac{{\rm Cl}}{{\rm H}}\right)_{\odot}
\label{ClH2}
\end{equation}
\noindent and $f=2N($H$_2)/(2N($H$_2)+N($H\,{\sc i}$))$ is the molecular fraction.
Therefore the ratio \ClH\ gives a direct constraint on the chlorine-based metallicity of H$_2$-bearing gas provided
the molecular fraction of H$_2$-bearing gas is known.
Conversely, if a constraint can be put on the actual chlorine abundance, \ClH\ can provide an estimate of the amount of H\,{\sc i} present
in H$_2$-bearing gas.
If Cl is depleted onto dust grains then the mentioned estimates of metallicity and molecular fraction will have to be corrected from the Cl depletion factor.
The Fig.~\ref{res} shows \ClH\ as a function of the
{\sl overall} metallicity for DLAs (given in Table~\ref{table_results}) at high redshift or as a function of [Cl/H] for clouds in our Galaxy.
Since $f\le 1$, \ClH\ gives an upper limit on the metallicity in H$_2$-bearing gas,
which is found to be roughly equal or less than solar metallicity. For 13 out of 21 \ClI\ bearing clouds in our Galaxy
associated \ClII\ was measured \citep{Moomey2012}. Therefore we have estimated the overall chlorine abundance [Cl/H]
of these clouds as $(N(\mbox{\ClI}) + N(\mbox{\ClII}))/(N(\mbox{\HI}) + 2N(\mbox{H}_2))$ (shown by blue circles in Fig.~\ref{res}).
Unfortunately, for high redshift DLAs, not only \ClII\ is not detected but also
DLAs contain several \HI\ clouds so that chlorine abundance of the very cloud of interest cannot be measured. Therefore we consider the overall metallicity (averaged over velocity components) measured using another non-depleted
element (usually Zn or S, see Table~\ref{table_results}). In Fig.~\ref{res} it can be seen that the \ClH\ ratio is likely not correlated with the {\sl overall} metallicity of the DLA (Pearson correlation coefficient 0.3 at 0.3 significance level).
For Milky-Way clouds it can be seen that [Cl/H] is typically one third solar, which can be interpreted
as evidence for chlorine depletion \citep{Moomey2012}.
The large difference between \ClH\ and [X/H]$_{\rm DLA}$ for the high redshift clouds can be explained either
by a molecular fraction $f<1$ in H$_2$-bearing clouds or by a higher metallicity in the
H$_2$-bearing gas compared to the overall DLA metallicity or by both effects. If we assume that the
metallicity in the H$_2$-bearing gas is equal to the overall DLA metallicity we find (using Eq.~\ref{Metalfl1})
that the molecular fraction in the H$_2$-bearing gas is typically an order of magnitude higher than
the overall inferred DLA molecular fraction.
Interestingly, two systems sitting close to the one-to-one relation are those where CO molecules
have been detected (Q\,1439+1118 and Q\,1237+0647).
In such systems, the H$_2$ component is probably fully molecularized and its metallicity is close to the overall metallicity of the DLA.
\setlength{\tabcolsep}{0pt}
\begin{figure}
\centering
\begin{tabular}{c}
\includegraphics[bb=62 178 490 570,clip=,width=0.8\hsize]{ClIH2_XH.ps} \\
\end{tabular}
\caption{\ClH\ as a function of the {\sl overall} metallicity for high-$z$ DLAs (red points) and the
chlorine-based metallicity for Milky-Way clouds (blue points). The dashed line represents the one-to-one relation.
\label{res}}
\end{figure}
\setlength{\tabcolsep}{3pt}
\section{Conclusion}
\noindent
We have studied the neutral chlorine abundance in high redshift (z$\sim$2\,-\,4)
strong H$_2$ bearing DLAs with log~$N$(H$_2$)~$>$~17.3.
These systems arise in the cold neutral medium of galaxies in the early Universe.
We have used 17 systems from the literature
and also present a new H$_2$ detection at $z=3.09145$ in the spectrum of J\,2100$-$0641.
We have detected \ClI\ absorption lines in half of these systems, including 5 new detections. The derived upper limits for N(\ClI)
for the remaining systems are shown to be consistent with the behavior of the
overall population.
Our measurements extend the \ClI-H$_2$ relation to lower column densities than
measurements towards nearby stars. We show that there is a 5\,$\sigma$ correlation between the column densities of both
species over the range 18.1~$<$~log~$N$(H$_2$)~$<$~20.1 with
indistinguishable behavior between high and zero redshift systems.
This suggests that at a given $N($H$_2$) the physical conditions are likely similar in our Galaxy and high-$z$ gas, in spite of possible differences in the dust depletion levels.
As we expect the Cl to be depleted less in the high-z absorbers studied here, we use the abundance of chlorine with
respect to H$_2$ to constrain the molecular fraction and the metallicity in H$_2$-bearing gas.
Our results suggest that the molecular fraction and/or the metallicity in the H$_2$ and \ClI\ bearing components could be
much higher than the mean value measured over the whole DLA system. This implies
that a large fraction of \HI\ is unrelated to the cold phase traced by H$_2$.
Finally, our understanding of the formation of H$_2$ onto dust grains, self-shielding and lifetime
of cold diffuse gas would certainly benefit from further observations of chlorine and molecular
hydrogen in different environments and over a wide range of column densities.
\vspace{2mm}{\footnotesize {\rm Acknowledgments.}
SB and VK thank RF President Program (grant MK-4861.2013.2) and ``Leading Scientific Schools of Russian Federation'' (grant NSh-294.2014.2).
RS and PPJ gratefully acknowledge support from the Indo-French Centre for the
Promotion of Advanced Research (Centre Franco-Indien pour la
Promotion de la Recherche Avanc\'ee) under contract No. 4304-2.
\bibliographystyle{aa}
|
\section{PARAMETERISATION OF EFFECTIVE POTENTIAL}
\label{sec:effpot}
\subsection{Theoretical motivation}
As described in the main text, we use an effective potential that captures the important features of the
depletion interaction between lock-and-key colloids.
The effective interaction potential is $v(\bm{r}_{12},\bm{n}_1,\bm{n}_2)$
where the vector between two particles is $\bm{r}_{12}$ and their orientations are $\bm{n}_{1,2}$.
Our goal is to parameterise the function $v$
so that our simplified model accurately represents the behaviour of a mixture
of hard indented colloids and ideal depletant particles. The optimal effective model would match
two-body interactions exactly:
\begin{equation}
\exp[-\beta v(\bm{r}_{12},\bm{n}_1,\bm{n}_2)] = {\tilde g}_0(\bm{r}_{12},\bm{n}_1,\bm{n}_2)
\label{equ:eff}
\end{equation}
where ${\tilde g}_0$ is the two-particle distribution function for the colloids, normalised so that $\tilde{g}_0=1$ at
large distances.
The subscript `0' indicates that this function is calculated in a dilute limit, where
the colloid density is taken to zero (at fixed depletant volume fraction $\eta_{\rm s}$).
The function ${\tilde g}_0$ can be obtained computationally, but it
depends on the distance $r_{12} = |\bm{r}_{12}|$ and on three different angles. Parameterising
this non-trivial function of a four variables is rather difficult -- we have used numerical simulation to identify the most
important features of this function, which we incorporate into a simplified effective potential.
The radial distribution function in the dilute limit may be obtained as
\begin{equation}
g_0(r) = \int \frac{\mathrm{d}\bm{n}_1 \mathrm{d}\bm{n}_2}{(4\pi)^2}\, {\tilde g}_0(\bm{r}_{12},\bm{n}_1,\bm{n}_2)
\label{equ:gr}
\end{equation}
where both integrals run over the entire unit sphere. Fig.~\ref{fig:grmap} shows an example of this function,
in a situation where both lock-key and back-to-back binding are signficant. The two separate peaks indicate
the different binding mechanisms, but this representation does not reveal the orientation dependence of the
interaction.
Fig.~\ref{fig:coords} shows the choice of co-ordinate system that we have used to capture this dependence. In lock-key binding, one particle
plays the role of the ``lock'' and the other the ``key''. The definition $\cos\theta_{\rm R} =
\max( \bm{n}_1\cdot \bm{r}_{12},-\bm{n}_2\cdot \bm{r}_{12})$ takes the smaller of the angles
between $\bm{r}_{12}$ and the directors $\bm{n}_{1,2}$; the particle associated with this smaller angle plays the part of
the lock. We define $\cos\theta_{\rm R}^c = d_c/\sigma$ so that $\theta_{\rm R}^c$ corresponds to angle
between the director and a vector pointing out through the ``lip'' of the lock mouth. Lock-and-key binding happens only for $\theta_{\rm R}<\theta_{\rm R}^c$
while back-to-back binding happens for $\theta_{\rm R}>\theta_{\rm R}^c$.
Since back-to-back binding takes place between two convex surfaces, this part of the interaction depends weakly on angular
co-ordinates, so the effective potential includes a square well of range $r_{\rm BB} = 1.1\sigma$ in that case. The contribution
to the effective potential $v$ is therefore $v_{\rm BB} =-\varepsilon_{\rm BB} \Theta( \theta_{\rm R} - \theta_{\rm R}^c) \Theta(1.1\sigma-r_{12})$
where $\Theta(x)$ is the step function.
For lock-and-key
bonding, the angular dependence is less simple, but the dominant effect can be accounted for through the angle $\phi$,
which measures the rotation of the ``key'' particle within the mouth of the lock. Hence the interaction
potential for lock-key binding is $v_{\rm LK}=-\varepsilon_{\rm LK} W(\phi) \Theta( \theta_{\rm R}^c - \theta_{\rm R}) \Theta(d_c+0.1\sigma-r_{12})$ where $W(\phi)$ is the piecewise linear function of $\cos\phi$ shown in Fig.~\figpot~of the main text.
In fact, this simple function provides an almost quantitative description of the $\phi$-dependence of the effective potential~{[38]}.
\begin{figure}
\includegraphics[width= 0.75 \columnwidth]{grmap}
\caption{Radial distribution function for two indented colloids, with depletant size ratio, $q=0.2$, and $\eta_{\rm s}=0.30$. (Recall $d_c=0.6\sigma$.)
A thin dashed line shows the corresponding function in the absence of depletant. }
\label{fig:grmap}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width= \columnwidth]{coords}
\caption{Co-ordinate system for interactions among indented colloids. The angles between the interparticle vector $\bm{r}_{12}$ and the directors $\bm{n}_{1,2}$
are labelled as shown: the labels $\theta_{\rm R,I}$ are chosen such that $\theta_{\rm R} < \theta_{\rm I}$.
The angle between the directors, is $\phi$.}
\label{fig:coords}
\end{center}
\end{figure}
\subsection{Numerical parameterisation}
Having specified the form of the effective potential, we now describe how the parameters $(\varepsilon_{\rm LK},\varepsilon_{\rm BB})$ are chosen.
The (effective) second virial coefficient for the indented colloids in the presence of the depletant is~{[39]}
\begin{equation}
B_2 = 2\pi
\int_0^\infty \mathrm{d}r\, r^2 [1-g_0(r)]
\end{equation}
We split $B_2$ into two pieces, associated with lock-key and back-to-back binding:
\begin{align}
B_2^{\rm LK} &= 2\pi
\int_0^{\sigma} \mathrm{d}r\, r^2 [1-g_0(r)]
\nonumber \\
B_2^{\rm BB} &= 2\pi
\int_{\sigma}^\infty \mathrm{d}r\, r^2 [1-g_0(r)]
\label{equ:B2split}
\end{align}
These two parameters indicate the strength of the two kinds of binding. A representative plot of $g_0(r)$
is shown in Fig.~\ref{fig:grmap}. The peak at $r\approx d_c$ is associated with lock-key binding, and its area sets
the strength of this interaction. The peak at $r\approx\sigma$ is associated with back-to-back binding.
From (\ref{equ:gr}), $B_2^{\rm LK}$ and $B_2^{\rm BB}$ can be written as integrals of $1-\tilde{g}_0$. Motivated by (\ref{equ:eff}),
our procedure is therefore to choose $\varepsilon_{\rm LK},\varepsilon_{\rm BB}$ so that the corresponding integrals of
$1-\exp[-\beta v(\bm{r}_{12},\bm{n}_1,\bm{n}_2)]$ are equal to $B_2^{\rm LK},B_2^{\rm BB}$. These integrals
are contributions to the second virial coefficient of the effective model, so we denote them
by $B_{2,{\rm eff}}^{\rm LK}$ and $B_{2,{\rm eff}}^{\rm BB}$.
Since the potential is a square well and the hard-particle interactions have range at most $\sigma$, it is easily verified
that
\begin{align}
B_{2,{\rm eff}}^{\rm LK} &= v_{\rm m} - v_{\rm LK} ( 1 - \mathrm{e}^{\varepsilon_{\rm LK}} )
\nonumber \\
B_{2,{\rm eff}}^{\rm BB} &= v_{\rm BB} ( 1 - \mathrm{e}^{\varepsilon_{\rm BB}} )
\label{equ:b2eff}
\end{align}
where $v_{\rm m}$ accounts for the excluded volume between two indented colloids,
and $v_{\rm LK},v_{\rm BB}$ are phase space volumes associated with
the two binding modes.
\begin{figure}
\includegraphics[width= 0.75 \columnwidth]{lk-eta}
\includegraphics[width= 0.75 \columnwidth]{bb-eta}
\caption{The contributions to the second virial coefficient for a system of two locks, interacting with a depletant of penetrable spheres.}
\label{fig:etalb2}
\end{figure}
\begin{figure}
\includegraphics[width= 0.75 \columnwidth]{lk-eff}
\includegraphics[width= 0.75 \columnwidth]{bb-eff}
\caption{Contributions [see (\ref{equ:B2split})] to the second virial coefficient for a system of two indented colloids interacting via the effective potential.
Dashed lines are fits to (\ref{equ:b2eff}) with $v_{\rm m}=1.94\sigma^3$, $v_{\rm LK}=(3.3\times10^{-4})\sigma^3$ and $v_{\rm BB}=0.44\sigma^3$}
\label{fig:lb2}
\end{figure}
\subsection{Computational implementation}
To estimate $g_0(r)$ for the system with depletant, we use the Geometrical Cluster Algorithm (GCA) to simulate two colloidal
particles in a box, mixed with depletant, as in~{[17]}~(this procedure accounts for finite-size corrections to $g(r)$).
Representative results for $B_2^{\rm LK},B_2^{\rm BB}$ are shown in Fig.~\ref{fig:etalb2}, for several size ratios $q$.
As expected, these parameters become large and negative as the nanoparticle volume fraction $\eta_{\rm s}$ increases.
The corresponding quantities $B_{2,{\rm eff}}^{\rm LK},B_{2,{\rm eff}}^{\rm BB}$ were calculated numerically
by the same method, but using the effective potential instead of an explicit depletant. Results are shown in Fig.~\ref{fig:lb2}, including
fits to (\ref{equ:b2eff}), from which we estimate the small volume associated with lock-and-key binding to be $v_{\rm LK} = (3.3\times 10^{-4})\sigma^3$.
Comparison of Figs~\ref{fig:etalb2} and~\ref{fig:lb2} allows straightforward mapping from the depletion
parameters $(q,\eta_{\rm s})$ to the parameters $(\varepsilon_{\rm LK},\varepsilon_{\rm BB})$ of the effective potential.
\section{SIMULATION METHOD}
\label{sec:sim}
\subsection{Biased insertions in the grand canonical ensemble}
The lock-key interactions in our system are typically strong, implying
a strongly negative chemical potential and consequently a very low
acceptance rate for standard particle transfers (insertions and
deletions) in grand canonical Monte Carlo (GCMC) simulations. To counter this we introduce an additional biased transfer
update in addition to the standard one. This operates by
preferentially attempting to insert an indented particles in a bound configuration around a target colloid.
To implement the ``biased insertion'', we first pick a target colloid randomly from
the existing ones in the box. The location of the centre of the colloid to be
inserted is then chosen randomly from within a narrow spherical shell
surrounding the target. The shell radii are chosen to represent the
typical separation of bound colloids. For a given point in the shell
there will be a range of solid angle for the particle director which
avoids hard particle overlaps, and we choose the director of the new
particle uniformly from within this range. The biasing factor needed
to maintain detailed balance is therefore a product of the shell
volume and the solid angle that avoids overlap. This biasing factor
can also be calculated for a bound particle which is to be removed
from the system, based on its separation from its bound neighbour, and
so enters symmetrically in the grand canonical acceptance
probabilities.
\subsection{Determination of binodals, critical points and field mixing}
\begin{figure}
\begin{center}
\includegraphics[width= 0.98 \columnwidth]{critical}
\caption{(a) Critical point order parameter distributions $P(x)$ with
$x=a(\rho-\rho_c))$. The figure shows data for $q=0.2$ and $q=0.126$
and the distributions are scaled to unit norm and variance via the
choice of the parameter $a$. Also shown for comparison is the fixed
point Ising magnetisation distribution $P^\star(m)$. (b) The
corresponding normalised distribution of the field mixed order
parameter $P(y)$ with $y=b({\cal M}-{\cal M}_c)$ and $b$ chosen to
scale the distributions to unit variance. Box sizes were
$\ell=12\sigma$ for the $q=0.126$ system and $\ell=7.5\sigma$ for
the $q=0.2$ system.}
\label{fig:critical}
\end{center}
\end{figure}
Our GCMC simulations sample the fluctuations in the colloid number density
$\rho=N/V$ and energy density $u=E/V$. Well within the two-phase
region, the probability distribution $P(\rho)$ exhibits a pair of
widely separated peaks centered on the coexistence
densities. Accordingly $\rho$ serves as an order parameter for the
transition, and the coexistence densities (binodal) can be simply read
off from the peak positions~{[29]}. Closer to the
liquid-vapor critical point, the fluctuations in $\rho$ and $u$ assume
a universal character which can be probed via the properties of the
joint distribution $P(\rho,u)$~{[29,30]}. In models
that exhibit particle hole symmetry, the order parameter distribution
at criticality matches the (independently known) universal fixed point
form of the Ising magnetisation distribution $P^\star(m)$. Effecting
this matching provides a means of estimating the critical
parameters. However, since our model lacks particle-hole symmetry,
field mixing is expected to occur. This implies that the scaling
equivalent of $m$ is the linear combination ${\cal M}=\rho-su$
{[29,30]}. (Strictly speaking, a non-linear combination
may be appropriate~{40]}, but the linear combination
is a simple and practical choice.). To estimate the critical
parameters the field mixing parameter $s$ must be tuned along with
$\mu$ and $T$ such as to best match $P({\cal M})$ to
$P^\star(\tilde{m})$ {[29]}.
We have used these techniques to determine the critical points and
coexistence binodals for a selection of values of $q$ in the interval
$q=[0.25,0.1]$; the results are shown in Fig.~\figpd(a,d)~of the main
text. Our universal distribution matching method also delivers
estimates for the field mixing properties of the critical fluid. For
$q=0.2$, a satisfactory matching of the number density distribution to
the Ising fixed point form can be obtained as shown in Fig.~\ref{fig:critical}(a),
so field mixing is negligible in this case. By contrast, for
$q=0.126$, no satisfactory matching can be obtained under the assumption $s=0$.
The situation is greatly improved, however, by forming the distribution
of the mixed field ordering operator $P({\cal M})$, as shown in
Fig.~\ref{fig:critical}(b), which corresponds to
$s=-0.058(1)$
\section{VOID SIZES IN LIQUID STATES}
\label{sec:void}
\begin{figure}
\includegraphics[width=0.99\columnwidth]{freevol-mod-all.pdf}
\caption{(a) Fraction of space $f^{\rm void}$ available to spherical ``ghost particles'' of size $\sigma^{\rm probe}$, in liquid states
at different size ratios $q$. (The probe diameter is measured in units of the colloid diameter $\sigma$.) For $\sigma^{\rm probe}=1$,
the porous liquids at $q=0.126$ have many more cavities where ghost particles will fit, compared with the dense liquids at $q=0.2$.
(b)~Depiction of void spaces in representative configurations, for $\sigma^{\rm probe}=1$. Each connected void space is shown with a different
color, to illustrate their sizes and connectivity.}
\label{fig:void}
\end{figure}
In Fig.~\ref{fig:void}, we show the fraction $f^{\rm void}$ of the total volume of the system that is accessible to ``probe particles'' (or ``ghost'' particles) of size
$\sigma^{\rm probe}$. These calculations were performed at states where $\eta_{\rm s}=1.05\eta_{\rm s}^c$.
The porous structure of the liquid at $q=0.126$ means that
$f^{\rm void}$ is greater by several orders of magnitude than the corresponding state at $q=0.2$. We also show the void spaces
associated with typical configuration, at $\sigma^{\rm probe}=1$. Note that the sizes of the typical voids increase along with their
volume fraction; at $q=0.126$, the largest cluster has percolated the system.
\section{WERTHEIM THEORY}
\label{sec:wert}
In the main text, arguments based on Wertheim's theory of associating fluids~{[34]}~were used to investigate the effects
of branching and non-specific binding on phase behaviour.
The theory consists of two main parts. The first is an exact representation
of a fluid's partition function, where the density is decomposed as a sum of partial densities that
describe particles in different local environments. The second part is an (approximate) thermodynamic perturbation
theory (TPT) which facilitates explicit calculations.
In {[17]}, we developed this theory for a system of indented colloids that forms linear chains, but both chain branching
and back-to-back binding were neglected.
The generalisation of this theory to include these effects is straightforward, although
some book-keeping is required to keep track of the various local environments. Briefly, the interaction between particles
is broken down into three parts, which take care of particle overlaps, lock-and-key binding, and back-to-back interactions. This decomposition
is performed at the level of the Mayer $f$-function, so the diagrammatic expansion of the free energy contains three kinds of interaction.
Within this diagrammatic expansion, Wertheim's theory requires that particles be classified according to their local bonding. In practice,
we keep track of the number of inward and and outward lock-and-key bonds, and the number of back-to-back bonds. The constraint (recall Fig.~\fignet~of main text) that
each particle has at most one outward bond is enforced at the level of the diagrammatic series (this is the key advantage of Wertheim's approach).
\subsection{Quasispecies densities and mass-action equations}
To develop the theory, let $\rho_{ijk}$ be the number of particle with $i$ inward bonds, $j$ outward bonds and $k$ back-to-back bonds. These are the quasispecies
densities of Wertheim's theory~{[34]}. For simplicity we assume that particles may have at most two inward
lock-and-key bonds ($k\leq2$), at most one outward lock-and-key bond ($i\leq 1$) and at most one back-to-back bond ($k\leq 1$).
Further, we assume that
that particles with back-to-back bonds have no inward lock-and-key bonds and vice versa, so that
only one of $j$ or $k$ may be greater than zero.
Even with these constraints, there are still eight distinct local environments, but
the advantage of the TPT is that
the various $\rho_{ijk}$ have fairly simple relations between them.
We postpone the detailed analysis to a later work and quote only the main results, with their physical interpretation. We
use the notation $\rho_0=\rho_{000}$, for brevity. ``Chemical equilibrium'' between
free monomers and particles with a single inward bond (and no other bonds) means that
\begin{equation}
\rho_{010} = K \rho_{0} \rho_{\rm L}
\label{equ:rho010}
\end{equation}
where
\begin{equation}
\rho_{\rm L} = \rho_0 + \rho_{100} + \rho_{020} + \rho_{001}
\end{equation}
is the density of unoccupied lock sites (see also the main text),
and $K$ is an equilbrium constant which, at the simplest level, is equal to $B_{2,\rm eff}^{\rm LK}$. More generally, $K$
may be obtained within the TPT as an integral involving the radial distribution function $g(r)$ in a hard-particle
reference system.
We also define
\begin{align}
\rho_{\rm K} & = \rho_0 + \rho_{010} \nonumber \\
\rho_{\rm LK} & = \rho_{010} + \rho_{110} \nonumber \\
\rho_{\rm LLK} & = \rho_{020} + \rho_{120}
\end{align}
which correspond to number densities of key surfaces with $0$, $1$, and $2$ inward bonds. Also,
\begin{equation}
\rho_{\rm KK} = \tfrac12 ( \rho_{001} + \rho_{101})
\end{equation}
is the number density of back-to-back bonds.
Chemical equilibria between other particle quasispecies yield
\begin{align}
\rho_{001} &= K_{\rm BB} \rho_{0} \rho_{\rm K} \nonumber \\
\rho_{020} &= a K^2 \rho_{0} \rho_{\rm L}^2
\label{equ:rho001}
\end{align}
where $K_{\rm BB}$ is an equilibrium constant for back-to-back binding, and $a$ is a geometrical factor that accounts
for the reduced volume available when binding a second lock onto a key where a lock is already bound. (The theory
provides integral expressions for these quantities; in general we expect $a$ to depend weakly on
the energy scales $\varepsilon_{\rm LK},\varepsilon_{\rm BB}$ while $K_{\rm BB}\propto \mathrm{e}^{\beta\varepsilon_{\rm BB}}$.)
The Wertheim theory also builds in independence of bonds for the lock- and key-sites of a given
particle. This means for example that $\rho_{110} = \rho_{100}\rho_{010}/\rho_0$, and similarly
$\rho_{120} = \rho_{100}\rho_{020}/\rho_0$ and $\rho_{101} = \rho_{100}\rho_{001}/\rho_0$.
Combining all the definitions and theoretical results given so far, we arrive at
\begin{align}
\rho_{\rm LK} & = K \rho_{\rm L} \rho_{\rm K} \nonumber \\
\rho_{\rm LLK} & = a K \rho_{\rm L}^2 \rho_{\rm K} \nonumber \\
\rho_{\rm KK} & = K_{\rm BB} \rho_{\rm K}^2
\label{equ:ma}
\end{align}
which are the ``mass-action'' equations described in the main text.
Since the total density $\rho$ is equal to the sum of all the partial densities $\rho_0,\rho_{001}$ etc, we also have
\begin{align}
\rho & = \rho_{\rm K} + \rho_{\rm LK} + \rho_{\rm LLK} + 2\rho_{\rm KK}
\label{equ:key-sum}
\end{align}
which simply means that all key-sites must be either free or involved in some kind of bonding. Combining (\ref{equ:rho010},\ref{equ:rho001})
yields $\rho_{100}\rho_{\rm L} = (\rho_{010} + 2\rho_{020})\rho_{\rm K}$ from which one may derive
\begin{equation}
\rho = \rho_{\rm L} + \rho_{\rm LK} + 2\rho_{\rm LLK}
\label{equ:lock-sum}
\end{equation}
This equation enforces the constraint that all lock-sites are either occupied or unoccupied, analogous to (\ref{equ:key-sum}).
\subsection{Pressure}
Following~{[23]}, we obtain the pressure
of the system within the TPT as a sum of two terms
\begin{equation}
\beta P = \rho_{\rm tree} + \Delta P_{\rm ref}
\end{equation}
where $\rho_{\rm tree}$ is an estimate of the number of clusters in the system and $\Delta P_{\rm ref} = \beta P_{\rm ref} - \rho$ is
a contribution from the hard-particle reference system, which has a reduced pressure $\beta P_{\rm ref}$ that depends on the total
density $\rho$ but not on the quasispecies densities. As discussed in the main text, $\rho_{\rm tree}$ is equal to the number of
clusters in the system if there are no internal loops within clusters (in that case, formation of any bond decreases the total number of clusters by $1$).
This ``tree-like'' assumption motivates the notation $\rho_{\rm tree}$.
The theory described here gives
\begin{equation}
\rho_{\rm tree} = \rho_{\rm L} - \rho_{\rm KK} .
\label{equ:tree}
\end{equation}
To see this, note that if all clusters are tree-like then
their number is equal to the difference between the total number of particles and the total number of bonds (either lock-key
or back-to-back). The number of lock-key bonds is $\rho_{\rm LK} + 2\rho_{\rm LKK} = \rho-\rho_{\rm L}$ and the number of back-to-back bonds
is $\rho_{\rm KK}$, yielding (\ref{equ:tree}).
\subsubsection{Conditions for phase separation}
\label{sec:sep}
Liquid vapour phase transitions occur if $(\partial P/\partial V)_{N,T}>0$ for some volume $V$. Here the derivative is taken
at a constant value of the total particle number $N$, but the numbers of particles within each quasispecies depend on $V$.
The reference contribution $\Delta P_{\rm ref}$ is decreasing in $V$ so a necessary condition for phase separation within Wertheim's theory is
that $(\partial \rho_{\rm tree}/\partial V)_{N,T}>0$. At constant $T$, $\rho_{\rm tree}$ is a simple function of $\rho$, which implies
$V(\partial \rho_{\rm tree}/\partial V)_{N,T} = -\rho(\mathrm{d}\rho_{\rm tree}/\mathrm{d}\rho)$, so phase separation can
occur only if
\begin{equation}
(\mathrm{d}\rho_{\rm tree}/\mathrm{d}\rho) < 0.
\label{equ:sep-suff}
\end{equation}
(More precisely, one requires $\mathrm{d}\rho_{\rm tree}/\mathrm{d}\rho <
-\mathrm{d}(\Delta P_{\rm ref})/\mathrm{d}\rho$.)
The relation (\ref{equ:sep-suff}) provides a useful necessary condition for phase separation.
In particular, for the system considered here, the densities $\rho_{\rm L}$, $\rho_{\rm K}$, $\rho_{\rm KK}$, etc are
all monotonic in $\rho$ (see Sec.~\ref{sec:mono} below). In the absence of back-to-back binding,
one has $\rho_{\rm tree}=\rho_{\rm L}$ from (\ref{equ:tree}), the pressure is monotonic in $\rho$, so
the theory predicts that no phase transition is possible.
Physically, the idea is that internal loops within bonded clusters of particles
are suppressed by the directed lock-key bond and the fact that a lock-site can bind to
at most one key-site. The Wertheim theory captures this property.
\subsubsection{Monotonicity of $\rho_{\rm L}$, $\rho_{\rm K}$}
\label{sec:mono}
To complete the argument that phase transitions within the theory requires $K_{\rm BB}>0$, we prove
that $\rho_{\rm L}$ and $\rho_{\rm K}$ depend monotonically on $\rho$.
Starting from (\ref{equ:lock-sum}), we use (\ref{equ:ma}) to express all quantities in terms of $\rho_{\rm L},\rho_{\rm K}$, and
we consider small variations in $(\rho,\rho_{\rm L},\rho_{\rm K})$, obtaining
\begin{equation}
\delta \rho = \delta\rho_{\rm L} ( 1 + K\rho_{\rm K} + 4aK^2\rho_{\rm L}\rho_{\rm K}) + \delta\rho_{\rm K} ( K\rho_{\rm L} + 2 a K^2 \rho_{\rm L}^2 )
\label{equ:dr}
\end{equation}
From (\ref{equ:key-sum},\ref{equ:lock-sum}), we have
\begin{equation}
\rho_{\rm L} + \rho_{\rm LLK} = \rho_{\rm K} + 2 \rho_{\rm KK}
\label{equ:sum-both}
\end{equation}
and again considering small variations in $\rho_{\rm L},\rho_{\rm K}$ yields
\begin{equation}
\delta\rho_{\rm L} ( 1 + 2 aK^2 \rho_{\rm L}\rho_{\rm K}) = \delta \rho_{\rm K} ( 1 - a K^2 \rho_{\rm L}^2 + 4 K_{\rm BB} \rho_{\rm K} ).
\end{equation}
Using (\ref{equ:sum-both}) again gives
\begin{equation}
\delta\rho_{\rm L} ( 1 + 2 aK^2 \rho_{\rm L}\rho_{\rm K}) = \frac{\delta \rho_{\rm K}}{\rho_{\rm K}} ( K\rho_{\rm L} + 2K_{\rm BB}\rho_{\rm K} ).
\label{equ:drlk}
\end{equation}
It follows from (\ref{equ:dr},\ref{equ:drlk}) that $\delta\rho$ and $\delta\rho_{\rm K}$ are both monotonic in $\delta\rho_{\rm L}$,
which implies that $(\rho_{\rm L},\rho_{\rm K})$ are both monotonic in $\rho$, as are $(\rho_{\rm KK},\rho_{\rm LK},\rho_{\rm LKK})$. Given that
$\rho_{\rm tree} = \rho_{\rm L} - \rho_{\rm KK}$, it follows that $\rho_{\rm tree}$ must be monotonic if $\rho_{\rm KK}=0$, in
which case $(\mathrm{d}\rho_{\rm tree}/\mathrm{d}\rho) > 0$ and the theory predicts no phase transition.
\end{appendix}
\end{document}
|
\section{Introduction}
\noindent
The most important questions of celestial mechanics are the ones concerning central configurations. They have a long history dating back to the 18th century and have been studied by various mathematicians, including Euler and Lagrange.
The first known central configurations were the three collinear classes discovered in 1767 by Euler (see \cite{[EULER]}) and the two classes with masses located at the vertices of an equilateral triangle found in 1772 by Lagrange (see \cite{[LAGRANGE]}). Later it was proved by Wintner that the $3$ collinear and $2$ triangular configurations are the only possible central configurations in the $3$-body problem (see \cite{[WINTNER]}).
In general, the question about the finiteness of number of central configurations has been included in the famous list of eighteen problems of the 21st century by Smale (\cite{[SMALE]}) and it is the following:
\begin{quote}
``Given positive real numbers $m_1, \ldots, m_n$ as the masses in
the n-body problem of celestial mechanics, is the number of relative equilibria finite?''
\end{quote}
As a matter of fact, this open question was already formulated by Wintner in his book (\cite{[WINTNER]}) in 1941. In the $4$- and $5$-body problem the answer to Smale's question (\cite{[SMALE]}) has been given by Hampton and Moeckel (\cite{[HAMPTONMOECKEL]}), and by Kaloshin and Albouy (\cite{[KALOSHINALBOUY]}), respectively.
Although as for now the question remains unanswered for $6$ and more bodies, there were given some estimations for the number of central configurations. For instance, Merkel in \cite{[MERKEL]} and Pacella in \cite{[PACELLA]}, using Morse theory, determined a lower bound for the number of spatial non-collinear central configurations (for a lower bound of the number of planar non-collinear central configurations see \cite{[PALMORE]}).
For a deeper discussion of central configurations we refer the reader to \cite{[MOECKEL]}.
Classifying central configurations is a very difficult problem, therefore an answer is sought even with some simplifications, that is imposing restrictions on the masses or considering only highly symmetrical solutions.
One of the many people to go in this direction was Palmore, who considered for $n=4, 5$ the configuration of $n-1$ bodies of mass $1$ at the vertices of a regular $(n-1)$-gon and $n$-th body of arbitrary mass $m$ at its centroid (\cite{[PALMORE1973]}). He showed that for $n=4$ and for $n=5$ there exist unique values of the mass parameter $m$ for which this central configuration becomes degenerate.
Meyer and Schmidt in 1988 reproduced these results and additionally showed that a bifurcation occurs (\cite{[MEYERANDSCHMIDT8*]}).
Since then a lot more central configurations have been discovered in spirit of these mentioned restrictions.
Examples of central configurations which are considered in our paper are the following:
\begin{itemize}
\item a planar configuration of two nested squares found in $2013$ by A. C. Fernandes, L. F. Mello and M. M. da Silva in \cite{[FERNANDES]},
\item a planar rosette configuration of $13$ bodies studied in $2006$ by J. Lei and M. Santoprete in \cite{[LEIANDSANTOPRETE]}, and in $2004$ by M. Sekiguchi in \cite{[SEKIGUCHI]}.
\end{itemize}
There are numerous reasons for studying central configurations.
There exists an important class of solutions of the $n$-body problem - the so-called homographic solutions (\cite{[WINTNER]}) - which is directly linked to the concept of central configurations. A solution is
called homographic if at every time the configuration of the bodies
remains similar. As a matter of fact, this can only happen for central configurations. The Euler and Lagrange solutions, which we have already mentioned, are the most famous examples. Thus central configurations generate the unique explicit solutions of the $N$-body problem known until now.
In particular, planar central configurations also give rise to families of periodic solutions. As a special case, the homographic solutions for which the configuration just rotates rigidly at constant angular speed are known as relative equilibria.
Moreover, central configurations are important for instance in analysing total collision orbits, namely a certain change of coordinates in the neighbourhood of a total collision point turns the colliding particles into a central configuration in the limit as $t$ approaches the collision time. This phenomenon was shown initially for the complete collapse of all bodies at the centre of mass by Wintner in \cite{[WINTNER]}
and was later generalised by Saari in the study of the parabolic escapes of the particles. The same holds for expanding subsystems, which also asymptotically tend to a central configuration (readers interested in these two phenomena should see \cite{[SAARI]} for details and more references). These examples demonstrate that knowledge of central configurations gives important insight to the dynamics near a total collision and asymptotic behaviour of the universe.
Another aspect motivating the search for these special solutions is the fact that the hypersurfaces of constant energy and angular momentum change their topology exactly at the energy level sets which contain central configurations (\cite{[SMALEII]}).
Central configurations are invariant under homotheties and rotations - the first claim comes from the homogeneity of the potential while the second one stems from the fact that the potential depends only on the mutual distances between the particles and not on its positions.
Additionally, in the planar case the action of the symmetry group, in this case $SO(2),$ is free, so the quotient space $\Omega/SO(2)$ of the action of the group $SO(2)$ on the configuration space $\Omega$ is a manifold.
Because of this a natural thing to consider is a new potential defined on the quotient space $\Omega/SO(2),$ which has been widely used in the literature
(\cite{[LEIANDSANTOPRETE], [MEYER1987], [MEYERANDSCHMIDT8*], [MEYERANDSCHMIDT1988]}). In this context central configurations have been viewed not as critical orbits, but as critical points of a ``quotient'' potential.
On the other hand, in the spatial case the action of the symmetry group $SO(3)$ is not free, therefore the quotient space $\Omega/SO(3)$ of the action of the group $SO(3)$ on the configuration space $\Omega$ is not a manifold (there are two isotropy groups appearing in the configuration space), so the use of ordinary theory is not justifiable (without any differential structure). Thus in this case we cannot use the approach from the planar one.
For example, the equivariant Morse theory has been presented to the study of spatial central configurations in \cite{[PACELLA]}. In general, a different approach which does not employ the concept of quotient space has also been successfully used (\cite{[GARCIAIZE], [MACIEJEWSKIRYBICKI], [CHAVELARYBICKI]}).
Apart from the symmetries, Newton's equations are of gradient nature, therefore it seems reasonable to apply methods which benefit from this.
In this paper we use the second approach to study classes of planar central configurations, which are treated as $SO(2)$-orbits of critical points of $SO(2)$-invariant potential.
Among the machinery which is designed to work in the equivariant context making use both of the symmetry and the gradient structure appearing in the problem and which we employ in our paper are equivariant versions of classical topological invariants such as the degree for equivariant gradient maps (see \cite{[GEBA]} and \cite{[RYBICKI]}) and the equivariant Conley index (see \cite{[BARTSCH]}, \cite{[FLOER]} and \cite{[GEBA]}).
This approach using equivariant theory has for instance been used in \cite{[MACIEJEWSKIRYBICKI]} and \cite{[CHAVELARYBICKI]}.
For general treatment of the degree for equivariant maps we refer the reader to \cite{[BALANOVKRAWCEWICZSTEINLEIN]} and \cite{[BALANOVKRAWCEWICZRYBICKISTEINLEIN]}.
In this article we apply those topological tools in equivariant bifurcation theory to produce new classes of planar central configurations from known families by obtaining global bifurcations of families of planar central configurations under a change of the degree for equivariant gradient maps. The same result has been reached for local bifurcations under a change of the equivariant Conley Index (\cite{[SMOLLER]}).
In the planar case because we have exactly one orbit type on the configuration space, the results obtained by using the equivariant Conley Index or the degree for equivariant gradient maps are equivalent to those which one could get considering the quotient space and applying ordinary Conley Index or the Brouwer degree. However, in the spatial case the approach via the quotient space is not possible. In general, the obtained theoretical results are stronger than these we need to study of the planar $N$-body problem.
The aim of this paper is twofold. On the one hand, we give new results about equivariant bifurcation theory (see Theorems \ref{twierbif1}, \ref{twchangedegree} and \ref{twierbif2}).
On the other hand, we apply those theoretical results to the study of planar central configurations.
In the next article we shall present an application of our abstract results to the spatial $N$-body problem and give necessary and sufficient conditions for the existence of local and global bifurcations of spatial central configurations. Thus the main goal of this article is to create a homogeneous theory which can be used for both planar and spatial $N$-body problems. In the planar case, as we mentioned before, we can consider the orbit space of the action of the group $SO(2)$ on the configuration space and then use both ordinary Morse theory or the Brouwer degree and the equivariant approach. However, in the spatial case we cannot work on the quotient space because the action of the group $SO(3)$ on the configuration space is not free.
Notice that we can consider the possibility of working on subsets of the full configuration space which are invariant for the gradient flow and seek bifurcations in these subsets, but it is possible that we can get no bifurcations while there are ones in the full configuration space. This phenomenon happened for instance in \cite{[MEYERANDSCHMIDT8*]}, that is bifurcations of less symmetrical families from highly symmetrical families of central configurations occur. Meyer and Schmidt treated the highly symmetrical family of equilateral triangle with fourth body at the centroid and proved that another families which are less symmetrical (that is the families of isosceles triangles with fourth body near the centroid and on the line of symmetry of the triangle) bifurcate from this family. Similarly, these authors taking the family of square with fifth body at the centroid proved bifurcations of families of kites and isosceles trapezoids (for more examples of this phenomenon see \cite{[MEYERANDSCHMIDT1988]}). In our case of the family of rosette
configuration of $13$ bodies, likewise, we have the existence of bifurcations of central configurations (not necessarily configurations of rosette type) from the rosette family (see Theorem \ref{thmrosette}) and it transpires that central configurations which bifurcate are not of rosette type. However, we do not know the shape of the bifurcating families. One can compute that there are no values of the mass parameter for which the rosette family is degenerate. Therefore the obtained results are stronger than these which could be get by taking into account the invariant subsets of the configuration space. In the case of the family of two nested squares we also get stronger results not considering invariant subsets.
We stress the fact that the algebraic structure of the potential is not important from the point of view of central configurations, the only things that matter are the rotation and scaling invariance. This makes the methods that we use applicable to mathematical models of physical problems involving interactions between multiple bodies, whose behaviour is not necessarily governed by Newton's gravity laws. Such situation arises for instance in molecular dynamics, where intermolecular
relations are modelled using a broad spectrum of potentials. The famous Lennard-Jones potential, which approximately describes the behaviour of two neutral atoms or molecules, was studied in this context in \cite{[CORBERALLIBRE1]} and \cite{[CORBERALLIBRE2]}.
This paper is organised as follows: Section \ref{preliminaries} contains an introduction to the equivariant setting along with the definition and properties of a topological degree suitable in this context and lemmas used in the proofs of our theorems. In Section \ref{bifurcations} we investigate the $G$-orbits of solutions of the equation
\begin{equation}\label{eq0}
\nabla _v \varphi (v,\rho )=0
\end{equation}
under the action of $G,$ strictly speaking we formulate abstract theorems giving necessary and sufficient conditions for the existence of bifurcations in the vicinity of a given family of solutions.
In Theorem \ref{wkkoniecznybif} we state that only a degenerate critical $G$-orbit of solutions of equation $(\ref{eq0})$ can be a $G$-orbit of local bifurcation.
The sufficient condition is given by Theorem \ref{twierbif1}, its assumptions, which include a change of the degree for $G$-equivariant gradient maps computed at critical $G$-orbits of $\varphi$, imply the existence of a global bifurcation. We also provide simple conditions in Theorem \ref{twchangedegree} and Remark \ref{twbif1remark1}, which allow us to verify inequality of degrees. Moreover, we prove a theorem in the spirit of Rabinowitz's famous alternative, see Theorem \ref{twierbif2}. In Section \ref{applications} we apply our abstract results to the planar $N$-body problem, namely we consider known families of central configurations - the two squares family \eqref{family_two_squares} and the rosette family \eqref{family_rosette} - and prove the existence of bifurcations of new classes of central configurations from these families. We stress that our method is applicable not only in this specific example, but is rather a general framework, which is additionally
easily implementable on any computer algebra system, which allows symbolic computations.
Notice that in this paper we consider topological bifurcations (local and global ones, see Definitions \ref{biflok} and \ref{bifglob}) rather than bifurcations in the sense of a change in the number of solutions. In other words, in the case of local one, we consider known family of solutions of equation \eqref{eq0} (so-called trivial solutions) and seek non-trivial ones nearby.
In the case of global one, we seek connected sets of non-trivial solutions nearby the trivial ones which additionally satisfy Rabinowitz type alternative (see condition \eqref{Rabinowitzalternative}).
Those two definitions, the bifurcation in the sense of a change in the number of solutions and the topological bifurcation, are independent, that is the first one can occur while the second one does not occur and inversely. Throughout this paper we will write about topological bifurcations briefly bifurcations.
\setcounter{Theorem}{0}\setcounter{Definition}{0
\section{Preliminaries}
\label{preliminaries}
\setcounter{Theorem}{0}\setcounter{Definition}{0
In this section we review some classical
facts on equivariant topology. The material is fairly standard and well known (for details see for example \cite{[DIECK2]} and \cite{[KAWAKUBO]}).
Throughout this paper we will remain in the real setting, that is all vector spaces will be considered over $\mathbb{R}$ and all matrices (treated as elements of Lie groups) will have real entries.
Also, from now on $G$ stands for a compact Lie group and $\mathbb{V}=(\mathbb{R}^n,\varsigma)$ is a finite-dimensional, real, orthogonal $G$-representation, that is a pair consisting of the Euclidean $n$-dimensional vector space and a continuous homomorphism
$\varsigma: G\rightarrow O(n),$
where $O(n)$ denotes the group of orthogonal matrices. By $v\in\mathbb{V}$ we mean $v\in\mathbb{R}^n.$
In addition to that by $SO(n)$ we will understand the group of special orthogonal matrices.
The linear action of $G$ on $\mathbb{R}^n$ is given by $\xi :G\times\mathbb{R}^n\rightarrow\mathbb{R}^n,$ where
$\xi(g,v)=\varsigma(g)v.$
We write $gv$ for short.
Moreover, $\mathbb{R}$ denotes the one-dimensional, trivial representation of the group $G,$ that is $\mathbb{R}=(\mathbb{R},\mathbf{1}),$ where $\mathbf{1}(g)=1$ for any $g\in G.$
Then the linear action of $G$ on $\mathbb{R}^n\times\mathbb{R}$ is given by the Cartesian product of $\varsigma$ and $\mathbf{1}$
\[G\times(\mathbb{R}^n\times\mathbb{R})\ni (g,(v,\rho))\mapsto (gv,\rho)\in \mathbb{R}^n\times\mathbb{R}.\]
A subset $\Omega\subset\mathbb{V}$ is called $G$-invariant if for any $g\in G$ and $v\in\Omega$ we have $gv\in\Omega.$
Let $\overline{sub}(G)$ be the set of closed subgroups of $G.$
Two subgroups $H, K \in \overline{sub}(G)$ are called \emph{conjugate} if there exists $g \in G$ such that $H=g^{-1}Kg.$ Conjugacy is an equivalence relation, the conjugacy class of $H$ is denoted by $(H)$ and it is called the orbit type of $H.$ Let $\overline{sub}[G]$ be the set of conjugacy classes of closed subgroups of $G.$ Similarly, $H$ is \emph{subconjugate} to $K$ if $H$ is conjugate to a subgroup of $K,$ denote by $(H)\leq (K).$ Subconjugacy defines a partial order in $\overline{sub}[G].$ Additionally, we write $(H)<(K)$ if $(H)\leq (K)$ and $(H)\neq (K).$
If $v\in\mathbb{V}$ then
$G_v=\{g\in G:gv=v\}\in\overline{sub}(G)$
is the isotropy group of $v.$
For each $v\in\mathbb{V}$ the set
$G(v)=\{gv:g\in G\}\subset\mathbb{V}$
is called the $G$-orbit through $v.$
The isotropy groups of points on the same $G$-orbit are conjugate subgroups of $G.$ We will also denote by $\mathbb{V}^G$ the set of fixed points of the
$G$-action on $\mathbb{V},$ that is
$\mathbb{V}^G=\{v\in\mathbb{V}: G_v=G\}=\{v\in\mathbb{V}:gv=v\ \forall g\in G\}.$
Moreover, for given $H\in\overline{sub}(G)$ and an open, $G$-invariant subset $\Omega\subset\mathbb{V}$ we have
$\Omega_{<(H)}=\{v\in\Omega: (G_v)<(H)\}$ and $\Omega_{(H)}=\{v\in\Omega: (G_v)=(H)\}.$
Let $H\in\overline{sub}(G)$ and $\mathbb{W}$ be an orthogonal $H$-representation. Now define an action of $H$ on the product $G\times\mathbb{W}$ by the formula
$(h,(g,w))\mapsto (gh^{-1},hw)$
and let $G\times_H\mathbb{W}$ denote the space of $H$-orbits of this action.
Fix $k,l\in \mathbb{N}\cup\{\infty\}$ and an open, $G$-invariant subset $\Omega\subset\mathbb{V}.$ A map $\varphi:\Omega\times\mathbb{R}\rightarrow\mathbb{R}$ of class $C^{k}$ is said to be a \emph{$G$-invariant $C^k$-map} if
$\varphi(gv,\rho)=\varphi(v,\rho)$
for any $g\in G$ and $(v,\rho)\in\Omega\times\mathbb{R}.$
The set of $G$-invariant $C^{k}$-maps is denoted by $C^k_G(\Omega\times\mathbb{R},\mathbb{R}).$
A map $\psi:\Omega\times\mathbb{R}\rightarrow\mathbb{V}$ of class $C^{l}$ is said to be a \emph{$G$-equivariant $C^l$-map} if
$\psi(gv,\rho)=g\psi(v,\rho)$
for any $g\in G$ and $(v,\rho)\in\Omega\times\mathbb{R}.$
The set of $G$-equivariant $C^{l}$-maps is denoted by $C^l_G(\Omega\times\mathbb{R},\mathbb{V}).$
For any $G$-invariant $C^k$-map $\varphi$ the gradient of $\varphi$ with respect to the
first coordinate denoted by $\nabla_v\varphi$ is $G$-equivariant $C^{k-1}$-map.
Let $\Omega\subset\mathbb{V}$ be open and $G$-invariant and fix $v_0\in\Omega,$ where by $H$ we denote $G_{v_0}.$
Notice that $\Omega_{(H)}$ is a $G$-invariant submanifold of $\Omega$ and $G(v_0)$ is a $G$-invariant submanifold of $\Omega_{(H)},$ therefore
we obtain the following orthogonal direct sum decomposition
\begin{equation}\label{decomp1}
T_{v_0}\mathbb{V}=T_{v_0}\Omega=(T_{v_0}\Omega_{(H)})\oplus(T_{v_0}\Omega_{(H)})^{\bot}=
(T_{v_0}G(v_0))\oplus(T_{v_0}\Omega_{(H)}\ominus T_{v_0}G(v_0))\oplus(T_{v_0}\Omega_{(H)})^{\bot},
\end{equation}
where $T_{v_0}\mathbb{V}$ is the tangent space to $\mathbb{V}$ at $v_0.$
On the other hand, we have
\begin{equation}\label{decomp2}
T_{v_0}\mathbb{V}=T_{v_0}\Omega=(T_{v_0}G(v_0))\oplus\mathbb{W}^H\oplus( \mathbb{W}^H)^{\bot},
\end{equation}
where $\mathbb{W}=(T_{v_0}G(v_0))^{\bot}$ is an orthogonal $H$-representation.
Because $(G\times_HB^{\varepsilon}_{v_0}(\mathbb{W}))_{(H)}=G/H\times B^{\varepsilon}_{v_0}(\mathbb{W}^H),$ we get that
$\mathbb{W}^H\subset T_{v_0}\Omega_{(H)},$ where $B^{\varepsilon}_{v_0}(\mathbb{W})$ denotes the $\varepsilon$-ball centred at $v_0$ in $\mathbb{W}.$
Now assume that $\varphi \in C^2_G(\Omega , \mathbb{R})$ and $\ v_0\in (\nabla \varphi)^{-1} (0).$
The following lemma gives the decomposition of the Hessian $\nabla^2\varphi$ of the potential $\varphi$ with respect to the sum decompositions given by the formulas \eqref{decomp1} or \eqref{decomp2} and has been proved in \cite{[GEBA]}.
\begin{Lemma} \label{hesjanpostac}
Under the above assumptions, the Hessian
\begin{equation*} \left.\begin{array}{lccccc}
& & T_{v_0}G(v_0) & & & T_{v_0}G(v_0) \\
& &\oplus & & & \oplus \\
\nabla ^2 \varphi (v_0) \ : & T_{v_0}\mathbb{V}= & \mathbb{W}^H & \rightarrow & T_{v_0}\mathbb{V}= & \mathbb{W}^H \\
& & \oplus & & & \oplus \\
& & ( \mathbb{W}^H)^{\bot} & & & (\mathbb{W}^H)^{\bot}
\end{array}\right. \end{equation*} is of the form
\begin{equation}\label{hesjanpostacwzor} \nabla ^2 \varphi (v_0)=
\left[ \begin{array}{lcc} 0 & 0 & 0 \\ 0 & B(v_0) & 0 \\ 0 & 0 &
C(v_0)
\end{array}\right] .
\end{equation}
\end{Lemma}
According to Lemma \ref{hesjanpostac}, we have $\dim \ker \nabla^2\varphi(v_0)\geq\dim G(v_0).$
We will thus call a critical $G$-orbit $G(v_0)$ \emph{degenerate} if the strict inequality holds and \emph{non-degenerate} if\break
$\dim \ker \nabla^2\varphi(v_0)=\dim G(v_0)$ (which is equivalent to $B(v_0)$ and $C(v_0)$ being non-singular). Note that non-degenerate critical $G$-orbits are isolated, that is there exists a $G$-invariant neighbourhood $\Theta\subset\mathbb{V}$ of $G$-orbit $G(v_0)$ such that $(\nabla\varphi)^{-1}(0)\cap \Theta=G(v_0).$ Moreover, a non-degenerate critical $G$-orbit $G(v_0)$ is called \emph{special} if $m^-(C(v_0))=0,$ where by the symbol $m^-(C(v_0))$ we denote the Morse index, that is the number of negative eigenvalues (counting multiplicities) of the matrix $C(v_0).$
Regarding planar central configurations, there is exactly one orbit type $(H)$ in the configuration space $\Omega$, that is $\Omega=\Omega_{(H)}$ ($(H)=(\{e\})$), so to illustrate this situation we give the following remark and example:
\begin{Remark}\label{hesjanpostaccccase}
Under the assumptions of Lemma \ref{hesjanpostac} with $\Omega=\Omega_{(H)},$ the Hessian
\begin{equation*} \left.\begin{array}{lccc}
& T_{v_0}G(v_0) & & T_{v_0}G(v_0) \\
\nabla ^2 \varphi (v_0) \ : & \oplus & \rightarrow & \oplus \\
& T_{v_0}\Omega_{(H)}\ominus T_{v_0}G(v_0) & & T_{v_0}\Omega_{(H)}\ominus T_{v_0}G(v_0)
\end{array}\right. \end{equation*} is of the form
\begin{equation}\label{hesjanpostacwzorcccase} \nabla ^2 \varphi (v_0)=
\left[ \begin{array}{lcc} 0 & 0 \\ 0 & B(v_0)
\end{array}\right] ,
\end{equation}
where $B(v_0)$ is from the formula \eqref{hesjanpostacwzor}.
\end{Remark}
\begin{Example}\label{hesjanpostacSO(2)}
Let $G=SO(2)$ and $\mathbb{V}=(\mathbb{R}^2,\varsigma),$ where $\varsigma: SO(2)\rightarrow O(2)$ is given by
$\varsigma(g)=g.$
Assume that $\Omega=\mathbb{V}\backslash\{0\},$ then for any $v\in\Omega$ we have $SO(2)_v=\{e\},$ so $\Omega=\Omega_{(\{e\})}.$
Fix $\varphi \in C^2_{SO(2)}(\Omega , \mathbb{R})$ given by $\varphi(v)=\psi(|v|^2),$ where
$\psi \in C^{\infty}(\mathbb{R} ,\mathbb{R})$ is defined by $\psi(t)=t(t-1)$ and by the symbol $|v|$ we denote the standard norm of an element $v.$
Then $\nabla \varphi(v_0)=0$ if and only if $|v_0|^2=\frac{1}{2},$ so we get that for $v_0=(0,\frac{1}{\sqrt{2}})$ the Hessian
\begin{equation*} \left.\begin{array}{lccc}
& T_{v_0}SO(2)(v_0) & & T_{v_0}SO(2)(v_0) \\
\nabla ^2 \varphi (v_0) \ : & \oplus &\rightarrow & \oplus \\
& T_{v_0}\Omega_{(\{e\})}\ominus T_{v_0}SO(2)(v_0) & & T_{v_0}\Omega_{(\{e\})}\ominus T_{v_0}SO(2)(v_0) \\
\end{array}\right. \end{equation*} is of the form
\begin{equation*}\nabla ^2 \varphi (v_0)=
\left[ \begin{array}{lcc}
0 & 0 \\ 0 & 4
\end{array}\right] .
\end{equation*}
\end{Example}
Note that if there is only one orbit type in $\Omega$ then the quotient space $\Omega/G$ of the action of the group $G$ on the open and $G$-invariant set $\Omega$ is a manifold, so for any $G$-invariant $C^2$-map we can consider a "quotient" potential defined on $\Omega/G.$
In the following lemma we present the decomposition of the Hessian of the "quotient" potential and it transpires that the ordinary Morse index of the "quotient" potential is equal to the Morse index of matrix $B(v_0)$ (see the formula \eqref{hesjanpostacwzor}).
\begin{Lemma}
Let $\Omega\subset\mathbb{V}$ be open and $G$-invariant. Assume that $\Omega=\Omega_{(H)}$ for some $(H)\in\overline{sub}[G]$ and fix $\varphi \in C^2_G(\Omega , \mathbb{R}).$ Then the potential $\psi:\Omega/G\rightarrow\mathbb{R}$ given by $\psi(G(v_0))=\varphi(v_0)$ is a map of class $C^2$ and $\nabla ^2 \psi (G(v_0)) \ : T_{v_0}\Omega/G \rightarrow T_{v_0}\Omega/G$
is of the form $\nabla ^2 \psi (G(v_0))=\left[ B(v_0) \right],$ where $B(v_0)$ is from the formula \eqref{hesjanpostacwzor}.
\end{Lemma}
Let $\mathcal{F}_\star(G)$ denote the class of finite, pointed $G$-$CW$-complexes (for the definition of $G$-$CW$-complex see \cite{[DIECK2]}). The $G$-homotopy type of $X\in\mathcal{F}_\star(G)$ is denoted by $[X].$ Let $F$ be a free abelian group generated by $G$-homotopy types of finite, pointed $G$-$CW$-complexes and let $N$ be a subgroup of $F$ generated by elements $[A]-[X]+[X/A],$ where $A,X\in \mathcal{F}_\star(G)$ and $A\subset X.$
Put $U(G)=F/N$ and let $\chi_G(X)$ be the class of an element $[X]\in F$ in $U(G).$
If $X$ is a $G$-$CW$-complex without base point we put $\chi_G(X)=\chi_G(X^+),$ where $X^+=X\cup\{\star\}.$
The ring $U(G)$ is called the Euler ring of $G$ ( for the definition of $U(G)$ see \cite{[DIECK1]} and \cite{[DIECK2]}).
The following theorem can be found in \cite{[DIECK2]}.
\begin{Theorem}
The group $(U(G),+)$ is the free abelian group with basis $\chi_G (G/H^+)$ for $(H)\in\overline{sub}[G].$
Moreover, if $X\in\mathcal{F}_\star(G)$ and $\bigcup _{k=0}^{p}\{(k,(H_{j,k})):\ j=1,\cdots,q(k)\}$ is a type of the cell decomposition of $X,$
then
\[\chi_G (X)=\sum_{(H)\in\overline{sub}[G]}\left(\sum_{k=0}^p(-1)^{k}\nu(k,(H))\right)\cdot \chi_G (G/H^+),\]
where $\nu(k,(H))$ is the number of cells of dimension $k$ and of orbit type $(H).$
\end{Theorem}
Let $\Omega\subset\mathbb{V}$ be open and $G$-invariant. A function $\varphi\in C^1_G(\mathbb{V},\mathbb{R})$ is said to be \emph{$\Omega$-admissible} if $(\nabla\varphi)^{-1}(0)\cap\partial\Omega=\emptyset.$
An $\Omega$-admissible map $\varphi\in C^2_G(\mathbb{V},\mathbb{R})$ is called a \emph{special Morse function} if for any $v_0\in(\nabla \varphi)^{-1} (0)\cap\Omega$ a critical $G$-orbit $G(v_0)$ is special.
Let $\Omega\subset\mathbb{V}$ be open, $G$-invariant and bounded. For an $\Omega$-admissible function $\varphi\in C^1_G(\mathbb{V},\mathbb{R})$ there is a notion of degree for equivariant gradient maps $\nabla_G\text{-}\mathrm{deg}(\nabla\varphi,\Omega)\in U(G).$ This invariant has been introduced by G\k{e}ba in \cite{[GEBA]} with coordinates as follows
\[\nabla_G\text{-}\mathrm{deg}(\nabla\varphi,\Omega)=\sum_{(H)\in\overline{sub}[G]}\nabla_G\text{-}\mathrm{deg}_{(H)}(\nabla\varphi,\Omega)\cdot \chi_G (G/H^+)\in U(G).\]
To make this article more readable we give a sketch of the definition of the degree. First, it is proved that there exists a special Morse function $\psi\in C^2_G(\mathbb{V},\mathbb{R})$ which is $G$-homotopic to the map $\varphi,$ that is there exists $H \in C^1_G(\mathbb{V}\times[0,1],\mathbb{R})$ being $\Omega$-admissible for each $t\in[0,1]$ such that $H(\cdot,0)=\psi$ and $H(\cdot,1)=\varphi.$ Additionally, because non-degenerate critical $G$-orbits are isolated and $\Omega$ is bounded, we get that $(\nabla\psi)^{-1}(0)\cap\Omega$ consists of a finite number of distinct critical $G$-orbits, that is $(\nabla\psi)^{-1}(0)\cap\Omega=G(v_1)\cup \ldots \cup G(v_l),$ where $G(v_i)\cap G(v_j)=\emptyset$ for $i\neq j.$ Let us remind that for any special critical $G$-orbit $G(v_i)$ we have $m^-(C(v_i))=0,$ so next we allocate $(-1)^{m^-(B(v_i))}$ to each critical $G$-orbit $G(v_i)$ and define the degree for equivariant gradient maps $\nabla_G\text{-}\mathrm{deg}(\nabla\varphi,\Omega)\in U(G)$ by
\[\nabla_G\text{-}\mathrm{deg}(\nabla\varphi,\Omega)=\nabla_G\text{-}\mathrm{deg}(\nabla\psi,\Omega)=\!\!\!\!\!\!\!\!\sum_{(H)\in\overline{sub}[G]}\left(\sum_{(G_{v_i})=(H)}(-1)^{m^-(B(v_i))}\right)\cdot \chi_G (G/H^+)\in U(G).\]
The definition of $\nabla_G\text{-}\mathrm{deg}(\nabla\varphi,\Omega)$ does not depend on the choice of the special Morse function.
For details of the definition and properties of the degree for equivariant gradient maps see \cite{[GEBA]} and \cite{[RYBICKI]}.
In the following theorem we formulate some basic properties of this degree.
\begin{Theorem}\label{properties_of_degree}
Fix an $\Omega$-\textit{admissible} function $\varphi\in C^1_G(\mathbb{V},\mathbb{R}),$ where $\Omega$ is an open, bounded and $G$-invariant set. Then the degree $\nabla_G\text{-}\mathrm{deg}(\nabla\varphi,\Omega)\in U(G)$ has the following properties:
\begin{enumerate}
\item[(1)] \text{[Existence]} if $\nabla_G\text{-}\mathrm{deg} (\nabla\varphi, \Omega)\neq\mathbf{0}\in U(G)$, then $(\nabla\varphi)^{-1}(0)\cap\Omega\neq\emptyset$,
\item[(2)] \text{[Additivity]} if $\Omega_1,\ \Omega_2$ are open, bounded, $G$-invariant and disjoint sets such that $\Omega=\Omega_1\cup\Omega_2,$ then
\[\nabla_G\text{-}\mathrm{deg} (\nabla\varphi, \Omega)=\nabla_G\text{-}\mathrm{deg} (\nabla\varphi, \Omega_1)+\nabla_G\text{-}\mathrm{deg} (\nabla\varphi, \Omega_2),\]
\item[(3)] \text{[Homotopy invariance]} if $\Phi \in C^1_G(\mathbb{V}\times[0,1],\mathbb{R})$ is $\Omega$-admissible for each $t\in[0,1],$ then
\[\nabla_G\text{-}\mathrm{deg}(\nabla_v\Phi(\cdot,0), \Omega)=\nabla_G\text{-}\mathrm{deg}(\nabla_v\Phi(\cdot,1), \Omega),\]
\item[(4)] \text{[Excision]} if $\Omega_1\subset\Omega$ is an open, $G$-invariant set such that $(\nabla\varphi)^{-1}(0)\cap\Omega\subset\Omega_1,$ then
$\nabla_G\text{-}\mathrm{deg} (\nabla\varphi, \Omega)=\nabla_G\text{-}\mathrm{deg} (\nabla\varphi, \Omega_1).$
\end{enumerate}
\end{Theorem}
In the non-equivariant case there are some invariants for continuous maps, for instance the Brouwer degree. This invariant is sometimes too weak to distinguish homotopy classes of two equivariant gradient maps. On the other hand, they can be distinguished by the degree for equivariant gradient maps. In the example below we define two equivariant gradient maps whose the Brouwer degrees are the same while they are distinguished by the degree for $SO(2)$-equivariant gradient maps.
\begin{Example}
Let $G=SO(2)$ and $\mathbb{V}=(\mathbb{R}^2,\varsigma),$ where $\varsigma: SO(2)\rightarrow O(2)$ is given by
$\varsigma(g)=g.$
Notice that for any $v\in\mathbb{V}$ we have
\begin{equation*}
SO(2)_v=\left\{
\begin{array}{ll}
\{e\} , & v\neq 0\\
SO(2), & v= 0
\end{array}\right.
\end{equation*}
and $B^{1}_0(\mathbb{V})\subset\mathbb{V}$ is an $SO(2)$-invariant set.
Consider two maps $\varphi_\pm\in C^2_{SO(2)}(B^{1}_0(\mathbb{V}),\mathbb{R})$ given by the formulas $\varphi_\pm(v)=\pm\frac{1}{2}|v|^2.$
According to Lemma \ref{hesjanpostac}, for $0\in(\nabla\varphi_\pm)^{-1}(0)$ we have
\begin{equation*}
\nabla ^2 \varphi_+ (0)=
\left[\begin{array}{cc}C_+(0)\end{array}\right]=
\left[\begin{array}{cc}1&0\\ 0&1\end{array}\right]
\text{and}\
\nabla ^2 \varphi_- (0)=
\left[\begin{array}{cc}C_-(0)\end{array}\right]=
\left[\begin{array}{rr}-1&0\\ 0&-1\end{array}\right] .
\end{equation*}
Then the Brouwer degrees of the maps $\nabla\varphi_\pm$ at $0$ in the set $B^{1}_0(\mathbb{V})$ are equal and
\[\mathrm{deg_B}(\nabla\varphi_+,B^{1}_0(\mathbb{V}),0)=\mathrm{deg_B}(\nabla\varphi_-,B^{1}_0(\mathbb{V}),0)=+1\in\mathbb{Z}.\]
Because $\varphi_+$ is a special Morse function and it has only one critical $SO(2)$-orbit $SO(2)(0),$ we get
\begin{multline*}
\nabla_G\text{-}\mathrm{deg} (\nabla\varphi_+,B^{1}_0(\mathbb{V}) )=(-1)^{m^-(B_+(0))}\chi_{SO(2)} (SO(2)/SO(2)^+)=\\
=(-1)^0\chi_{SO(2)} (SO(2)/SO(2)^+)=\mathbf{1}\in U(SO(2)).
\end{multline*}
Next, let $H\in C^2_{SO(2)}(B^{1}_0(\mathbb{V})\times [0,1],\mathbb{R})$ be given by
$H(v,t)=(1-t\gamma(|v|))\varphi_-(v)+t\gamma(|v|)\varphi_+(v),$ where $\gamma\in C^{\infty}(\mathbb{R},\mathbb{R})$ is such that $\gamma_{|(-\infty,\frac{1}{2}]}\equiv 1,\ \gamma_{|[1,+\infty)}\equiv 0$ and $\gamma_{|(\frac{1}{2},1)}$ is a smooth, decreasing map.
Notice that $H(\cdot,0)=\varphi_-,\ H(\cdot,1)=(1-\gamma(|\cdot|))\varphi_-(\cdot)+\gamma(|\cdot|)\varphi_+(\cdot)$ and $H$ is $B^{1}_0(\mathbb{V})$-admissible for each $t\in[0,1],$ so
the homotopy invariance property of the degree (Theorem \ref{properties_of_degree}.(3)) implies that
$\nabla_G\text{-}\mathrm{deg}(\nabla\varphi_- ,B^{1}_0(\mathbb{V}))=\nabla_G\text{-}\mathrm{deg}(\nabla_vH(\cdot,1) ,B^{1}_0(\mathbb{V})).$
One can compute that the map $H(\cdot,1)$ is a special Morse function and it has exactly two critical $SO(2)$-orbit $SO(2)(0)$ and $SO(2)(v_0),$ where $v_0=(0,y_0)$ and $\frac{1}{2}<y_0<1.$
According to Lemma \ref{hesjanpostac}, for $0\in(\nabla_vH(\cdot,1))^{-1}(0)$ and $v_0\in(\nabla_vH(\cdot,1))^{-1}(0)$ we have
\begin{equation*}
\nabla^2_vH(0,1)=\nabla ^2 \varphi_+ (0)\ \text{and}\
\nabla ^2_vH(v_0,1)=
\left[\begin{array}{cc}0 & 0\\ 0 & B(v_0)\end{array}\right] .
\end{equation*}
Then
\begin{multline*}
\nabla_G\text{-}\mathrm{deg} (\nabla_vH(\cdot,1),B^{1}_0(\mathbb{V}) )=\\ =(-1)^{m^-(B(v_0))}\chi_{SO(2)} (SO(2)/\{e\}^+)+ (-1)^{m^-(B_+(0))}\chi_{SO(2)} (SO(2)/SO(2)^+) =\\=
\pm\chi_{SO(2)} (SO(2)/\{e\}^+)+\chi_{SO(2)} (SO(2)/SO(2)^+)\in U(SO(2)).
\end{multline*}
\end{Example}
Now we consider equivariant gradient maps whose the Brouwer degree vanishes and the degree for equivariant gradient maps can be nontrivial, see the following example:
\begin{Example}
Let $G=SO(2)$ and $\Omega\subset\mathbb{V}$ be an open, $SO(2)$-invariant and bounded subset of an orthogonal $SO(2)$-representation $\mathbb{V}$ such that $\Omega^{SO(2)}=\emptyset.$ Then for any $\Omega$-admissible map $\varphi\in C^2_G(\Omega,\mathbb{R})$ we have $\mathrm{deg_B}(\nabla\varphi,\Omega,0)=0\in\mathbb{Z}.$
\end{Example}
Below we formulate lemmas which are standard in equivariant topology and will be useful for us in further investigations.
\begin{Lemma}\label{fkoustalonymgihomeo}
Fix an element $g\in G.$ Then the map $\gamma_g \ : \mathbb{V}\rightarrow \mathbb{V}$ given by the formula $\gamma_g(v)=gv$ is a homeomorphism. In particular, the set $GU=\bigcup_{g\in G}\gamma_g(U)\subset \mathbb{V}$ is open and $G$-invariant for any open subset $U\subset \mathbb{V}.$
\end{Lemma}
\begin{Lemma}\label{lmozerach}
Let $\varphi \in C^2_G(\mathbb{V} , \mathbb{R})$ and assume that $U\subset \mathbb{V}.$ Then $(\nabla \varphi)^{-1} (0)\cap GU=\emptyset$ if $(\nabla \varphi)^{-1} (0)\cap U=\emptyset.$
\end{Lemma}
The following theorem can be found in \cite{[BROWN]}.
\begin{Theorem}\label{separationlemma}
Let $K$ be a compact space and $A,\ B\subset K$ be closed, disjoint sets such that there is no connected set $S\subset K$ such that $S\cap A\neq\emptyset$ and $S\cap B\neq\emptyset.$ Then there exist compact sets $K_A,\ K_B \subset K$ such that $A\subset K_A,\ B\subset K_B,\ K_A\cap K_B=\emptyset$ and $K_A\cup K_B=K.$
\end{Theorem}
\begin{Lemma} \label{wystepujacetypyorbitowe}
Fix a map $\varphi\in C^{2}_{G}(\mathbb{V},\mathbb{R})$ and $v_0\in\mathbb{V}$ such that $\nabla\varphi(v_0)=0.$
Then for any $v\in G B^{\varepsilon}_{v_0}(\mathbb{W})$ we have $(G_v)\leq (G_{v_0}),$
where $\mathbb{W}=(T_{v_0}G(v_0))^{\bot}.$
\end{Lemma}
The following lemma can be found in \cite{[MAYER]}.
\begin{Lemma}\label{istnieniespecfunkcjiMorse'a}
Fix a map $\varphi\!\in\! C^{2}_{G}(\mathbb{V},\mathbb{R})$ and assume that $G(v_0)$ is a non-degenerate critical $G$-orbit of $\varphi$ such that
$m^-(C(v_0))\neq 0.$ Then for all open, $G$-invariant neighbourhoods $\Theta \subset\mathbb{V}$ of the
orbit $G(v_0)$ such that $(\nabla\varphi)^{-1}(0)\cap\Theta=G(v_0)$ there exist an open, $G$-invariant neighbourhood $U\subset cl(U)\subset \Theta$ of the orbit $G(v_0),\ \varepsilon >0$ and a map
$\phi\!\in\! C^{2}_{G}(\mathbb{V},\mathbb{R})$ such that
\begin{enumerate}
\item $\varphi (v)=\phi(v)$ for all $v\in \mathbb{V}\backslash U(\varepsilon),$ where $U(\varepsilon)$ denotes an $\varepsilon$-neighbourhood of the set $U,$ that is $\bigcup_{v\in U}B^{\varepsilon}_{v}(\mathbb{V}),$
\item $G(v_0)$ is a special critical $G$-orbit of $\phi,$
\item $((\nabla \phi )^{-1}(0)\cap (U(\varepsilon)\backslash G(v_0)))_{<(H)}=(\nabla \phi )^{-1}(0)\cap (U(\varepsilon)\backslash G(v_0)),$ where $(H)=(G_{v_0}).$
\end{enumerate}
\end{Lemma}
\begin{Lemma} \label{postacstopnia}
Fix a map $\varphi\!\in\! C^{2}_{G}(\mathbb{V},\mathbb{R})$ and assume that $G(v_0)$ is a non-degenerate critical $G$-orbit of $\varphi$ such that $m^-(C(v_0))\neq 0$ and $(H)=(G_{v_0}).$ Then there exists an open, $G$-invariant neighbourhood $\Theta$ of the orbit $G(v_0)$ such that $(\nabla\varphi)^{-1}(0)\cap\Theta= G(v_0).$ Moreover,
\begin{align*}
\nabla_G\text{-}\mathrm{deg}(\nabla\varphi ,\Theta)& =(-1)^{m^-(B(v_0))} \chi_G(G/H^+)\ +\\ & + \sum_{(H')\in\overline{sub}[G],(H')<(H)}\nabla_G\text{-}\mathrm{deg}_{(H')}(\nabla\varphi ,\Theta)\cdot \chi_G(G/H'^+).
\end{align*}
\end{Lemma}
\begin{proof}
We can choose an open, $G$-invariant neighbourhood $\Theta$ of the $G$-orbit $G(v_0)$ such that $(\nabla\varphi)^{-1}(0)\cap\Theta= G(v_0).$ Without loss of generality we can assume $(\nabla\varphi)^{-1}(0)\cap\partial\Theta= \emptyset.$
Now, using Lemma \ref{istnieniespecfunkcjiMorse'a}, we define a $\Theta$-admissible homotopy
$H(v,t)=t\varphi(v)+(1-t)\phi(v).$
Next, the homotopy invariance property of the degree (Theorem \ref{properties_of_degree}.(3)) implies that
\[\nabla_G\text{-}\mathrm{deg}(\nabla\varphi ,\Theta)=\nabla_G\text{-}\mathrm{deg}(\nabla_vH(\cdot ,1) ,\Theta)=\nabla_G\text{-}\mathrm{deg}(\nabla_vH(\cdot ,0) ,\Theta)=\nabla_G\text{-}\mathrm{deg}(\nabla\phi ,\Theta).\]
In addition, $(\nabla\phi)^{-1}(0)\cap\Theta\subset U\left(\frac{\varepsilon}{2}\right)$ and $(\nabla\phi)^{-1}(0)\cap U\left(\frac{\varepsilon}{4}\right)=\{G(v_0)\},$ just like in Lemma
\ref{istnieniespecfunkcjiMorse'a}. Therefore using excision and additivity properties of the degree (Theorems \ref{properties_of_degree}.(4) and \ref{properties_of_degree}.(2)) and Lemma \ref{wystepujacetypyorbitowe} we can conclude that
\begin{align*}
\nabla_G \text{-deg}(\nabla \phi ,\Theta) &=\nabla_G\text{-}\mathrm{deg}(\nabla\phi ,U\left(\varepsilon/ 2\right))= \nabla_G\text{-}\mathrm{deg}(\nabla\phi ,U\left(\varepsilon/ 4\right)) +\\
& +\nabla_G\text{-}\mathrm{deg}(\nabla\phi ,U\left(\varepsilon/ 2\right)\backslash \mathrm{cl}( U\left(\varepsilon/ 4\right)))=(-1)^{m^-(B(v_0))}\chi_G(G/H^+)+\\
&+\sum_{(H')\in \overline{sub}[G],(H')<(H)}\!\!\!\!\!\!\!\!\!\!\nabla_G\text{-}\mathrm{deg}_{(H')}(\nabla\phi ,U\left(\varepsilon/ 2\right)\backslash \mathrm{cl}( U\left(\varepsilon/ 4\right)))\cdot\chi_G(G/H'^+),
\end{align*}
where by $ \mathrm{cl}( U\left(\varepsilon/ 4\right))$ we denote the closure of $ U\left(\varepsilon/ 4\right).$
\end{proof}
\setcounter{Theorem}{0}\setcounter{Definition}{0
\section{Statement of main results}
\label{bifurcations}
The following assumptions will be needed throughout this section. Let $G$ be a compact Lie group and suppose that $\Omega\subset\mathbb{V}$ is an open, $G$-invariant subset of a
finite-dimensional, real, orthogonal $G$-representation $\mathbb{V}.$
Consider a $G$-invariant $C^2$-potential $\varphi:\Omega\times\mathbb{R}\rightarrow\mathbb{R}.$ Since $\nabla_v\varphi$ is $G$-equivariant, if $(v_0,\rho_0)\in (\nabla_v\varphi)^{-1}(0)$ then $G(v_0)\times\{\rho_0\}\subset(\nabla_v\varphi)^{-1}(0).$
A set of the form $G(v_0)\times\{\rho_0\}$ consisting of critical points is called a \emph{critical $G$-orbit of $\varphi.$}
Using the techniques of equivariant bifurcation theory we will study solutions of the equation
\begin{equation}\label{eq}
\nabla _v \varphi (v,\rho )=0.
\end{equation}
Additionally, assume there exists a continuous map
$w:\mathbb{R}\rightarrow\Omega$ such that \[\mathcal{F}=\bigcup\limits_{\rho\in\mathbb{R}}G(w(\rho ))\times\{\rho\}\subset (\nabla_v \varphi )^{-1}(0).\]
The family $\mathcal{F}$ is called a \emph{trivial family of $G$-orbits} of solutions of \eqref{eq} and for any subset $X\subset\mathbb{R}$ let $\mathcal{F}_{X}$ denote the set of $G$-orbits
$\bigcup_{\rho\in X }G(w(\rho))\times\{\rho\}\subset\mathcal{F}\subset\Omega\times\mathbb{R}.$
In this section we will formulate the necessary and sufficient conditions for the existence of
bifurcations of nontrivial solutions of (\ref{eq}) from the trivial family $\mathcal{F}.$
Let us introduce a notion of a local bifurcation of $G$-orbits of solutions of (\ref{eq}).
\begin{Definition} \label{biflok}
Fix parameters $\rho^{\pm}\in\mathbb{R}$ such that $\rho^- < \rho^+.$ A local bifurcation from the segment of $G$-orbits $\mathcal{F}_{[\rho^-,\rho^+]} \subset \mathcal{F}$ of solutions
of (\ref{eq}) occurs if there exists a $G$-orbit $\mathcal{F}_{\rho_0}\subset\mathcal{F}_{[\rho^-,\rho^+]}$ such that the point $(w(\rho _0) , \rho _0)\in\mathcal{F}_{\rho_0}$ is
an accumulation point of the set $\{(v,\rho)\in (\Omega\times\mathbb{R}) \backslash\mathcal{F} :\ \nabla _v \varphi (v,\rho )=0\}$ (see Figure \ref{deflocbif}).\\
We call $\rho _0$ a parameter of local bifurcation and $\mathcal{F}_{\rho_0}$ a $G$-orbit of local bifurcation. We denote by $\mathcal{BIF}$ the set of all parameters of local bifurcation.
\end{Definition}
\begin{figure}[h!]
\centering
\setlength{\unitlength}{0.1\textwidth}
\begin{picture}(5,5)
\put(0,0){\includegraphics[height=0.4\textwidth]{def-locbif}}
\put(4.17,3.5){$\mathcal{F}$}
\put(5,0.6){$\mathbb{R}^n$}
\end{picture}
\caption{A local bifurcation from the segment of $G$-orbits $\mathcal{F}_{[\rho^-,\rho^+]} \subset \mathcal{F}.$}\label{deflocbif}
\end{figure}
According to Lemma \ref{hesjanpostac}, for each parameter $\rho\in\mathbb{R}$ we have $\dim \ker \nabla_v^2\varphi(w(\rho),\rho)\geq\dim G(w(\rho )).$
Now we will formulate the necessary condition for the existence of parameters of local bifurcation.
\begin{Theorem} \label{wkkoniecznybif}
Under the above assumptions, if, moreover, $\rho _0\in \mathcal{BIF},$ then \[\dim \ker \nabla ^2_v\varphi (w(\rho_0),\rho_0)>\dim G(w(\rho_0)).\]
\end{Theorem}
\begin{proof}
To prove this theorem we apply two kinds of implicit function theorems - the classical one and the $G$-invariant one.
Suppose, contrary to our claim, that $\rho_0\!\in\!\mathcal{BIF}$ and $\dim \ker \nabla_v^2\varphi (w(\rho _0),\rho_0)\!=\! \dim G(w(\rho _0)).$
Consider two cases: $w(\rho_0)\in \mathbb{V}^G$ and $w(\rho_0)\notin \mathbb{V}^G.$\\
\emph{Case $w(\rho_0)\in \mathbb{V}^G.$} Since $w(\rho_0)\in\mathbb{V}^G,$ we have
\[G(w(\rho _0))=\{w(\rho_0)\}\ \text{and}\ 0=\dim G(w(\rho _0))=\dim\ker \nabla_v^2\varphi (w(\rho _0),\rho _0),\]
that is $\det \nabla_v^2\varphi (w(\rho_0),\rho _0)\neq 0.$ The map $\varphi$ is of class $C^2$ and $w$ is continuous, so there exists $\varepsilon>0$ such that
$\det\nabla_v^2\varphi (w(\rho),\rho)\neq 0$ for all parameters $\rho\in (\rho_0-\varepsilon,\rho_0+\varepsilon).$
The implicit function theorem applied at point $(w(\rho_0),\rho_0)$ yields that there exist open neighbourhoods $U_{w(\rho_0)}\subset\Omega$ and $U_{\rho_0}\subset\mathbb{R}$ of the points $w(\rho_0)\in\Omega$ and $\rho_0\in\mathbb{R},$ respectively, and exactly one continuous map $\psi \ : U_{\rho_0}\rightarrow U_{w(\rho_0)}$ such that
\[\nabla_v\varphi(v,\rho)=0 \mbox{ and } (v,\rho)\in U_{w(\rho_0)}\times U_{\rho_0} \mbox{ if and only if } v=\psi(\rho).\] Therefore $w(\rho)\in \mathbb{V}^G$ for all
parameters $\rho\in U_{\rho_0},\ \psi=w\Big|_{U_{\rho_0}}$ and hence
\[(\nabla_v\varphi)^{-1}(0)\cap(U_{w(\rho_0)}\times U_{\rho_0}\;\backslash\mathcal{F})=\emptyset ,\]
which contradicts that $\rho_0\in \mathcal{BIF}.$\\
\emph{Case $w(\rho_0)\notin \mathbb{V}^G.$} As a consequence of the $G$-invariant implicit function theorem (see \cite{[DANCER]}) and Remark 4 of \cite{[DANCER]} there exist an open, $G$-invariant neighbourhood $\Theta\subset\Omega$ of the $G$-orbit $G(w(\rho_0))$ and a continuous, $G$-equivariant map
$\psi : G(w(\rho_0))\times (\rho_0-\varepsilon, \rho_0+\varepsilon)\rightarrow \mathbb{V}$ such that if
\[\mathcal{N}=\{(\psi(v,\rho),\rho):\rho\in(\rho_0-\varepsilon,\rho_0+\varepsilon),v\in G(w(\rho_0))\},\]
then for all $(v,\rho)\in\Theta\times(\rho_0-\varepsilon,\rho_0+\varepsilon)$
\begin{equation}
\nabla_v\varphi(v,\rho)=0 \ \text{if and only if}\ (v,\rho)\in\mathcal{N}.
\end{equation}
Hence for all $\rho\in (\rho_0-\varepsilon, \rho_0+\varepsilon)$ if $\nabla_v\varphi(v_1,\rho)=0$ and $\nabla_v\varphi(v_2,\rho)=0,$
where $v_1, \ v_2 \in \Theta,$ then $v_1=\psi(v_1',\rho)$ and $v_2=\psi(v_2',\rho)$ for some $v_1',\ v_2'\in G(w(\rho_0)).$ Thus there exists $g\in G$ such that $v_1'=gv_2'.$
Consequently, $v_1=\psi(v_1',\rho)=\psi(gv_2',\rho)=g\psi(v_2',\rho)=gv_2.$ On the other hand, we have the map $w :\mathbb{R}\rightarrow\Omega$ such that
$\nabla_v\varphi(w(\rho),\rho)=0$ for all parameters $\rho\in \mathbb{R}.$
Thus for all $\rho\in(\rho_0-\varepsilon,\rho_0+\varepsilon)$ if $w(\rho)\in\Theta$ then $(w(\rho),\rho)\in\mathcal{N},$ that is $w(\rho)=\psi(gw(\rho_0),\rho)$ for some $g\in G.$ Since $w$ is continuous and $w(\rho_0)\in\Theta,$ there exists small enough $\tilde{\varepsilon}\leq\varepsilon$ such that
$w(\rho)\in\Theta$ for all $\rho\in(\rho_0-\tilde{\varepsilon},\rho_0+\tilde{\varepsilon}).$
Hence $(\nabla_v\varphi)^{-1}(0)\cap(\Theta\times (\rho_0-\tilde{\varepsilon},\rho_0+\tilde{\varepsilon})\backslash\mathcal{F})=\emptyset,$ which again contradicts that $\rho_0\in \mathcal{BIF}.$
\end{proof}
After introducing the local bifurcation of $G$-orbits of solutions of (\ref{eq}) we now turn to the notion of global bifurcation of these solutions. Denote by $C(\rho_0)$ the connected component of the set
$cl(\{(v,\rho)\!\in\!\! (\Omega\times\mathbb{R}) \backslash\mathcal{F}: \nabla _v \varphi (v,\rho )=0\})\cup\mathcal{F}_{\rho_0}$ containing $\mathcal{F}_{\rho_0}$
and by $C([\rho^-,\rho^+])$ the connected component of the set $cl(\{(v,\rho)\!\in\! (\Omega\times\mathbb{R})\backslash\mathcal{F} : \nabla _v \varphi (v,\rho )=0\})\cup\mathcal{F}_{[\rho^-,\rho^+]}$ containing $\mathcal{F}_{[\rho^-,\rho^+]}.$
\begin{Definition}\label{bifglob}
Fix parameters $\rho^{\pm}\in\mathbb{R}$ such that $\rho^- < \rho^+.$ A global bifurcation from the segment of $G$-orbits $\mathcal{F}_{[\rho^-,\rho^+]} \subset \mathcal{F}$ of solutions
of (\ref{eq}) occurs if the component $C([\rho^-,\rho^+])\subset\Omega\times\mathbb{R}$
satisfies the following condition:
\begin{equation}\label{Rabinowitzalternative}
C([\rho^-,\rho^+])\ \text{is not compact or}\
(C([\rho^-,\rho^+])\backslash \mathcal{F}_{[\rho^-,\rho^+]})\cap \mathcal{F}\neq \emptyset
\end{equation}
(see Figure \ref{defglobbif}).\\
We call a parameter $\rho_0\in[\rho^-,\rho^+]$ a parameter of global bifurcation if the component $C(\rho_0)\subset\Omega\times\mathbb{R}$ is not compact or
$(C(\rho_0)\backslash \mathcal{F}_{[\rho^-,\rho^+]})\cap \mathcal{F}\neq \emptyset.$ A $G$-orbit $\mathcal{F}_{\rho_0}\subset\mathcal{F}_{[\rho^-,\rho^+]}$ is called a $G$-orbit of global bifurcation. We denote by $\mathcal{GLOB}$ the set of all parameters of global bifurcation.
\end{Definition}
\begin{figure}[h!]
\centering
\setlength{\unitlength}{0.1\textwidth}
\begin{picture}(4,4)
\put(0,0){\includegraphics[height=0.4\textwidth]{def-globbif}}
\put(3.6,3.5){$\mathcal{F}$}
\put(4.5,0.6){\makebox(0,0)[b]{$\mathbb{R}^n$}}
\end{picture}
\caption{A global bifurcation from the segment of $G$-orbits $\mathcal{F}_{[\rho^-,\rho^+]} \subset \mathcal{F}.$}\label{defglobbif}
\end{figure}
The condition \eqref{Rabinowitzalternative} is referred to as the Rabinowitz type alternative (\cite{[RABINOWITZ]}).
Notice that $\mathcal{GLOB}\subset\mathcal{BIF}.$ Below we define the bifurcation index from the segment $[\rho^-,\rho^+]$ to be an element $BIF_{[\rho^-,\rho^+]}\in U(G).$
\begin{Definition} \label{indbif}
Fix parameters $\rho^{\pm}\in\mathbb{R}, \ \rho^- < \rho^+$ such that $\rho^{\pm}\notin \mathcal{BIF}$ and fix open, $G$-invariant neighbourhoods $\Theta^{\pm}\subset\Omega$ of
$G$-orbits $G(w(\rho^{\pm}))$ such that
\[(\nabla_v\varphi)^{-1}(0)\cap(\mathrm{cl}(\Theta^{\pm})\times\{\rho^{\pm}\})=\mathcal{F}_{\rho^{\pm}}.\] We define the bifurcation index from the segment $[\rho^-,\rho^+]$ by
\[BIF_{[\rho^-,\rho^+]}=\nabla_G\text{-}\mathrm{deg}(\nabla_v\varphi(\cdot , \rho^{+}),\Theta^+)- \nabla_G\text{-}\mathrm{deg}(\nabla_v\varphi(\cdot , \rho^{-}) ,\Theta^-)\in U(G).\]
\end{Definition}
The definition of $BIF_{[\rho^-,\rho^+]}$ does not depend on the choice of the $G$-invariant neighbourhood of the $G$-orbit. In what follows we formulate a sufficient condition for the existence of global
bifurcation from the segment of solutions of $(\ref{eq})$.
\begin{Theorem}\label{twierbif1}
Fix parameters $\rho^{\pm}\in\mathbb{R}, \ \rho^- < \rho^+$ such that $\rho^{\pm}\notin \mathcal{BIF}$ and assume that $BIF_{[\rho^-,\rho^+]}\neq \mathbf{0} \in U(G).$ Then a global bifurcation
from the segment of $G$-orbits $\mathcal{F}_{[\rho^-,\rho^+]} \subset \mathcal{F}$ of solutions of (\ref{eq}) occurs.
\end{Theorem}
\begin{proof}
First, we will prove that there exists a parameter $\rho_0\in (\rho^{-},\rho^{+}) $ such that $\rho_0\in\mathcal{BIF}.$
Suppose, contrary to our claim, that $BIF_{[\rho^-,\rho^+]}\neq \mathbf{0}$ and $\mathcal{BIF}\cap [\rho^-, \rho^+]= \emptyset.$
Hence there exists an open, bounded $G$-invariant neighbourhood $U\subset\Omega\times\mathbb{R}$ of the set $\mathcal{F}_{[\rho^-,\rho^+]}$ such that
$(\nabla _v\varphi)^{-1}(0)\cap (\mathrm{cl}(U)\backslash\mathcal{F})=\emptyset .$
In particular, we can pick open, bounded $G$-invariant neighbourhoods $\Theta^{\pm}\subset\Omega$ of $G$-orbits $G(w(\rho^{\pm}))$ such that
$(\nabla_v\varphi)^{-1}(0)\cap(\mathrm{cl}(\Theta^{\pm})\times\{\rho^{\pm}\})=\mathcal{F}_{\rho^{\pm}}.$
The generalised homotopy invariance property of the degree (Theorem \ref{properties_of_degree}.(3)) yields that
\[\nabla_G\text{-}\mathrm{deg}(\nabla\varphi(\cdot , \rho^-) ,U\cap(\Omega\times\{\rho^-\}))= \nabla_G\text{-}\mathrm{deg}(\nabla\varphi(\cdot , \rho^+) ,U\cap(\Omega\times\{\rho^+\})),\]
while from excision property (Theorem \ref{properties_of_degree}.(4)) we obtain the following equalities:
\begin{align*}
\nabla_G\text{-}\mathrm{deg}(\nabla\varphi(\cdot ,\rho^-), U\cap(\Omega\times\{\rho^-\}))=\nabla_G\text{-}\mathrm{deg}(\nabla\varphi(\cdot ,\rho^-) ,\Theta^-),\\
\nabla_G\text{-}\mathrm{deg}(\nabla\varphi(\cdot , \rho^+),U\cap(\Omega\times\{\rho^+\}))= \nabla_G\text{-}\mathrm{deg}(\nabla\varphi(\cdot , \rho^+) ,\Theta^+),
\end{align*}
a contradiction.
We have thus proved that there exists a $G$-orbit of local bifurcation $\mathcal{F}_{\rho_0}$ of solutions of (\ref{eq}), or equivalently, the parameter $\rho_0$ is a parameter of local bifurcation, that is $\rho_0\in\mathcal{BIF}.$\\
Secondly, we will prove that a global bifurcation from
$\mathcal{F}_{[\rho^-,\rho^+]}$ occurs. The proof will be divided into two steps.\\
\emph{Step 1.} We first prove that $C([\rho^-,\rho^+])\neq \mathcal{F}_{[\rho^-,\rho^+]}.$ Suppose the contrary and choose an open, bounded set
$Q\subset\Omega\times\mathbb{R}$ such that $\mathcal{F}_{(\rho^-,\rho^+)}\subset Q$ and
$\Omega\times((-\infty,\rho^-)\cup(\rho^+,+\infty))\cap Q=\emptyset.$
Additionally, let
$(\Omega\times\{\rho^{\pm}\})\cap\partial Q=\mathrm{cl}(\Theta^\pm)\times\{\rho^\pm\},$
where $\Theta^{\frac{+}{}}\subset \Omega$ are chosen similarly as above, that is $\Theta^{\frac{+}{}}\subset \Omega$ are open, bounded, $G$-invariant neighbourhoods of the $G$-orbits
$G(w(\rho^{\frac{+}{}}))$
such that $(\nabla_v\varphi)^{-1}(0)\cap(\mathrm{cl}(\Theta^{\pm})\times\{\rho^{\pm}\})=\mathcal{F}_{\rho^{\pm}}.$
Observe that $\mathrm{cl}(Q)\cap (\nabla _v \varphi )^{-1}(0)$ is compact and put
\begin{align*}
K & =\mathrm{cl}(Q)\cap (\nabla _v \varphi )^{-1}(0),\\
A & =\mathcal{F}_{[\rho^{-},\rho^{+}]},\\
B & =\partial Q\cap \mathrm{cl}(\{(v,\rho)\in (\Omega\times\mathbb{R}) \backslash\mathcal{F} :\ \nabla _v \varphi (v,\rho )=0\}).
\end{align*}
These sets satisfy assumptions of Theorem \ref{separationlemma}, hence we obtain sets $K_A$ and $K_B$ from the conclusion of this theorem. Since these sets are compact and
disjoint, there exists $\eta >0$ such that $\eta$-neighbourhoods $K_A(\eta),\ K_B(\eta)$ of the sets $K_A, \ K_B$ are disjoint.
Now put
\[U=G(Q\backslash \mathrm{cl}(K_B(\frac{1}{2}\eta)))\]
and observe that $\partial U \cap (\nabla _v\varphi)^{-1}(0)= \mathcal{F}_{\rho^{\pm}}$ by Lemma \ref{lmozerach}.
The set $U$ is $G$-invariant and open by Lemma \ref{fkoustalonymgihomeo}. Again, using the generalised homotopy invariance property of the degree (Theorem \ref{properties_of_degree}.(3)), we obtain the equality
\[\nabla_G\text{-}\mathrm{deg}(\nabla\varphi(\cdot , \rho^-) ,\Theta^-)=\nabla_G\text{-}\mathrm{deg}(\nabla\varphi(\cdot , \rho^+) ,\Theta^+),\] a contradiction.\\
\emph{Step 2.} Now we prove that either the component $C([\rho^-,\rho^+])$ is not compact or that the sets $C([\rho^-,\rho^+])\backslash \mathcal{F}_{[\rho^-,\rho^+]}$ and $\mathcal{F}$ have nonempty intersection. Suppose the contrary and pick an open, bounded $\varepsilon$-neighbourhood
$(\mathrm{cl}(C([\rho^-,\rho^+])\backslash \mathcal{F}_{[\rho^-,\rho^+]}))(\varepsilon)\subset\Omega\times\mathbb{R}$ of $\mathrm{cl}(C([\rho^-,\rho^+])\backslash \mathcal{F}_{[\rho^-,\rho^+]})$ such that
$\mathrm{cl}((\mathrm{cl}(C([\rho^-,\rho^+])\backslash \mathcal{F}_{[\rho^-,\rho^+]}))(\varepsilon))\cap \mathcal{F}_{(-\infty,\rho^{-}]\cup[\rho^+,+\infty)}=\emptyset$
for some (small enough) $\varepsilon>0.$ Let $Q\subset\Omega\times\mathbb{R}$ and $\Theta^{\pm}\subset\Omega$ be the sets chosen as in Step 1 satisfying an additional condition:
$(\Theta^{\frac{+}{}}\times\{\rho^{\frac{+}{}}\})\cap (\mathrm{cl}(C([\rho^-,\rho^+])\backslash \mathcal{F}_{[\rho^-,\rho^+]}))(\varepsilon)=\emptyset.$
Now define
$Q_1=Q\cup (\mathrm{cl}(C([\rho^-,\rho^+])\backslash \mathcal{F}_{[\rho^-,\rho^+]}))(\varepsilon)$
and notice that $\mathrm{cl}(Q_1)\cap (\nabla _v \varphi )^{-1}(0)$ is a compact set. Again put
\begin{align*}
K & =\mathrm{cl}(Q_1)\cap (\nabla _v \varphi )^{-1}(0),\\
A & =C([\rho^-,\rho^+]),\\
B & =\partial Q_1\cap \mathrm{cl}(\{(v,\rho)\in (\Omega\times\mathbb{R})\backslash\mathcal{F} :\ \nabla _v \varphi (v,\rho )=0\}).
\end{align*}
The rest follows analogously to Step 1, but with $Q_1$ instead of $Q$. So defined sets satisfy the assumptions of Theorem \ref{separationlemma}, therefore we obtain $K_A$ and $K_B$ in an
identical manner as earlier. Once again they are compact and disjoint, which enables us to pick their disjoint $\eta$-neighbourhoods $K_A(\eta)$ and $K_B(\eta)$. Similarly to Step 1, we now put
\[U=G(Q_1\backslash \mathrm{cl}(K_B(\frac{1}{2}\eta)))\]
and notice that $U$ is open, $G$-invariant and $\partial U \cap (\nabla _v \varphi )^{-1}(0)= \mathcal{F}_{\rho^{\pm}}.$
Once again the generalised homotopy invariance property of the degree (Theorem \ref{properties_of_degree}.(3)) implies equality of degrees, a contradiction.
\end{proof}
A consequence of Theorem \ref{wkkoniecznybif} is that there is no $G$-orbit of any type of bifurcation among non-degenerate $G$-orbits, which is made precise by the following:
\begin{Corollary} \label{twbif1corollary}
Fix parameters $\rho^{\pm},\rho_0\in\mathbb{R}, \ \rho^- < \rho_0< \rho^+$ and assume that $\mathcal{F}_{\rho_0}$ is the only degenerate $G$-orbit in the segment $[\rho^-,\rho^+].$
Thus, if $BIF_{[\rho^-,\rho^+]}\neq \mathbf{0} \in U(G),$ then there exists exactly one $G$-orbit of local bifurcation in the segment $[\rho^-,\rho^+]$ and this orbit is precisely $\mathcal{F}_{\rho_0}.$
Moreover, it is also a $G$-orbit of global bifurcation.
\end{Corollary}
What remains is a question how to distinguish the degrees of critical orbits. It transpires that under a very mild assumption this can be done by means of a simple condition.
\begin{Theorem} \label{twchangedegree}
Fix parameters $\rho^{\pm}\in\mathbb{R}, \ \rho^- < \rho^+$ and assume additionally that the $G$-orbits $G(w(\rho^\pm))$ are non-degenerate and $(G_{w(\rho^+)})=(G_{w(\rho^-)})=(H).$ Thus, if\\
$\det B(w(\rho^-))\cdot \det B(w(\rho^+)) < 0,$ then
\[\nabla_G\text{-}\mathrm{deg}(\nabla_v\varphi(\cdot , \rho^{-}) ,\Theta^{-})\neq \nabla_G\text{-}\mathrm{deg}(\nabla_v\varphi(\cdot , \rho^{+}) ,\Theta^{+}).\]
\end{Theorem}
\begin{proof}
According to Lemma \ref{hesjanpostac}, there is a special form of $\nabla^2_v\varphi(w(\rho^{\pm}),\rho^{\pm})$ given by the formula (\ref{hesjanpostacwzor}). Observe that if $m^-(C(w(\rho^\pm)))=0,$
then
\begin{equation}\label{formula1}
\nabla_G\text{-}\mathrm{deg}(\nabla_v\varphi(\cdot , \rho^\pm) ,\Theta^\pm)= (-1)^{m^-(B(w(\rho^\pm)))}\chi_G(G/H^+).
\end{equation}
When, on the other hand, $m^-(C(w(\rho^\pm)))\neq 0,$ by Lemma \ref{postacstopnia}, we obtain that
\begin{align}\label{formula2}
\nabla_G\text{-}\mathrm{deg}(\nabla_v\varphi(\cdot, \rho^\pm) ,\Theta^\pm) & = (-1)^{m^-(B(w(\rho^\pm)))}\chi_G(G/H^+)\ +\nonumber \\ & + \sum_{(H')\in \overline{sub}[G],(H')<(H)}\nabla_G\text{-}\mathrm{deg}_{(H')}(\nabla_v\varphi(\cdot,
\rho^\pm) ,\Theta^\pm)\cdot\chi_G(G/H'^+).
\end{align}
Since $\det B(w(\rho^-))\cdot \det B(w(\rho^+)) < 0,$ we get that $(-1)^{m^-(B(w(\rho^-)))}\neq(-1)^{m^-(B(w(\rho^+)))}.$ Using the formulas \eqref{formula1} and \eqref{formula2} we complete the proof.
\end{proof}
\begin{Remark} \label{twbif1remark1}
In the case of Theorem \ref{twchangedegree} with $(G_{w(\rho^-)})\neq(G_{w(\rho^+)}),$ we get inequality of the degrees $\nabla_G\text{-}\mathrm{deg}(\nabla_v\varphi(\cdot , \rho^{\pm}) ,\Theta^{\pm}),$ what is the consequence of the formulas (\ref{formula1}) and (\ref{formula2}).
\end{Remark}
The following theorem is in the spirit of Rabinowitz's theorems regarding the problem of global bifurcation of zeros without symmetries (see \cite{[RABINOWITZ]}).
\begin{Theorem} \label{twierbif2}
Fix parameters $\rho^{\pm}\!\in\!\mathbb{R},\ \rho^{-}<\rho^{+}$ such that $\rho^\pm\notin\mathcal{BIF}$ and $BIF_{[\rho^-,\rho^+]}\neq \mathbf{0} \in U(G).$
Then a global bifurcation from the segment of $G$-orbits $\mathcal{F}_{[\rho^-,\rho^+]} \subset \mathcal{F}$ of solutions of (\ref{eq}) occurs,
that is the component $C([\rho^-,\rho^+])\subset\Omega\times\mathbb{R}$ is not compact or $(C([\rho^-,\rho^+])\backslash \mathcal{F}_{[\rho^-,\rho^+]})\cap \mathcal{F}\neq \emptyset.$
Moreover, if the component $C([\rho^-,\rho^+])$ is compact and there exist finitely many parameters
\[\rho^-_0<\rho^+_0<\ldots<\rho^-_l=\rho^-<\rho^+=\rho^+_l<\ldots<\rho^-_n<\rho^+_n\] such that
\begin{enumerate}
\item $C([\rho^-,\rho^+])\cap\mathcal{F}\subset\bigcup_{k=0}^{n}\mathcal{F}_{[\rho^-_k,\rho^+_k]},$
\item $C([\rho^-,\rho^+])\cap\mathcal{F}_{[\rho^-_k,\rho^+_k]}\neq\emptyset$ for $k=0,\ldots,n,$
\item $\mathcal{GLOB}\cap \bigcup_{k=0}^{n}\mathcal{F}_{[\rho^-_k,\rho^+_k]}=\mathrm{cl}(C([\rho^-,\rho^+])\backslash \mathcal{F}_{[\rho^-,\rho^+]})\cap\bigcup_{k=0}^{n}\mathcal{F}_{[\rho^-_k,\rho^+_k]},$
\item $\rho^\pm_k\notin\mathcal{BIF}$ for $k=0,\ldots,n,$
\end{enumerate}
then
\begin{equation}\sum_{k=0}^{n}BIF_{[\rho^-_k,\rho^+_k]}=\mathbf{0}\in U(G).\label{sumastopni}\end{equation}
\end{Theorem}
\begin{proof}
According to Theorem \ref{twierbif1}, we only need to show that the formula (\ref{sumastopni}) is valid. To prove this, fix an open and bounded $\varepsilon$-neighbourhood $(\mathrm{cl}(C([\rho^{-},\rho^{+}])\backslash \mathcal{F}_{[\rho^{-},\rho^{+}]}))(\varepsilon),$ denoted by $W$ for short, of
the set $\mathrm{cl}(C([\rho^{-},\rho^{+}])\backslash \mathcal{F}_{[\rho^{-},\rho^{+}]})$ such that
$\mathrm{cl}(W)\cap\mathcal{F}_{\mathbb{R}\backslash\bigcup_{k=0}^{n}(\rho^{-}_k,\rho^{+}_k)}=\emptyset$
for some (small enough) $\varepsilon >0.$
By assumption $(3),$ for any $k\neq l$ we can choose an open, bounded $\tilde{\varepsilon}$-neighbourhood
$(\mathrm{cl}(C([\rho^{-}_k,\rho^{+}_k])\backslash (\mathcal{F}_{[\rho^{-}_k,\rho^{+}_k]}\cup C([\rho^{-},\rho^{+}]))))(\tilde{\varepsilon}),$ denoted by $W_k$ for short, of $\mathrm{cl}(C([\rho^{-}_k,\rho^{+}_k])\backslash (\mathcal{F}_{[\rho^{-}_k,\rho^{+}_k]}\cup C([\rho^{-},\rho^{+}])))$ such that
$\mathrm{cl}(W_k)\cap\mathcal{F}_{\mathbb{R}\backslash(\rho^{-}_k,\rho^{+}_k)}=\emptyset$ for some (small enough) $\tilde{\varepsilon} >0.$
For any $k\in\{0,\ldots ,n\}$ set an open and bounded set $Q_k\subset\Omega\times\mathbb{R}$ such that
$\mathcal{F}_{(\rho^{-}_k,\rho^{+}_k)}\subset Q_k$ and
$(\Omega\times((-\infty,\rho^-_k)\cup(\rho^+_k,+\infty)))\cap Q_k=\emptyset.$
Additionally, let
$(\Omega\times\{\rho^{\pm}_k\})\cap\partial Q_k=\mathrm{cl}(\Theta^\pm_k)\times\{\rho^\pm_k\},$
where $\Theta^{\pm}_k\subset \Omega$ are an open, bounded, $G$-invariant neighbourhoods of the $G$-orbits
$G(w(\rho^{\frac{+}{}}_k))$
such that $(\nabla_v\varphi)^{-1}(0)\cap(\mathrm{cl}(\Theta^{\pm}_k)\times\{\rho^{\pm}_k\})=\mathcal{F}_{\rho^{\pm}_k}$ and
$(\Theta^{\pm}_k\times\{\rho^{\pm}_k\})\cap W=\emptyset.$
For $k\neq l$ let $\Theta_k^\pm$ satisfy an additional condition
$(\Theta_k^\pm\times \{\rho_k^\pm\})\cap W_k=\emptyset.$
Now for $k\neq l$ define $\tilde{Q}_k=Q_k\cup W_k$ and for $k=l$ let $\tilde{Q}_l=Q_l.$
Next, define
$Q= \bigcup_{k=0}^{n} \tilde{Q}_k\cup W$ and observe that $\mathrm{cl}(Q)\cap (\nabla _v \varphi )^{-1}(0)$ is a compact set. Put
\begin{align*}
K & =\mathrm{cl}(Q)\cap (\nabla _v \varphi )^{-1}(0),\\
A & =\bigcup_{k=0}^{n}\mathcal{F}_{[\rho^{-}_k,\rho^{+}_k]}\cup C([\rho^-,\rho^+])\\
B & =\partial Q\cap \mathrm{cl}(\{(v,\rho)\in (\Omega\times\mathbb{R})\backslash\mathcal{F} :\ \nabla _v \varphi (v,\rho )=0\}).
\end{align*}
Sets so defined satisfy the assumptions of Theorem \ref{separationlemma}, hence we obtain sets $K_A$ and $K_B$ from the conclusion of this theorem. Since these sets are
compact and disjoint, there exists $\eta >0$ such that $\eta$-neighbourhoods $K_A(\eta),\ K_B(\eta)$ of the sets $K_A, \ K_B$ are disjoint.
Now put
$U=G(Q\backslash \mathrm{cl}(K_B(\frac{1}{2}\eta)))$
and observe that $(\nabla_v \varphi )^{-1}(0)\cap\partial U= \bigcup_{k=0}^{n}\mathcal{F}_{\rho^-_k}\cup\mathcal{F}_{\rho^+_k}$ by Lemma
\ref{lmozerach}. The set $U$ is $G$-invariant and open by Lemma \ref{fkoustalonymgihomeo}. Using the generalised homotopy invariance property of the degree (Theorem \ref{properties_of_degree}.(3)) we obtain that \[\nabla_G\text{-}\mathrm{deg}(\nabla_v\varphi(\cdot , \rho),U\cap(\Omega\times\{\rho\}))\in U(G)\] is well defined and is constant as a function of the
$\rho\in [\rho_{min}+\varepsilon_0,\rho_{max}-\varepsilon_0],$ where
\begin{align*}
\rho_{min}&=inf\{\rho\in\mathbb{R}: U\cap(\Omega\times\{\rho\})\neq\emptyset\},\\ \rho_{max}&=sup\{\rho\in\mathbb{R}:U\cap(\Omega\times\{\rho\})\neq\emptyset\}
\end{align*}
and $\varepsilon_0$ is
a positive number satisfying
$(\nabla_v\varphi)^{-1}(0)\cap (U\cap(\Omega\times\{\rho_{\substack{min \\ max}}\pm\varepsilon_0\}))=\emptyset$
in the case of $\rho_{\text{min}}< \rho^-_0$ and $\rho_{\text{max}}>\rho_n^+.$ In other cases, $\rho\in[\rho^-_0,\rho_{max}-\varepsilon_0]$ or $\rho\in[\rho_{min}+\varepsilon_0, \rho^+_n]$ or $\rho\in[\rho^-_0,\rho^+_n].$
Observe that, using the additivity property of the degree (Theorem \ref{properties_of_degree}.(2)), we have
\begin{multline*}\nabla_G\text{-}\mathrm{deg}(\nabla_v\varphi(\cdot , \rho^+_k),U\cap(\Omega\times\{\rho^+_k\}))=\\ \underbrace{\nabla_G\text{-}\mathrm{deg}(\nabla_v\varphi(\cdot , \rho^+_k) ,\Theta^+_k)}_{a^+_k}+ \underbrace{\nabla_G\text{-}\mathrm{deg}(\nabla_v\varphi(\cdot , \rho^+_k),(U\cap(\Omega\times\{\rho^+_k\}))\backslash\mathrm{cl}(\Theta^+_k))}_{c^+_k},
\end{multline*}
analogously for parameters $\rho^-_k,k=0,\dots,n.$ We assume that $\nabla_G\text{-}\mathrm{deg}(\nabla_v\varphi(\cdot , \rho),\emptyset)=\mathbf{0}\in U(G).$
Since $a_k^++c_k^+=a_{k+1}^-+c_{k+1}^-,$ we have $a_k^+=a_{k+1}^-,$ hence we conclude that
\begin{equation*}
\sum_{k=0}^{n}a^+_k-a^-_k =\sum_{k=0}^{n}a^+_k-\sum_{k=0}^{n}a^-_k=\sum_{k=0}^{n}a^+_k-\sum_{k=-1}^{n-1}a^-_{k+1}=\sum_{k=0}^{n-1}(a^+_k-a^-_{k+1})+a^+_n -a^-_0=a^+_n-a^-_0.
\end{equation*}
Observe also that $c^+_n=c^-_0=0.$ Then
\begin{multline*}
\sum_{k=0}^{n} BIF_{[\rho^-_k,\rho^+_k]}=
\sum_{k=0}^{n}\nabla_G\text{-}\mathrm{deg}(\nabla_v\varphi(\cdot , \rho^+_k),\Theta^+_k)-\nabla_G\text{-}\mathrm{deg}(\nabla_v\varphi(\cdot , \rho^-_k),\Theta^-_k)=\\
=\sum_{k=0}^{n}a_k^+-a_k^-=a^+_n-a^-_0=(a_n^++c_n^+)-(a_0^-+c_0^-)=\mathbf{0}\in U(G),
\end{multline*}
which proves our assertion.
\end{proof}
\begin{Remark}
Observe that in Theorem \ref{twierbif2} if for $k=0,\ldots,n$ the set $\mathcal{F}_{[\rho^-_k,\rho^+_k]}$ contains exactly one degenerate critical orbit, then the assumption $(3)$ of this theorem is satisfied.
\end{Remark}
Now fix $(v_0,\rho_0)\in(\nabla_v\varphi)^{-1}(0)$ and let
\[\mathbb{V}^+=\bigoplus_{\lambda\in \sigma^{+}(-\nabla_v^2\varphi(v_0,\rho_0))}\mathbb{V}_{-\nabla^2_v\varphi(v_0,\rho_0)}(\lambda),\]
where $\mathbb{V}_{-\nabla^2_v\varphi(v_0,\rho_0)}(\lambda)$ denotes the eigenspace of $-\nabla^2\varphi(v_0,\rho_0)$ associated to $\lambda.$
Notice that a non-degenerate critical $G$-orbit $G(v_0)\subset\Omega$ of the map $\varphi(\cdot,\rho_0)$ is an isolated $\eta_{\rho_0}$-invariant set, where $\eta_{\rho_0}:U\subset\Omega\times\mathbb{R}\rightarrow\mathbb{V}$ is a local $G$-flow induced by the equation
\[\dot{v}(t)=-\nabla_v\varphi(v(t),\rho_0).\]
For simplicity of notation, $G_{v_0}=H.$ According to Lemma \ref{hesjanpostac}, there is a special form of $\nabla_v^2\varphi(v_0,\rho_0)$ given by the formula (\ref{hesjanpostacwzor}), where \[\mathbb{V}=T_{v_0}G(v_0)\oplus\mathbb{W}^H\oplus(\mathbb{W}^H)^{\bot}=T_{v_0}G(v_0)\oplus\mathbb{V}^+\oplus(\mathbb{V}^+)^{\bot}.\]
Assume that the non-degenerate critical $G$-orbit $G(v_0)$ is special, that is $m^-(C(v_0))=0.$
Then the $G$-equivariant Conley index of the critical $G$-orbit $G(v_0)$ has the following $G$-homotopy type:
\[CI_G(G(v_0),\eta_{\rho_0})=([(G\times_H D(\mathbb{V}^+))/(G\times_H S(\mathbb{V}^+))]_G ,[*]),\]
where $D(\mathbb{V}^+),\ S(\mathbb{V}^+)$ denote a closed ball and a sphere in the space $\mathbb{V}^+,$ respectively.
Observe that since $m^-(C(v_0))=0,$ we have $S(\mathbb{V}^+)\subset D(\mathbb{V}^+)\subset\mathbb{V}^+ \subset\mathbb{W}^H$ and $m^-(B(v_0))=\dim\mathbb{V}^+,$ hence
\begin{equation}
CI_G(G(v_0),\eta_{\rho_0})=([(G/H\times D(\mathbb{V}^+))/(G/H\times S(\mathbb{V}^+))]_G ,[*]).
\end{equation}
\begin{Remark}\label{G-CW-complex}
Additionally, there is a connection between the $G$-equivariant Conley index of the special critical $G$-orbit $G(v_0)$ and a notion of finite, pointed $G$-$CW$-complex. The $G$-equivariant Conley index $CI_G(G(v_0),\eta_{\rho_0})$ has the $G$-homotopy type of a pointed $G$-$CW$-complex which consists of the base point $*$ and one $m^-(B(v_0))$-dimensional cell of orbit type $(H).$
\end{Remark}
\begin{Remark}\label{AchangeofG-CW-complex}
Thus, if Morse indices of two special critical $G$-orbits are different, then the $G$-equivariant Conley indices have different $G$-homotopy types. It is known (see \cite{[SMOLLER]}) that a change in the $G$-equivariant Conley index implies the existence of $G$-orbits of local bifurcation of critical points of $G$-equivariant variational problems.
\end{Remark}
As a consequence of Remark \ref{G-CW-complex} and Remark \ref{AchangeofG-CW-complex}, there are simple conditions verifying the existence of a local bifurcation.
\begin{Theorem}\label{twbiflok1}
Fix parameters $\rho^{\pm}\!\in\!\mathbb{R},\ \rho^{-}<\rho^{+}$ such that:
\begin{enumerate}
\item[(1)] $\dim \ker \nabla_v ^2\varphi (w(\rho ^{\pm}),\rho ^{\pm}) = \dim G(w(\rho^{\pm})),$
\item[(2)] $(G_{w(\rho^-)})=(G_{w(\rho^+)}),$
\item[(3)] $m^-(C(w(\rho^{\pm})))=0.$
\end{enumerate}
If $m^{-}(B(w(\rho^-)))\neq m^{-}(B(w(\rho^+))),$ then $(\rho^-,\rho^+)\cap\mathcal{BIF}\neq\emptyset,$ that is there exists a local bifurcation parameter $\rho_0\in(\rho^-\!,\rho^+).$
\end{Theorem}
\begin{proof}
First observe that, according to Remark \ref{G-CW-complex}, the $G$-equivariant Conley indices of the special critical $G$-orbits $G(w(\rho^{\pm}))$ have the $G$-homotopy type of a pointed $G$-$CW$-complex which consists of the base point $*$ and one $m^-(B(w(\rho^\pm)))$-dimensional cell of orbit type $(G_{w(\rho^-)})=(G_{w(\rho^+)}).$ Since $m^{-}(B(w(\rho^-)))\neq m^{-}(B(w(\rho^+))),$ we conclude, by Remark \ref{AchangeofG-CW-complex}, that $(\rho^-,\rho^+)\cap\mathcal{BIF}\neq\emptyset.$
\end{proof}
\begin{Theorem}\label{twbiflok2}
Fix parameters $\rho^{\pm}\!\in\!\mathbb{R},\ \rho^{-}<\rho^{+}$ such that:
\begin{enumerate}
\item[(1)] $\dim \ker \nabla_v ^2\varphi (w(\rho ^{\pm}),\rho ^{\pm}) = \dim G(w(\rho^{\pm})),$
\item[(2)] $(G_{w(\rho^-)})\neq(G_{w(\rho^+)}),$
\item[(3)] $m^-(C(w(\rho^{\pm})))=0.$
\end{enumerate}
Then $(\rho^-,\rho^+)\cap\mathcal{BIF}\neq\emptyset,$ that is there exists a local bifurcation parameter $\rho_0\in(\rho^-\!,\rho^+).$
\end{Theorem}
\begin{proof}
Analogously to the proof of Theorem \ref{twbiflok1}, the $G$-equivariant Conley indices of the special critical $G$-orbits $G(w(\rho^{\pm}))$ have the $G$-homotopy type of a pointed $G$-$CW$-complex which consists of the base point $*$ and one $m^-(B(w(\rho^\pm)))$-dimensional cell of orbit type $(G_{w(\rho^\pm)}),$ see Remark \ref{G-CW-complex}. Since $(G_{w(\rho^-)})\neq (G_{w(\rho^+)}),$ applying Remark \ref{AchangeofG-CW-complex} once again, we get that $(\rho^-,\rho^+)\cap\mathcal{BIF}\neq\emptyset.$
\end{proof}
\setcounter{Theorem}{0}\setcounter{Definition}{0
\section{Applications}
\label{applications}
In this section we apply our results to show the existence of new families of central configurations, which bifurcate from certain known families of central configurations. We consider $N$ bodies of positive masses $m_1,\ldots ,\ m_N$ in the $2$-dimensional Euclidean vector space, whose positions are denoted by $q_1,\ldots ,q_N\in\mathbb{R}^2$.
The equations of motion are given by
\[m_j\ddot{q}_j=-\sum_{i=1,i\neq j}^N\mathbf{G}m_jm_i\frac{q_j-q_i}{|q_j-q_i|^3},\ j=1,\ldots, N,\]
where $\mathbf{G}$ is the gravitational constant, which can be set equal to one.
From now on we will treat the space $\mathbb{R}^{2N}$ as an $SO(2)$-representation $\mathbb{V},$ that is a representation of the Lie group $SO(2)$, which is a direct sum of $N$
copies of the natural, orthogonal $SO(2)$-representation (that is the usual multiplication of a vector by a matrix). The action of $SO(2)$ on $\mathbb{V}$ is given by
\[
SO(2)\times\mathbb{V} \ni (g,q)=(g,(q_1,\ldots ,q_N))\mapsto g\cdot q =(g\cdot q_1,\ldots ,g\cdot q_N)\in\mathbb{V}.
\]
Newtonian equations of motion associated with a potential $U$ have the following form:
\begin{equation} \label{eqr} M\ddot{q}=\nabla_q U(q,m),
\end{equation}
where $M$ is a mass matrix, that is $M=diag (m_1,m_1, \ldots , m_N, m_N),$ and the Newtonian potential $U : \Omega\times(0,+\infty)^N\rightarrow\mathbb{R}$ is given by
\[U(q,m)=U(q_1,\ldots ,q_N,m_1,\ldots ,m_N)=\sum\limits_{1 \leq i<j\leq N}\frac{m_im_j}{|q_i-q_j|},\]
where the set $\Omega\subset\mathbb{V}$ is defined by
\[\Omega=\{q\!=\!(q_1,\ldots ,q_N)\in\mathbb{V} : q_i\neq q_j \ for \ i\neq j\}.\]
We can assume without loss of generality that the centre of mass of the bodies is at the origin of the coordinate chart. Observe that the set
$\Omega\subset \mathbb{V}$ is $SO(2)$-invariant and open, while the potential $U$ is an $SO(2)$-invariant $C^\infty$-map.
\begin{Definition}
A configuration $q=(q_1,\ldots,q_N)\in\Omega$ is called a central configuration of the system
(\ref{eqr}) if there exists a positive constant $\lambda$ such that $\ddot{q}=-\lambda q.$
\end{Definition}
This is equivalent to saying that the following condition is fulfilled:
\begin{equation}\label{eqeq}
-\lambda\nabla_q I(q,m)=\nabla _qU(q,m),
\end{equation}
where $I:\Omega\times(0,+\infty)^N\rightarrow\mathbb{R}$ given by the formula
\[
I(q,m)=\frac{1}{2}\sum\limits_{j=1}^{N}m_j|q_j|^2
\]
is the moment of inertia.
Using Euler's theorem and the fact that $U$ is homogeneous of degree $-1$ one can show that $\lambda=\frac{U(q,m)}{2I(q,m)}.$ Moreover,
for any central configuration $q$ it is known that $rq$ and
$g\cdot q$ are central configurations with new $\lambda$ coefficients equal to $\frac{\lambda}{|r|^3}$
and $\lambda$ respectively, where $r\in\mathbb{R}$ and $ g\in SO(2).$
Thus this way we can introduce an equivalence relation by saying that two central configurations are equivalent if
we can pass from one to another by a composition of a rotation and a scaling.
However, instead of passing to the quotient and treating classes as single points we shall study the whole $SO(2)$-orbits of central configurations. In this approach our problem becomes a problem of finding critical $SO(2)$-orbits of a smooth, $SO(2)$-invariant potential
$\hat{\varphi} : \Omega\times (0,+\infty)^N\rightarrow\mathbb{R}$ given by
\[
\hat{\varphi}(q,m)=U(q,m)+\lambda I(q,m).
\]
As we mentioned earlier we search bifurcations of central configurations from known families, so we assume additionally that there exist continuous maps $w:(0,+\infty)\rightarrow\Omega$ and $m:(0,+\infty)\rightarrow(0,+\infty)^N$ such that
\[\nabla_q\hat{\varphi}(w(\rho),m(\rho))=0.\]
Now define a potential $\varphi:\Omega\times(0,+\infty)\rightarrow\mathbb{R}$ by
\[\varphi(q,\rho)=\hat{\varphi}(q,m(\rho)).\]
Then
\[\mathcal{F}=\bigcup_{\rho\in(0,+\infty)}SO(2)(w(\rho ))\times \{\rho \}\subset (\nabla _q \varphi )^{-1}(0).\]
We shall refer to $\mathcal{F}$ as the \emph{trivial family of solutions}.
Thus, we will consider the following equation:
\begin{equation} \label{eqcc}
\nabla_{q}\varphi(q,\rho)=0,
\end{equation}
where $\lambda=\lambda(\rho)=\frac{U(w(\rho),m(\rho))}{2I(w(\rho),m(\rho))}.$
Now we will apply our main Theorems \ref{wkkoniecznybif} and \ref{twierbif1} to formulate necessary and sufficient conditions for the existence of bifurcations of central configurations from our known families $w$. Note that for families $w$ considered in this paper we have $SO(2)_{w(\rho)}=Id$ thus the matrix $C(w(\rho))$ is zero dimensional and in particular, its Morse index is equal to zero.
Now, as has been indicated earlier, we pass from seeking bifurcations of central configurations to seeking bifurcations of critical $SO(2)$-orbits of the $SO(2)$-invariant potential
\[\varphi(q,\rho)\!=\!U(q,m(\rho))+ \lambda(\rho) I(q,m(\rho)).\]
For this we seek parameters $\rho$ satisfying the necessary condition for the existence of a local bifurcation, which is given by Theorem \ref{wkkoniecznybif}.
Because of tedious computations we aid ourselves by the algebraic processor MAPLE$\texttrademark$ to calculate the Hessian $\nabla^2_q\varphi$ of the potential $\varphi$ and its characteristic polynomial $W_\rho$ along the curve $w.$ Denote the symbolic form of the latter by
\[W_\rho(q)=q^{2N}-a_1(\rho)q^{2N-1}+\cdots + (-1)^{2N-1}a_{2N-1}(\rho)q+(-1)^{2N} a_{2N}(\rho).\]
Notice that $\dim\ker \nabla^2_q \varphi(w(\rho),\rho)=k$ if and only if
$a_{2N}=\cdots=a_{2N-k+1}=0$ and $a_{2N-k}\neq 0.$ Additionally in this case $a_{2N-k}$ is the product of the nonzero eigenvalues. Also recall that
$\dim \ker \nabla_q^2\varphi (w(\rho),\rho)\geq \dim SO(2)(w(\rho))=1,$ so $a_{2N}(\rho)=0.$ For this reason we consider the equation
\[a_{2N-1}(\rho)=0,\]
whose solutions are the only parameters satisfying the necessary condition from Theorem \ref{wkkoniecznybif}.
In the next step we apply Theorems \ref{twierbif1}, \ref{twchangedegree} and \ref{twbiflok1} to each candidate for the parameter of bifurcation to separate the ones which also satisfy the sufficient conditions.
For each candidate $\rho^*$ we choose its small enough $\varepsilon$-neighbourhood for which we have $\dim \ker \nabla_q ^2\varphi (w(\rho^*\pm\varepsilon),\rho^*\pm\varepsilon) = \dim SO(2)(w(\rho^*\pm\varepsilon)).$ If
\[ m^{-}(B(w(\rho^*-\varepsilon)))\neq m^{-}(B(w(\rho^*+\varepsilon))),\] then there exists a parameter of local bifurcation in this neighbourhood. Moreover, if these Morse indices differ by an odd number a global bifurcation occurs.
\begin{figure}[h!]
\centering
\includegraphics[height=0.4\textwidth]{two_squares}
\caption{Eight bodies located at the vertices of two nested squares.}\label{two_squares}
\end{figure}
We will now focus on our goal, which is studying bifurcations from the known families of central configurations, which we will now consider as our trivial family of solutions $w.$ First, we consider nested planar central configuration for the problem of $8$ bodies.
This configuration was found in $2013$ by A. C. Fernandes, L. F. Mello and M. M. da Silva in \cite{[FERNANDES]} and consists of $4$ bodies with equal masses located at the vertices of a square whose side is equal to $1$ and the other $4$ bodies, also with equal masses, located at the vertices of a smaller square whose side is equal to $0<b<0.53177...,$ whose centre coincides with the centre of the first square (see Figure \ref{two_squares}). Denote by $r$ the ratio of the circumcircle of the inner square. We treat $0<r<r_0=0.37602...$ as a parameter. Let $m_1=m_2=m_3=m_4=M_r=-\frac{B(r)}{A(r)}$ and $m_5=m_6=m_7=m_8=1,$
where
\begin{align*}
A(r)&=(R_{1,2}-R_{1,5})\Delta_{1,5,2}+(R_{1,3}-R_{1,6})\Delta_{1,6,3}+(R_{1,7}-R_{1,2})\Delta_{1,7,2},\\
B(r)&=(R_{6,7}-R_{1,5})\Delta_{1,5,6}+(R_{1,7}-R_{6,7})\Delta_{5,6,3}+(R_{1,6}-R_{5,7})\Delta_{1,6,8}
\end{align*}
and $R_{i,j}=1/|q_i-q_j|^3,\ \Delta_{i,j,k}=(q_i-q_j)\wedge (q_i-q_k)$ for $1\leq i,j,k\leq 8.$
We define $w:(0,r_0)\rightarrow\Omega$ by
\begin{align}\label{family_two_squares}\displaystyle
w(r)\!\!&=\!\!\left(\!\!-\frac{1}{2},-\frac{1}{2},\frac{1}{2},-\frac{1}{2},\frac{1}{2},\frac{1}{2},-\frac{1}{2},\frac{1}{2}, -\frac{\sqrt{2}}{2}r,-\frac{\sqrt{2}}{2}r, \frac{\sqrt{2}}{2}r,-\frac{\sqrt{2}}{2}r, \frac{\sqrt{2}}{2}r,\frac{\sqrt{2}}{2}r, -\frac{\sqrt{2}}{2}r,\frac{\sqrt{2}}{2}r\!\!\right)\!\!\end{align}
and $m:(0,r_0)\rightarrow(0,+\infty)^8$ as follows
\[m(r)=(M_r,M_r,M_r,M_r,1,1,1,1).\]
\begin{Lemma}\label{lemma1}
Put $ r_1=\frac{1}{7}\sqrt{2}, \ r_2=\frac{1}{6}\sqrt{2}$ and $r_3=\frac{1}{5}\sqrt{2}.$
Then $\dim \ker \nabla_q ^2\varphi (w(r_i),r_i) = \dim SO(2)(w(r_i))=1$ for $i=1,2,3$ and
the Morse index of $\nabla^2_{q}\varphi(w(\cdot),\cdot)$ evaluated at $r_i$ is
\begin{align*}
m^- (\nabla^2_{q}\varphi (w(r_i),r_i))=\left\{ \begin{array}{lrl}
1, & for & i=1 \\
3, & for & i=2\\
4, & for & i=3
\end{array}\right. .\end{align*}
\end{Lemma}
So, there is no matrix $C(w(r))$ and $SO(2)_{w(r)}=Id$ for any $0<r<r_0=0.37602...$ Also, by Lemma \ref{lemma1}, we have $\dim \ker \nabla_q ^2\varphi (w(r_i),r_i) = \dim SO(2)(w(r_i))=1$ for $i=1,2,3.$ Thus, by Theorem \ref{twbiflok1}, since $m^- (\nabla^2_{q}\varphi (w(r_1),r_1))\neq m^- (\nabla^2_{q}\varphi (w(r_2),r_2)),$ there exists a local bifurcation parameter in the segment $(r_1,r_2).$ Additionally, by Theorem \ref{twierbif1} and Theorem \ref{twchangedegree}, since the numbers $m^- (\nabla^2_{q}\varphi (w(r_2),r_2))$ and $m^- (\nabla^2_{q}\varphi (w(r_3),r_3))$ are of different parity, we have that there exists a global bifurcation parameter in $(r_2,r_3).$
Summarising, we have proved the following theorem:
\begin{Theorem}\label{thmtwosquare}$ $
\begin{enumerate}
\item There exists a local bifurcation parameter in the segment $(r_1,r_2),$ that is $(r_1,r_2)\cap \mathcal{BIF}\neq\emptyset .$
\item There exists a global bifurcation parameter in the segment $(r_2,r_3),$ that is $(r_2,r_3)\cap \mathcal{GLOB}\neq\emptyset .$
\end{enumerate}
\end{Theorem}
Notice that we can consider a subset of the full configuration space $\Omega$ which is invariant for the gradient flow (the set of central configurations of two nested squares type $(r,m,M)$), then studying central configurations in this set becomes a problem of studying zeros of a function $F:(0,r_0)\times(0,+\infty)^2\rightarrow\mathbb{R}$ given by the formula $F(r,m,M)=MA(r)+mB(r).$ For the trivial family of solutions $(r,m,\frac{-mB(r)}{A(r)}),$ for any $(r,m)\in (0,r_0)\times(0,+\infty),$ we have
$F\left(r,m,M(r,m)\right)=0,$ where $M(r,m)=\frac{-mB(r)}{A(r)},$ and $F'_r\left(r,m,M(r,m)\right)=M(r,m)A'(r)+mB'(r)>0,$ so by the implicit function theorem there is no bifurcation of central configurations of two nested squares type from the trivial family. Combining Theorem \ref{thmtwosquare} and the above, we obtain that the families which bifurcate from the family \eqref{family_two_squares} are not of two nested squares type.
\begin{figure}[h!]
\centering
\includegraphics[height=0.4\textwidth]{rosette}
\caption{Thirteen bodies in rosette configuration.}\label{rosette}
\end{figure}
Now consider a rosette central configuration consisting of $n$ particles of mass $m_1$ lying at the vertices of a regular $n$-gon, $n$ particles of mass $m_2$ lying at the vertices of another $n$-gon, where the second one is rotated by $\frac{\pi}{n},$ and an additional particle of mass $m_0$ lying at the common centre of the two $n$-gons (see \cite{[LEIANDSANTOPRETE]} and \cite{[SEKIGUCHI]}). We analyse bifurcations from this family in the case of $13$ bodies (see Figure \ref{rosette}) which are considered in the following positions:\label{appendixB}
\begin{center}
$\hat{q}_{k+1}=\Phi\left(\frac{2\pi}{6}k\right)(r_1,0),\ \hat{q}_{7+k}=\Phi\left(\frac{\pi}{6}+\frac{2\pi}{6}k\right)(r_2,0)$ for $k=0,\ldots ,5$ and $\hat{q}_{13}=(0,0),$
\end{center}
where by $\Phi$ we denote the universal covering of $SO(2)$ by $\mathbb{R}$, that is a mapping $\Phi : \mathbb{R} \to SO(2)$ given by the formula
\[
\Phi(s) =\left( \begin{array}{lr} \cos s & -\sin s \\ \sin s & \cos s \end{array}\right).
\]
We transform the coordinates $r_1$ and $r_2$ in the following way: $r_1=r\cos \theta,\ r_2=r\sin\theta,$ where $(r,\theta)\in (0,+\infty)\times(0,\frac{\pi}{2})$ (see \cite{[LEIANDSANTOPRETE]}).
We choose $\theta=\frac{\pi}{3},\ r=1$ and define $w:(0,+\infty)^2\rightarrow\Omega$ by
\begin{align}\label{family_rosette}
w(m_0,m_1)=\left(\hat{q}_1,\ldots,\hat{q}_{12},\hat{q}_{13}\right),
\end{align}
where masses $m_0$ and $m_1$ are treated as parameters.
Notice that for the two-parameter family $w$ we obtain the same results as in Theorem \ref{wkkoniecznybif} and Theorem \ref{twierbif1}. Namely, for any points $(m_0^-,m_1^-)$ and $(m_0^+,m_1^+)$ in the parameter space we can choose a one-parameter path joining these points, that is the segment connecting them.
Then the configuration $\hat{q}\!=\!(\hat{q}_1,\!\ldots\!,\hat{q}_{12},\hat{q}_{13})$ is central for each parameter $(m_0,m_1)\!\in\!(0,+\infty)^2$ if
\begin{multline*}
m_2=m_2(m_0,m_1)= \frac{1}{ -1862\sqrt {3}-7203+810\sqrt {7}+90\sqrt {7}\sqrt {3} }\left( -7644m_0 \right. \\
\left. + 6\left(81\sqrt {7}-441\sqrt {3}+9\sqrt {3}\sqrt {7}-147 \right)m_1 \right) .
\end{multline*}
So, we define $m:(0,+\infty)^2\rightarrow(0,+\infty)^{13}$ by the formula
\begin{multline}\nonumber
m(m_0,m_1)=(m_1,m_1,m_1,m_1,m_1,m_1,m_2(m_0,m_1),m_2(m_0,m_1),m_2(m_0,m_1),\\m_2(m_0,m_1),m_2(m_0,m_1),m_2(m_0,m_1),m_0)
\end{multline}
and compute the characteristic polynomial $W_{(m_0,m_1)}$ along the curve $w.$
Then, we get that
\[a_{26}(m_0,m_1)=0\]
and \[a_{25}(m_0,m_1)=C(m_0,m_1)C_1^2(m_0,m_1)C_2(m_0,m_1) C_3^2(m_0,m_1),\]
where polynomials $C_1,\ C_2$ and $C_3$ are equal to $0$ on some curves (see Figure \ref{rosette_zeros}) and $C(m_0,m_1)\neq 0$ for $(m_0,m_1)\in(0,+\infty)^2.$
\begin{figure}[h!]
\centering
\includegraphics[height=0.4\textwidth]{rosette_zeros}
\caption{The set of zeros of the coefficient $a_{25}$ and regions $c_3,\ c_5,\ c_2$ and $c_0$ where $a_{25}\neq 0$.}\label{rosette_zeros}
\end{figure}
\begin{Lemma}\label{lemma2}
The Morse index of $\nabla^2_{q}\varphi $ along $w$ depends on $(m_0,m_1)$ as follows
\begin{align*}
m^- (\nabla^2_{q}\varphi (w(m_0,m_1),(m_0,m_1)))=\left\{ \begin{array}{lrl}
3, & for & (m_0,m_1) \in c_3\\
5, & for & (m_0,m_1) \in c_5\\
2, & for & (m_0,m_1) \in c_2\\
0, & for & (m_0,m_1) \in c_0\\
\end{array}\right.
\end{align*}
and
\[\dim \ker \nabla_q^2\varphi (w(m_0,m_1),(m_0,m_1))= \dim SO(2)(w(m_0,m_1))=1\]
for $(m_0,m_1)\in(0,+\infty)^2$ such that $C_i(m_0,m_1)\neq 0$ for all $i=1,2,3.$
\end{Lemma}
So, for any $(m_0,m_1)\in (0,+\infty)^2$ we have that $SO(2)_{w(m_0,m_1)}=Id$ and
there is no matrix $C(w(m_0,m_1)).$
Lemma \ref{lemma2} now shows that $\dim \ker \nabla_q^2\varphi (w(m_0,m_1),(m_0,m_1))=\dim SO(2)(w(m_0,m_1))=1$ for $(m_0,m_1)\!\!\in\!\!(0,+\infty)^2$ such that $C_i(m_0,m_1)\neq 0$ for all $i\in\{1,2,3\}.$ Thus, by Theorem \ref{twbiflok1}, since the numbers $m^- (\nabla^2_{q}\varphi (w(m_0,m_1),(m_0,m_1)))$ are different in the regions $c_3,\ c_5,\ c_2$ and $c_0,$ we have $(m_0,m_1)\in\mathcal{BIF}$ for $(m_0,m_1)\in(0,+\infty)^2$ with $C_i(m_0,m_1)=0,$ where $i\in\{1,2,3\}.$ Moreover, by Theorem \ref{twierbif1} and Theorem \ref{twchangedegree}, since the numbers $m^- (\nabla^2_{q}\varphi (w(m_0,m_1),(m_0,m_1)))$ are of different parity in the regions $c_3$ and $c_2,$ we have $(m_0,m_1)\in\mathcal{GLOB}$ for $(m_0,m_1)\in(0,+\infty)^2$ such that $C_2(m_0,m_1)=0.$
Summarising, we have proved the following theorem:
\begin{Theorem}\label{thmrosette} $ $
\begin{enumerate}
\item For any $i\in\{1,2,3\},$ if $C_i(m_0,m_1)=0,$ then $(m_0,m_1)\in\mathcal{BIF}.$
\item If $C_2(m_0,m_1)=0,$ then $(m_0,m_1)\in\mathcal{GLOB}.$
\end{enumerate}
\end{Theorem}
Notice that we can consider a subset of the full configuration space $\Omega$ which is invariant for the gradient flow (the set of central configurations of rosette type $(\frac{r_2}{r_1},m_0,m_1,m_2)$), then studying central configurations in this set becomes a problem of studying zeros of a function $F:(0,+\infty)^3\rightarrow\mathbb{R},\ F(x,\varepsilon,\mu)$ given by the formula $(6)$ from \cite{[LEIANDSANTOPRETE]}, where $x=\frac{r_2}{r_1},\ \varepsilon=\frac{m_2}{m_1}$ and $\mu=\frac{m_0}{m_1}.$ For the trivial family of solutions \eqref{family_rosette}, for any $m_0,\ m_1 \in(0,+\infty),$ we have $F\left(\sqrt{3},\frac{m_2(m_0,m_1)}{m_1},\frac{m_0}{m_1}\right)=0$ and $F'_x\left(\sqrt{3},\frac{m_2(m_0,m_1)}{m_1},\frac{m_0}{m_1}\right)>0,$ so by the implicit function theorem there is no bifurcation of central configurations of rosette type from this family. Combining Theorem \ref{thmrosette} and the above, we obtain that the families which bifurcate from the trivial one are not of
rosette type.
\begin{figure}[h!]
\centering
\includegraphics[height=0.4\textwidth]{rosette_zeros_oneparameter}
\caption{One-parameter family of rosette configuration where $m_1$ is a parameter and the set of four parameters of local bifurcation.}\label{rosette_zeros_oneparameter}
\end{figure}
Now we fix $m_0^*\in (0,+\infty)$ and treat $w(m_0^*,\cdot )$ and $m(m_0^*,\cdot )$ as one-parameter families where $m_1$ is a parameter. Observe that there are exactly four parameters $m_1$ for which critical orbits $SO(2)(w(m_0^*,m_1))$ are degenerate and there is at most one parameter $m_1^*$ such that $(m_0^*,m_1^*)\in\mathcal{GLOB}$ (see Figure \ref{rosette_zeros_oneparameter}). Then choose $\varepsilon>0$ (small enough) such that $(m_0^*,m_1^*\pm\varepsilon)\notin\mathcal{BIF}$ and there is only one degenerate critical orbit $SO(2)(w(m_0^*,m_1^*))$ in the segment $[m_1^*-\varepsilon,m_1^*+\varepsilon].$ It follows that
\begin{multline*}
BIF_{[m_1^*-\varepsilon,m_1^*+\varepsilon]}\!\!=\!\nabla_G\text{-}\mathrm{deg}(\nabla_v\varphi(\cdot ,(m_0^*, m_1^*+\varepsilon)),\Theta^+)- \nabla_G\text{-}\mathrm{deg}(\nabla_v\varphi(\cdot, (m_0^* ,m_1^*-\varepsilon)) ,\Theta^-)=\\
=\!(-1)^{m^-(B(w(m_0^*,m_1^*+\varepsilon)))}\!\chi_{SO(2)}(SO(2)/\{Id\}^+)-(-1)^{m^-(B(w(m_0^*,m_1^*-\varepsilon)))}\!\chi_{SO(2)}(SO(2)/\{Id\}^+)\\
=((-1)^{m^-(B(w(m_0^*,m_1^*+\varepsilon)))}-(-1)^{m^-(B(w(m_0^*,m_1^*-\varepsilon)))})\chi_{SO(2)}(SO(2)/\{Id\}^+)=\\
=((-1)^{3}-(-1)^{2})\chi_{SO(2)}(SO(2)/\{Id\}^+)=-2\chi_{SO(2)}(SO(2)/\{Id\}^+)\neq\mathbf{0}\in U(G).
\end{multline*}
Similarly, for any $m_1\neq m_1^*$ since the numbers $m^-(B(w(m_0^*,m_1\pm\tilde{\varepsilon})))$ are of the same parity, we conclude that $BIF_{[m_1-\tilde{\varepsilon},m_1+\tilde{\varepsilon}]}=\mathbf{0},$ where $\tilde{\varepsilon}$ is chosen as above.
Therefore, by Theorem \ref{twierbif2}, $C([m_1^*-\varepsilon,m_1^*+\varepsilon])$ is not compact.
|
\section{\@startsection {section}{1}{\z@}{-3.5ex plus -1ex minus
-.2ex}{2.3ex plus .2ex}{\large\bf}}
\makeatother
\def\bfm#1{\mbox{\boldmath$#1$}}
\def\R{\bfm R}
\def\N{\bfm N}
\def\k{\kappa}
\def\x{\bfm x}
\def\Z{\bfm Z}
\def\z{\bfm z}
\def\Si{\bfm \Sigma}
\def\si{\sigma}
\def\y{\bfm y}
\def\u{\bfm u}
\def\0{\bfm 0}
\def\e{\bfm e}
\def\bmu{\bfm \mu}
\def\al{\alpha}
\newcommand{\qedd}{\hfill\rule{1 ex}{1 ex}}
\newcommand{\comment}[1]{\textcolor{black}{\textbf{#1}}}
\newcommand{\bluecomment}[1]{\textcolor{black}{\textrm{#1}}}
\newcommand{\redcomment}[1]{\textcolor{black}{\textrm{#1}}}
\DeclareMathAlphabet{\mathpzc}{OT1}{pzc}{m}{it}
\newcommand{\de}{\backslash}
\def\theequation{\thesection.\arabic{equation}}
\def\bfm#1{\mbox{\boldmath$#1$}}
\begin{document}
\title{\bf Finding Connected Dense $k$-Subgraphs \thanks{Research supported in part by by NNSF of China under Grant No. 11222109, 11021161 and 10928102,
by 973 Project of China under Grant No. 2011CB80800, and by CAS Program for Cross \& Cooperative Team of Science \& Technology Innovation.}}
\author{Xujin Chen, Xiaodong Hu, Changjun Wang}
\date{Institute of Applied Mathematics, AMSS, Chinese Academy of Sciences\\ Beijing 100190, China\\
${}$\\
\mailsa}
\maketitle
\begin{abstract}
Given a connected graph $G$ on $n$ vertices and {a} positive integer $k\le n$, a subgraph of $G$ on $k$ vertices is {called} a $k$-subgraph in $G$. We design combinatorial approximation algorithms for finding a connected $k$-subgraph in $G$ such that its density is at least a factor $\Omega(\max\{n^{-2/5},k^2/n^2\})$ of the density of the densest $k$-subgraph in $G$ (which is not necessarily connected). These particularly provide the first non-trivial approximations for the densest connected $k$-subgraph problem on general graphs.
\end{abstract}
\noindent{\bf Keywords:} {Densest $k$-subgraphs, Connectivity, Combinatorial approximation algorithms
\newcounter{my}
\newenvironment{mylabel}
{
\begin{list}{(\roman{my})}{
\setlength{\parsep}{-0mm}
\setlength{\labelwidth}{8mm}
\setlength{\leftmargin}{8mm}
\usecounter{my}}
}{\end{list}}
\newcounter{my2}
\newenvironment{mylabel2}
{
\begin{list}{(\alph{my2})}{
\setlength{\parsep}{-1mm} \setlength{\labelwidth}{12mm}
\setlength{\leftmargin}{14mm}
\usecounter{my2}}
}{\end{list}}
\newcounter{my3}
\newenvironment{mylabel3}
{
\begin{list}{(\alph{my3})}{
\setlength{\parsep}{-1mm}
\setlength{\labelwidth}{8mm}
\setlength{\leftmargin}{10mm}
\usecounter{my3}}
}{\end{list}}
\section{Introduction}
Let $G=(V,E)$ be a connected simple undirected graph with $n$ vertices, {$m$ edges}, and nonnegative edge weights. The ({\em weighted}) {\em density} of $G $ is defined as its average (weighted) degree. Let $k\le n$ be a positive integer.
A subgraph of $G$ is called a {\em $k$-subgraph} if it has exactly $k$ vertices. The \emph{densest $k$-subgraph problem} (D$k$SP) is to find a $k$-subgraph of $G$ that has the maximum density, equivalently, a maximum number of edges. If the $k$-subgraph requires to be connected, then the problem is referred as to the \emph{densest connected $k$-subgraph problem} (DC$k$SP). Both D$k$SP and DC$k$SP have their weighted generalizations, denoted respectively as H$k$SP and HC$k$SP, which ask for a heaviest (connected) $k$-subgraph, i.e., a (connected) $k$-subgraph with a maximum total edge weight. Identifying $k$-subgraphs with high
{densities} is a useful primitive, which arises in diverse applications -- from social networks, to protein interaction
graphs, to the world wide web, etc. While dense subgraphs can give valuable information about
{interactions} in these networks, the additional connectivity requirement turns out to be natural in \bluecomment{various scenarios.
One of typical examples is searching for a large community. If most vertices belong to a dense connected subnetwork, only a few selected inter-hub links are needed to have a short average distance between any two arbitrary vertices in the entire network. Commercial airlines employ this hub-based routing scheme \cite{lrja10}.}
\paragraph{Related work.} An easy {reduction} from the maximum clique problem shows that D$k$SP, DC$k$SP and their weighted generalizations are all NP-hard in general. {The NP-hardness remains even for some very restricted graph classes
such as chordal graphs, triangle-free graphs, comparability graphs \cite{cp84} and bipartite graphs of
maximum degree three \cite{fs97}.
}
Most literature on finding dense subgraphs focus on the versions without requiring subgraphs to be connected. For D$k$SP and its generalization H$k$SP, narrowing the large gap between the lower and upper bounds on the approximabilty is an important open problem.
On the negative side, the decision problem version of D$k$SP, in which one is asked if there is a $k$-subgraph with more than $h$ edges, is NP-complete even if $h$ is restricted by \redcomment{$h\le k^{1+\varepsilon}$ \cite{ai95}.}
Feige \cite{f02} showed that computing a $(1+\varepsilon)$-approximation for D$k$SP is at least as hard as refuting random 3-SAT clauses for some $\varepsilon>0$.
Khot \cite{k06} showed that there does not exist any polynomial time approximation scheme (PTAS) for D$k$SP assuming
NP does not have randomized algorithms that run in sub-exponential time. Recently, constant factor approximations in polynomial
time for D$k$SP have been ruled out by Raghavendra and Steurel \cite{rs10} under Unique Games with Small
Set Expansion conjecture, and by Alon et al. \cite{aammw11} under certain ``average case'' hardness assumptions. On the positive side, considerable efforts have been devoted to finding good quality approximations for H$k$SP. Improving the $O(n^{0.3885})$-approximation of Kortsarz and Peleg \cite{kp93}, Feige et al. \cite{fkp01} proposed a combinatorial algorithm with approximation ratio $O(n^{\delta})$ for some $\delta<1/3$. The latest algorithm of Bhaskara et al. \cite{bccfv10} provides an $O(n^{1/4+\varepsilon})$-approximation in $n^{O(1/\varepsilon)}$ time. If allowed to run for $n^{O(\log n)}$ time, their algorithm guarantees an approximation ratio of $O(n^{1/4})$. The $O(n/k)$-approximation algorithm by Asahiro et al. \cite{aitt00} is remarkable for its simple greedy removal method. {Linear and semidefinite programming (SDP) relaxation approaches have been adopted in \cite{fl01,hyz02,sw98} to design randomized rounding {algorithms}, where Feige and Langberg \cite{fl01} obtained an approximation ratio somewhat better than $n/k$, while the algorithms of Srivastav and Wolf \cite{sw98} and Han et al. \cite{hyz02} outperform this ratio for a range of values $k=\Theta(n)$.} \bluecomment{On the other hand, the SDP relaxation methods have a limit of $n^{\Omega(1)}$ for D$k$SP as shown by Feige and Seltser \cite{fs97} and Bhaskara et al.~\cite{bcvgz12}.}
For some special \bluecomment{cases in terms of graph classes, values of $k$ and optimal objective values,}
better approximations have been obtained for D$k$SP and H$k$SP. Arora et al. \cite{akk95} gave a PTAS for the restricted D$k$SP where $m=\Omega(n^2)$ and $k=\Omega(n)$, or each vertex of $G$ has degree $\Omega(n)$. Kortsarz and Peleg \cite{kp93} approximated D$k$SP with ratio $O((n/k)^{2/3})$ when the number of edges in the optimal solution is larger than \redcomment{$2\sqrt{k^5/n}$.}
\mbox{Demaine et al. \cite{dhk05}} developed a 2-approximation algorithm for D$k$SP on $H$-minor-free graphs, where $H$ is any given fixed undirected graph. Chen et al. \cite{cfl11} showed that D$k$SP on a large family of intersection graphs, including chordal graphs, circular-arc graphs and claw-free graphs, admits constant factor approximations. Several PTAS have been designed for D$k$SP on unit disk graphs \cite{cfl11}, interval graphs \cite{n11}, and a subclass of chordal graphs \cite{lmpz07}.
The work on approximating densest/heaviest connected $k$-subgraphs are relatively very limited. To the best of our knowledge, the existing polynomial time algorithms deal only with special graphical topologies, including: (a) 4-approximation \cite{rrt94} and 2-approximation \cite{hrt97} for the metric H$k$SP and HC$k$SP, where the underlying graph $G$ is complete, and the connectivity is trivial; (b) exact algorithms for H$k$SP and HC$k$SP on trees \cite{cp84}, for D$k$SP and DC$k$SP on $h$-trees, cographs and split graphs \cite{cp84}, and for
DC$k$SP on interval graphs {whose clique graphs are simple paths \cite{lmz05}.}
Among the well-known relaxations of D$k$SP and H$k$SP is the problem of finding a (connected) subgraph (without any cardinality constraint) of maximum weighted density. It is strongly polynomial time solvable using max-flow based techniques \cite{g84,l76}. Andersen and Chellapilla \cite{ac09} and Khuller and Saha \cite{ks09} studied two relaxed variants of H$k$SP for finding a weighted densest subgraph with at least or at most $k$ vertices. The former variant was shown to be NP-hard even in {the} unweighted case, and admit 2-approximations in {the} weighted setting. The approximation of the latter variant was proved to be as hard as that of D$k$SP/H$k$SP up to a constant factor.
\paragraph{Our results.} Given the interest in finding densest/heaviest connected $k$-subgraphs from both
the theoretical and practical point of view,
a better understanding of the {problems} is
an important challenge for the field. In this paper, we design {$O(mn\log n)$} time combinatorial approximation algorithms for finding a connected $k$-subgraph of $G$ whose density (resp. weighted density) is at least a factor $\Omega(\max\{n^{-2/5},k^2/n^2\})$ (resp. $\Omega(\max\{n^{-2/3},k^2/n^2\})$) {of} the density (resp. weighted density) of the densest (resp. heaviest) $k$-subgraph of $G$ which is not necessarily connected. These particularly provide the first non-trivial approximations for the densest/heaviest connected $k$-subgraph problem on general graphs: $O(\min\{n^{2/5},n^2/k^2\})$ for DC$k$SP and $O(\min\{n^{2/3},n^2/k^2\})$ for HC$k$SP.
{To evaluate the quality of our algorithms' performance guarantees $O(n^{2/5})$ and $O(n^{2/3})$, which are compared with the optimums of D$k$SP and H$k$SP, we investigate the maximum ratio $\Lambda$ (resp. $\Lambda_w$), over all graphs $G$ (resp. over all graphs $G$ and all nonnegative edge weights), between the maximum density (resp. weighted density) of {\em all} $k$-subgraphs and that of {\em all connected} $k$-subgraphs in $G$. The following examples show $\Lambda\ge n^{1/3}/3$ and $\Lambda_w\ge n^{1/2}/2$.}
\begin{example}\label{eg1}
(a) The graph $G$ is formed from $\ell$ vertex-disjoint $\ell$-cliques $L_1,\ldots,L_\ell$ by adding, for each $i=1,\ldots,\ell-1$, a path $P_i$ of length $\ell^2+1$ {to connect $L_i$ and $L_{i+1}$,} where $P_i$ intersects all the $\ell$ cliques only at a vertex in $L_i$ and a vertex in $L_{i+1}$.
Let $k=\ell^2$. Note that $G$ has $n=\ell^2+\ell^2(\ell-1)=\ell^3$ vertices. {The unique densest $k$-subgraph of $G$ is the disjoint union of $L_1,\ldots, L_{\ell}$ and has density $\ell-1$. One of densest connected $k$-subgraphs of $G$ is induced by the $\ell$ vertices in $L_1$ and certain $\ell^2-\ell$ vertices in $P_1$, and has density $(\ell(\ell-1)+2(\ell^2-\ell))/\ell^2$. Hence $\Lambda\ge \ell^2/(\ell+2\ell)= n^{1/3}/3$.}
(b)
The graph $G$ {is a tree} formed from a star on $\ell+1$ vertices by dividing each edge into a path of length $\ell+1$. All pendant edges have weight 1 and other edges have weight 0. Let $k=2\ell$. Note that $G$ has $n=\ell^2+1$ vertices. The unique heaviest $k$-subgraph of $G$ is induced by the $\ell$ pendant edges of $G$, and has weighted density $1$. Every heaviest connected $k$-subgraph {of $G$ is a path containing exactly one pendant edge of $G$, and has weighted density $1/\ell$. Hence $\Lambda_w\ge\ell\ge n^{1/2}/2$.}\end{example}
\medskip
{The remainder of this paper is organized as follows. Section 2 gives notations, definitions and basic properties necessary for our discussion. Section 3 is devoted to designing approximation algorithms for finding connected dense $k$-subgraphs. Section 4 discusses extension to the weighted case, and future research directions.}
\section{Preliminaries}
Graphs studied in this paper are simple and undirected. {For any graph $G'=(V',E')$ and any vertex $v\in V'$, we use $d_{G'}(v)$ to denote $v$'s degree in $G'$.
The \mbox{{\em density}} ${\sigma}(G')$ of $G' $ refers to its average degree, i.e. ${\sigma}(G')=\sum_{v\in V'}d_{G'}(v)/|V'|=2|E'|/|V'|$.} {Following convention, we define $|G'|=|V'|$. By a {\em component} of $G'$ we mean a maximal connected subgraph of $G'$.}
Throughout let $G=(V,E)$ be a connected graph on $n$ vertices and $m$ edges, and let $k\in[3,n]$ be an integer. Our goal is to find a connected $k$-subgraph $C$ of $G$ such that its density $\sigma(C)$ is as large as possible.
Let $\sigma^*(G)$ and ${\sigma}_k^*(G)$ denote the maximum densities of a subgraph and a $k$-subgraph of $G$, respectively, {where the subgraphs are not necessarily connected}. It is clear that
\begin{equation}
{\sigma}^*(G)\ge {\sigma}_k^*(G)\text{ and }n-1\ge{\sigma}(G)\ge k\cdot {\sigma}_k^*(G)/n.\label{bound1} \end{equation}
Let $S$ be a subset of $V$ or a subgraph of $G$. We use $G[S]$ to denote the subgraph of $G$ induced by the vertices in $S$, and use $G\setminus S$ to denote the graph obtained from $G$ by removing all vertices in $S$ and their incident edges. If $S$ consists of {a single} vertex $v$, we write $G\setminus v$ instead of $G\setminus\{v\}$.
\begin{lemma} \label{lem1}
${\sigma}_k^*(G)<\sigma_{k-1}^*(G)+2$ and ${\sigma}_k^*(G)\le3\cdot\sigma_{k-1}^*(G)$.
\end{lemma}
\begin{proof}The first inequality in the lemma implies the second {since $\sigma_{k-1}^*(G)\ge1$}.
To prove ${\sigma}_k^*(G)<\sigma_{k-1}^*(G)+2$, consider a densest $k$-subgraph $H$ of $G$, and $v\in V(H)$. Then $d_H(v)\le k-1$, and
\begin{center}$\sigma_{k-1}^*(G)\ge\sigma(H\setminus v)=\frac{k\cdot\sigma(H)-2d_H(v)}{k-1}>\sigma(H)-\frac{2(k-1)}{k-1}={\sigma}_k^*(G)-2 $,\end{center}
establishing the lemma. \end{proof}
The vertices whose removals increase \bluecomment{the density of the graph} play an important role in our algorithm design.
\begin{definition}\label{remove}
A vertex $v\in V$ is called {\em removable} in $G$ if $\sigma(G\setminus v)>\sigma(G)$.
\end{definition}
Since $\sigma(G\setminus v)=2(|E|-d_G(v))/(|V|-1)$, the following is straightforward. It also provides an efficient way to identify removable vertices.
\begin{lemma}\label{del}
{A vertex $v\in V$ is removable in $G$} if and only if $d_G(v)\!<\!{\sigma}(G)/2$.~
\end{lemma}
\begin{lemma}\label{add}
Let $G_1$ be a connected $k$-subgraph of $G$. For any connected subgraph $G_2$ of $G_1$, it holds that $\sigma(G_1)\geq \sigma(G_2)/\sqrt{k}$.
\end{lemma}
\begin{proof} Suppose that $G_2$ is a $k_2$-subgraph of $G$ with $m_2$ edges. By {the definition of density}, $\sigma(G_2)\le k_2-1$. The connectivity of $G_1$ implies $|E(G_1)|\ge|E(G_2)|+|V(G_1\setminus G_2)|$, and
\[ \sigma(G_1)\ge \frac{2(m_2+k-k_2)}{k}=\frac{k_2\cdot\sigma(G_2)+2(k-k_2)}{k}.
\]
In case of $k_2\ge\sqrt{k}$, we have {$\sigma(G_1)\ge k_2\cdot \sigma(G_2)/k\ge \sigma(G_2)/\sqrt{k}$. In case of $k_2<\sqrt{k}$, since $k\ge3$, it follows that $G_1$ has no isolated vertices, and $\sigma(G_1)\ge1> k_2/\sqrt{k} > \sigma(G_2)/\sqrt{k}$.}
\end{proof}
For a cut-vertex $v$ of $G$, we use $G_v$ to denote a densest component of $G\setminus v$, and use $G_{v+}$ to denote the {connected} subgraph of $G$ induced by $V(G_v)\cup\{v\}$. {Note that} $G\setminus G_v$ is a connected subgraph of $G$.
\section{Algorithms}
We design an $O(n^2/k^2)$-approximation algorithm (in Section 3.1) and further an $O(n^{2/5})$-approximation algorithm (in Section 3.2) for D$k$SP that always finds a connected $k$-subgraph of $G$. For ease of description we assume $k$ is even. The case of odd $k$ can be treated similarly. Alternatively, if $k$ is odd, we can first find a connected $(k-1)$-subgraph $G_1$ satisfying $\sigma_{k-1}^*(G)/\sigma(G_1)\le O(\alpha)$, where $\alpha\in\{n^2/k^2,n^{2/5}\}$; it follows from Lemma \ref{lem1} that $\sigma^*_{k}(G)/\sigma(G_1)\le O(\alpha)$. Then we attach an appropriate vertex to $G_1$, making a connected $k$-subgraph $G_2$ with {density} $\sigma(G_2)\ge \frac{k-1}k\sigma(G_1)\ge\frac23\sigma(G_1)$. This guarantees that the approximation ratio is still ${\sigma}_k^*(G)/\sigma(G_2)\le O(\alpha)$.
\subsection{$O(n^2/k^2)$-approximation}\label{k/n}
We first give an outline of our algorithm (see Algorithm \ref{alg1}) for finding a connected $k$-subgraph $C$ of $G$ with density {$\sigma(C)\ge \Omega(k^2/n^2)\cdot\sigma^*_k(G)$} (see Theorem \ref{the1}).
\paragraph{Outline.} We start with a connected graph $G'\leftarrow G$ and repeatedly delete removable vertices from $G'$ {to increase its density without destroying its connectivity.}
\begin{itemize}
\item If we can reach $G'$ with $|G'|=k$ in this way, we output $C$ as the resulting~$G'$.
\vspace{-2mm}\item If we can find a removable cut-vertex $r$ in $G'$ such that $|G'_r|\ge k$, then we recurse with $G'\leftarrow G'_r$.
\vspace{-2mm}\item If we stop at a $G'$ without any removable vertices, then we construct $C$ from an arbitrary connected $(k/2)$-subgraph by greedily attaching $k/2$ more vertices (see Procedure \ref{pro1}).
\vspace{-2mm}\item If we are in none of the above three cases,
we find a connected subgraph of $G'$ induced by a set $S$ of at most $ k/2$ vertices, and \bluecomment{then} expand the subgraph in two ways: (1) attaching $G'_r$ for all removable vertices $r$ of $G'$ which are contained in $S$, and (2) greedily attaching no more than $k/2$ vertices. From the resulting connected subgraphs, we choose the one that has more edges \bluecomment{(breaking ties arbitrarily)}, and further expand it to be a connected $k$-subgraph (see Procedure \ref{pro2}), which is returned as the output~$C$.
\end{itemize}
\paragraph{Greedy attachment.} We describe how the greedy attaching mentioned in the above outline {proceeds}. Let $S$ and $T$ be disjoint nonempty vertex subsets (or subgraphs) of $G$. Note that $1\le|S|<n$. The set of edges of $G$ with one end in $S$ and the other in $T$ is written as $[S,T]$. For any positive integer $j\le n-|S|$, a set $S^{\star}$ of $j$ vertices in $G\setminus S$ with {\em maximum} $|[ S, S^{\star}]|$ can be found greedily by sorting the vertices in $ G \setminus S $ as $v_1,v_2,\ldots,v_{j},\ldots$ in a non-increasing order of the number of neighbors they have in $S$. For each $i=1,\ldots,j$, it can be guaranteed that $v_i$ has either a neighbor in $S$ or a neighbor in $\{v_1,\ldots,v_{i-1}\}$; in the latter case $i\ge2$. Setting $S^{\star}=\{v_1,v_2,\ldots,v_j\}$. It is easy to see that
\begin{equation}\label{max}
\begin{array}{c}|[S,S^{\star}]|\ge \frac{j}n\cdot|[S,G\setminus S]| .\end{array}
\end{equation}
Moreover, if $G[S]$ is connected, the choices of $v_i$'s guarantee that $G[S\cup S^{\star}]$ is connected. We refer to this $S^{\star}$ as a {\em $j$-attachment} of $S$ in $G$. Given $S$, finding a $j$-attachment of $S$ takes {$O(m+n\log n)$} time, which implies the following procedure runs in {$O( |E(G')|+|G'|\cdot\log |G'|)$} time.
\begin{procedure}\label{pro1}
{\em Input:} {a} connected graph $G' $ without removable vertices, where $| G' | > k$. \\
{\em Output:} a connected $k$-subgraph of $G'$, written as {\sc Prc1}($G'$).
\end{procedure}
\vspace{-0mm}\hrule
\begin{enumerate}
\vspace{-2mm} \item \vspace{0mm} $G_1=(V_1,E_1)\leftarrow$ an arbitrary connected $(k/2)$-subgraph of $G'$
\vspace{-2mm} \item \vspace{0mm} $V_1^{\star}\leftarrow$ a $(k/2)$-attachment of $V_1$ in $G'$
\vspace{-2mm} \item \vspace{0mm
Output {\sc Prc1}$(G')\leftarrow$ $G[V_1\cup V_1^{\star}]$\vspace{-1mm}
\end{enumerate}
\vspace{-1mm}\hrule
\medskip
Note that the definition of attachment guarantees that $V_1\cap V_1^{\star}=\emptyset$, $|[V_1,V_1^{\star}]|$ is maximum, and $G[V_1\cup V_1^{\star}]$ is connected.
\begin{lemma}\label{lem3}
$\sigma(\text{\sc Prc1}(G'))\ge \frac{k}{4|G'|}\cdot\sigma(G')$.
\end{lemma}
\begin{proof}
Since $G'$ has no removable vertices, we deduce from Lemma \ref{del} that every vertex of $G'$ has degree at least ${\sigma}(G')/2$. Therefore {$|[G_1, G'\setminus G_1]|\ge \frac{k}2\cdot\frac{\sigma(G')}2-2|E_1|$.} Recalling (\ref{max}),
we see that the number of edges in $\text{\sc Prc1}(G') $ is at least $ |[V_1,V_1^{\star}]|\ge{(\frac{k\cdot{\sigma}(G')}4-2|E _1 |)}\cdot\frac{k/2}{| G' |}+|E_1|\ge\frac{ k^2}{8|G'|}\cdot{\sigma}(G') $, proving the lemma.
\end{proof}
\begin{procedure}\label{pro2}
{\em Input:} {a} connected graph $G'$ with $|G'|>k$, where every removable vertex $r$ is a cut-vertex satisfying $|G'_r|<k$. \quad
{\em Output:} a connected $k$-subgraph of $G'$, written as {\sc Prc2}($G'$).
\end{procedure}
\vspace{-0mm}\hrule
\begin{enumerate}
\vspace{-2mm} \item \vspace{0mm} $H\leftarrow G'$, {$R'\leftarrow R=$ the set of removable vertices of $G'$}
\vspace{-2mm} \item \vspace{0mm} \textbf{While} $R'\neq\emptyset$ \textbf{do}
\vspace{-2mm} \item \vspace{0mm} \hspace{6mm} Take $r\in R'$
\vspace{-2mm} \item \vspace{0mm} \hspace{6mm} $H\leftarrow H\setminus V(G'_r)$, \ $R'\leftarrow R'\setminus V(G'_{r+})$ \label{delete}
\vspace{-2mm} \item \vspace{0mm} \textbf{End-While} \label{end}
\vspace{-2mm} \item \vspace{0mm} For each $v\in V(H)$, define $\theta(v)=|G'_{v+}|$ if $v\in R$, and $\theta(v)=1$ otherwise\label{size}
\vspace{-2mm} \item \vspace{0mm} Let $S$ be a {\em minimal} subset of $V(H)$ s.t. \!$H[S]$ is connected \& $\sum_{v\in S}\theta(v) \!\ge \! \frac{k}2$ \label{h1}
\vspace{-2mm} \item \vspace{0mm}
Let $S^{\ast}$ be a $\min\{k/2,|H\setminus S|\}$-attachment of $S$ in $H$\label{step2}
\vspace{-2mm} \item \vspace{0mm} {$V_1\leftarrow S\cup(\cup_{r\in R\cap S}V(G'_r))$,\; $V_2\leftarrow S\cup S^{\star}$}
\vspace{-2mm} \item \vspace{0mm} Let $H'$ be one of $G'[V_1]$ and $G'[V_2]$ {whichever has more edges (break ties arbitrarily)}\label{more}
\vspace{-2mm} \item \vspace{0mm} Expand $H'$ to be a connected $k$-subgraph of $G'$\label{exp}
\vspace{-2mm} \item \vspace{0mm} Output {{\sc Prc2}$(G')\leftarrow H'$}
\end{enumerate}
\vspace{-1mm}\hrule
\medskip
Under the condition that the resulting graph is connected, the expansion in Step~\ref{exp} can be done in an arbitrary way. It is easy to see that Procedure \ref{pro2} runs in $O(|G'|\cdot|E(G')|)$ time.
\begin{lemma}\label{lem4}
At the end of the while-loop (Step \ref{end}) in Procedure \ref{pro2}, we have
\vspace{-2mm}\begin{mylabel}
\item $H$ is a connected subgraph of $G'$.
\item If $H$ contains two distinct vertices $r$ and $s$ that are removable in $G'$, then (by the condition of the procedure both $r$ and $s$ are cut-vertices of $G'$, {and moreover}) {$G'_r$ and $ G'_s$} are vertex-disjoint.
\end{mylabel}
\end{lemma}
\begin{proof} Note that in every execution of the while-loop, $r\in R'$ is a cut-vertex of $H$, and $V(H)\cap V(G'_r)$ induces a component of $H\setminus r$. Thus $H$ is connected throughout the procedure. For any two removable vertices $r,s $ of $G'$ with $|G'_r|\le|G'_s|$ and $r,s\in V(H)$, if $G'_r$ and $G'_s$ are not vertex-disjoint, then $V(G'_r)\cup\{r\}\subseteq V(G'_s)$. It follows that all vertices of $V(G'_r)\cup\{r\}$ have been removed by Step \ref{delete} when considering $s\in R'$, a contradiction.
\end{proof}
\bluecomment{Observe that for any two distinct $r,s\in R$, either $G'_{r+}$ and $G'_{s+}$ are vertex-disjoint, or $G'_{r+}$ contains $G'_{s+}$, or $G'_{s+}$ contains $G'_{r+}$. This fact, along with an inductive argument, shows that, throughout Procedure \ref{pro2}, for any $s\in R\de V(H)$, there exists at least a vertex $r\in V(H)\cap R$ such that $G'_{r+}$ contains $G'_{s+}$, implying that $(U_{r\in R\cap V(H)}V(G_{r+}))\cup(V(H)\de R)=V(G')$ holds always. By Lemma \ref{lem4}(ii), in Step \ref{h1}, we see that $V(G')$ is the disjoint union of $V(G_{r+})$, $r\in R\cap V(H)$ and $V(H)\de R$, giving $\sum_{v\in V( H)}\theta(v)=|G'|>k$. Hence, the connectivity of $H$ (Lemma \ref{lem4}(i)) implies that the set $S$ at Step \ref{h1} does exist.}
\bluecomment{Take $u\in S$ such that $u$ is not a cut-vertex of $H$. If $|S|\ge(k/2)+1$, then we have $\sum_{v\in S\de\{u\}}\theta(v)\ge|S\de\{u\}|\ge k/2$, a contradiction to the minimality of $S$. Hence
\begin{equation*}|S|\le k/2.\end{equation*}
Since Step 4 has removed from $H$ all vertices in $V(G'_r)$ for all $r\in R$, we see that $V_1$ is the disjoint union of $S$ and $\cup_{r\in R\cap S}V(G'_r)$ Recall that $|G'_r|<k$ for all $r\in R\cap S$. If $|V_1|>k$, then $|S|\ge2$, and either $\theta_u\ge k/2$ or $ \sum_{v\in S\de\{u\}}\theta(v)\ge k/2$, contradicting to the minimality of $S$. Noting that $|V_1|=\sum_{v\in S}\theta(v)$, we have} \begin{equation}\label{v1}
k/2\le|V_1|\le k.\end{equation} We deduce that the output of Procedure \ref{pro2} is indeed a connected $k$-subgraph of~$G'$.
\begin{algorithm}\label{alg1}
{\em Input:} connected graph $G=(V,E)$ with $|V|\ge k$.\\
{\em Output:} a connected $k$-subgraph of $G$, written as {\sc Alg1}$(G)$.
\end{algorithm}
\vspace{-0mm}\hrule
\begin{enumerate}
\vspace{-2mm} \item $G'\leftarrow G$
\vspace{-2mm} \item \vspace{0mm} \textbf{While} $|G'|>k$ {and $G'$ has a removable vertex $r$ that is not a cut-vertex}
\textbf{do}\label{condition}
\vspace{-2mm}\item \vspace{0mm} \hspace{4mm} $G'\leftarrow G'\setminus r$
\vspace{-2mm}\item \vspace{0mm} \textbf{End-While}\label{note} {\hfill{\small // either $|G'|=k$ or any removable vertex of $G'$ is a cut-vertex}}
\vspace{-2mm}\item \vspace{0mm} \textbf{If} $|G'|=k$ {{\bf then} output {\sc Alg1}$(G)\leftarrow G'$} \label{s1}
\vspace{-2mm} \item \vspace{0mm} {\textbf{If} $|G'|>k$ and $G'$ has no removable vertices
\hspace{4mm}{\bf then} output {\sc Alg1}$(G)\leftarrow$ {\sc Prc1}($G'$)} \label{s2}
\vspace{-2mm} \item \vspace{0mm} {\textbf{If} $|G'|>k$ and $|G'_r|< k$ for each removable vertex $r$ of $G'$
\hspace{4mm}{\bf then} output {\sc Alg1}$(G)\leftarrow$ {\sc Prc2}($G'$)} \label{s3}
\vspace{-2mm} \item \vspace{0mm} {\textbf{If} $|G'|>k$ and $|G'_r|\ge k$ for some removable vertex $r$ of $G'$
\hspace{4mm}{\bf then} output {\sc Alg1}$(G)\leftarrow$ {\sc Alg1}($G'_r$)} \label{s4}
\end{enumerate}
\vspace{-2mm}\hrule
\vspace{3mm}
In the {while-loop}, we repeatedly delete removable non-cut vertices from $G'$ until $|G'|=k$ or {$G'$ has} no removable non-cut vertex anymore. {The deletion process keeps $G'$ connected,
and its density $\sigma(G')$} increasing (cf. Definition \ref{remove}). When the {deletion process finishes, there are four possible cases, which are handled by Steps \ref{s1}, \ref{s2}, \ref{s3} and \ref{s4}, respectively.
\begin{itemize}
\vspace{-2mm}\item In case of Step \ref{s1}, the output $G'$ is clearly a connected $k$-subgraph of $G$.
\vspace{-2mm}\item In case of Step \ref{s2}, $G'$ qualifies to be an input of Procedure \ref{pro1}. With this input, Procedure \ref{pro1} returns the connected $k$-subgraph {\sc Prc1}$(G')$ of $G'$ as the algorithm's output.
\vspace{-2mm}\item In case of Step \ref{s3}, $G'$ qualifies to be an input of Procedure \ref{pro2}. With this input, Procedure \ref{pro2} returns the connected $k$-subgraph {\sc Prc2}$(G')$ of $G'$ as the algorihtm's output.
\vspace{-2mm}\item In case of Step \ref{s4}, the algorithm recurses with {smaller} input $G'_r$, which satisfies $\sigma(G'_r)\ge\sigma(G')\ge\sigma(G)$ and $k\le|G'_r|<|G'|\le|G|$.
\end{itemize}
Hence after $O(n)$ recursions, the algorithm terminates at one of Steps \ref{s1} -- \ref{s3}, and outputs a connected $k$-subgraph of $G$.}
\begin{theorem}\label{the1}
Algorithm \ref{alg1} finds in {$O(mn)$
time} a connected $k$-subgraph $C$ of $G$ such that $ {\sigma}_k^*(G)/\sigma(C)\le12n^2/k^2 $.
\end{theorem}
\begin{proof} Let $C=\text{\sc{Alg\ref{alg1}}}(G)$ be the output connected $k$-subgraph of $G$.
If $C$ is output at Step \ref{s1}, then its density is ${\sigma}(C)\ge {\sigma}(G)\ge (k/n)\cdot{\sigma}_k^*(G)$, {where the last inequality is by (\ref{bound1}).} If $C$ is output by Procedure \ref{pro1} at Step \ref{s2}, then from Lemma \ref{lem3} we know its density is at least $ \frac{k}{4|G'|}\cdot\sigma(G')\ge \frac{k}{4n}\cdot\sigma(G)\ge \frac{k^2}{4n^2}\cdot{\sigma}_k^*(G)$.
Now we are only left with the case that $C =\text{\sc Prc2}(G')$ is output by Procedure~\ref{pro2} at Step \ref{s3} of Algorithm~\ref{alg1}. Let $R$ denote the set of removable vertices of $G'$. For every $r\in R$, we see that $r$ is a cut-vertex of $G'$ (cf. the note at Step \ref{note} of the algorithm), and ${\sigma}(G'_r)\ge{\sigma}(G'\setminus r)> {\sigma}(G')$, where the first inequality is from the definition of $G'_r$ (it is the densest component of $G'\setminus r$), and the second inequality is due to the removability of {$r$.} Thus \[{\sigma}(G'_{r+})> \sigma(G'_r)\cdot|G'_r|/(|G'_r|+1)\ge {\sigma}(G')/2\;\text{ for every }r\in R.\] Using the notations in Procedure \ref{pro2}, we note that each vertex of $S\setminus R$ is non-removable in $G'$, and therefore has degree at least $\sigma(G')/2$ in $G'$ by Lemma \ref{del}. Since $V_1= S\cup(\cup_{r\in R\cap S}V(G'_r))=(S\setminus R)\cup(\cup_{r\in S\cap R}V(G'_{r+}))$ contains at least $k/2$ vertices {(recall (\ref{v1}))}, it follows that $G'$ contains at least {$(\frac{k}2\cdot\frac{\sigma(G')}2)/2\ge \frac{k}8\cdot\sigma(G)\ge \frac{k^2}{8n}\cdot{\sigma}_k^*(G)$} edges each with at least one end in $V_1$.
If there are at least {$ \frac{k^2}{24n}\cdot{\sigma}_k^*(G)$ edges with both ends in $V_1$, then by Step~\ref{more} of Procedure \ref{pro2} we have $|E(C)|\ge \frac{k^2}{24n}\cdot{\sigma}_k^*(G)$ and ${\sigma}(C)=2|E(C)|/k\ge \frac{k}{12n}\cdot{\sigma}_k^*(G) \ge \frac{k^2}{12n^2}\cdot{\sigma}_k^*(G)$}. It remains to consider the case where $G'$ contains at least {$\frac{k^2}{12n}\cdot{\sigma}_k^*(G)$} edges between $V_1$ and $G'\setminus V_1$. All these edges are between $S$ and $G'\setminus V_1=H\setminus S$, since each edge incident with any vertex in $G'_r$ ($r\in R$) must have both ends in $V_1$.
So, by the definition of $S^{\star}$ at Step \ref{step2} of Procedure \ref{pro2}, we deduce from (\ref{max}) that there are at least {a number $|[S,S^{\star}]|\ge \frac{k/2}n\cdot|[S,H\setminus S]|\ge\frac{ k^3}{24n^2}\cdot{{\sigma}_k^*}(G)$ of} edges in the subgraph of $G'$ induced by $V_2=S\cup S^{\star}$. Hence {$ \sigma(C)\ge 2|[S,S^{\star}]|/k\ge \frac{k^2}{12n^2}\cdot{\sigma}_k^*(G) $}, \bluecomment{justifying the performance of the algorithm.}
Algorithm \ref{alg1} runs Procedure \ref{pro1} or Procedure \ref{pro2} at most once, which takes $O(mn)$ time. {At least one of Procedures \ref{pro1} and \ref{pro2} has never been called by the algorithm}. Using appropriate data structures {and $O(n^2)$ time preprocessing, we construct a list $L$ of removable vertices in $G'$ (cf. Lemma \ref{del}). It takes $O(m)$ time for Step 2 to determine whether a removable vertex $r\in L$ is a cut-vertex of $G'$, and obtain $G'_r$ if it is. If $r$ is not a cut-vertex, then we remove $r$ from $G'$, and update $G'$ and $L$ in $O(n)$ time. If $r$ is a cut-vertex with $|G'_r|<k$, then $r$ remains a cut-vertex of $G'$ in the subsequent process (note $|G'|\ge k$ holds always) unless it is removed from the graph by certain recursion at Step \ref{s4}; so the subsequent while-loops {will} never consider it. If $r$ is a cut-vertex with $|G'_r|\ge k$, then we recurse on $G'_r$, and update $G'\rightarrow G'_r$ and $L$ in {$O(| G'_r|)=O(n)$} time, throwing away $G'\setminus G'_r$ which contains $r$. Overall, the algorithm runs in $O(mn)$ time.}
\end{proof}
\subsection{$O(n^{2/5})$-approximation}\label{2/5}
In this subsection we design algorithms for finding connected $k$-subgraphs of $G$ that jointly provide an $O(n^{2/5})$-approximation to D$k$SP. Among the outputs of all these algorithms (with input $G$), we select the densest one, denoted as $C$. Then it can be guaranteed that {$\sigma^*_k(G)/\sigma(C)\le O(n^{2/5})$}. In view of the $O(n^2/k^2)$-approximation of Algorithm \ref{alg1}, we may focus on the case of $k< n^{4/5}$. (Note that $ n^2/k^2 \le n^{2/5}$ if $k\ge n^{4/5}$.)
Let $D$ be a densest connected subgraph of $G$, which is computable in time $O(mn\log(n^2/m))$ \cite{g84,l76} (because every component of a densest subgraph of $G$ is also a densest subgraph of $G$). Thus
\begin{center}${\sigma}(D)={\sigma}^*(G)\ge\sigma^*_k(G).$\end{center}
Moreover, the maximality of $\sigma(D)$ implies that $D$ has no removable vertices.
\begin{algorithm}\label{pro3}
{\em Input:} connected graph $G$ along with its densest connected subgraph $D$.\\
{\em Output:} a connected $k$-subgraph of $G$, denoted as {\sc Alg\ref{pro3}}($G$).
\end{algorithm}
\vspace{-0mm}\hrule
\begin{enumerate}
\vspace{-2mm} \item \textbf{If} $|D|\le k$ \textbf{then} Expand $D$ to be a connected $k$-subgraph $H$ of $G$
\vspace{-1mm}\hspace{25mm} Output {\sc Alg\ref{pro3}}$(G)\leftarrow H$
\vspace{-2mm} \item \hspace{3mm} \textbf{Else} Output {\sc Alg\ref{pro3}}$(G)\leftarrow $ {\sc Prc\ref{pro1}}($D$)
\end{enumerate}
\vspace{-2mm}\hrule
\medskip
\begin{lemma}\label{per3}
If $k< n^{4/5}$, then
{${\sigma}(\text{\sc Alg\ref{pro3}}(G))\ge \min\{k/(4n),n^{-2/5}\}\cdot{\sigma}^*(G)$.}
\end{lemma}
\begin{proof}In case of $|D|\le k$, by Lemma \ref{add}, it follows from ${\sigma}^*(G)\ge {\sigma}_k^*(G)$ that the density of the output subgraph {${\sigma}(H)\ge \sigma(D)/\sqrt{k}={\sigma}^*(G)/\sqrt k
$. Since $k\le n^{4/5}$, we see that $\sigma(H)\ge n^{-2/5}\cdot\sigma^*(G)$.}
In case of $|D|>k$, we deduce from Lemma \ref{lem3} that the connected $k$-subgraph {\sc Alg\ref{pro3}}($G$)={\sc Prc\ref{pro1}}($D$) of $D$ has density at least {$\frac{k}{4|D|}\cdot \sigma(D) \ge \frac{k}{4n} \cdot{\sigma}^*(G)$.} \end{proof}
Our next algorithm is simply an expansion of Procedure 2 by Feige et al.~\cite{fkp01}.
{Let $V_h$ be a set of $k/2$ vertices of highest degrees in $G$, and let $d_h=\frac2k\sum_{v\in V_h}\!d_G(v)$ denote the average degree of {the} vertices in $V_h$.}
\begin{algorithm}\label{pro4}
{\em Input:} connected graph $G$ with $|G|\ge k$. \\
{\em Output:} a connected $k$-subgraph of $G$, denoted as {\sc Alg\ref{pro4}}($G$)..
\end{algorithm}
\vspace{-0mm}\hrule
\begin{enumerate}
\vspace{-1mm}\item \vspace{0mm} $V_h^{\star}\leftarrow$ a $(k/2)$-attachment of $V_h$ in $G$
\vspace{-2mm}\item \vspace{0mm} {$H\leftarrow$ a densest component of $G[V_h\cup V_h^{\star}]$
\vspace{-2mm}\item \vspace{0mm} {Output {\sc Alg\ref{pro4}}$(G)\leftarrow $ a $k$-connected subgraph of $G$ that is expanded from $H$}
\end{enumerate}
\vspace{-1mm}\hrule
\medskip
In the above algorithm, the subgraph $G[V_h\cup V_h^{\star}]$ is exactly the output of Procedure 2 in~\cite{fkp01}, for which it has been shown (cf, Lemma 3.2 of \cite{fkp01}) that
\[\bar\sigma:=\sigma(G[V_h\cup V_h^{\star}])\ge kd_h/(2n).\] Together with Lemma \ref{add}, we have the following result.
\begin{lemma}\label{expand}
{${\sigma}(\text{\sc Alg\ref{pro4}}(G))\geq \frac{\bar\sigma}{\sqrt k}\geq\frac{\sqrt{k}}{2n}\cdot d_h $.}
\end{lemma}
\begin{proof} It follows from Lemma \ref{add} that ${\sigma}(\text{\sc Alg\ref{pro4}}(G))\ge\sigma(H)/\sqrt{k}\ge\bar\sigma/\sqrt{k}$.
\end{proof}
Our last algorithm is a slight modification of Procedure 3 in \cite{fkp01}, where we link things up via a ``hub'' vertex.
For vertices $u,v$ of $G$, let $W(u,v)$ denote the number of walks of length $2$ from $u$ to $v$ in $G$.
\begin{algorithm}\label{pro5}
{\em Input:} connected graph $G=(V,E)$ with $|G|\ge k$. \\
{\em Output:} a connected $k$-subgraph of $G$, denoted as {\sc Alg\ref{pro5}}($G$).
\end{algorithm}
\vspace{-0mm}\hrule
\begin{enumerate}
\vspace{-1mm}
\item \vspace{0mm} $G_\ell\leftarrow G[ V\setminus V_h]$
\vspace{-2mm}\item \vspace{0mm} Compute $W(u,v)$ for all pairs of vertices {$u,v$} in $G_\ell$.
\vspace{-2mm} \item \vspace{0mm} For every $v\in V\setminus V_h$, construct a {connected $k$-subgraph {$C^v$} of $G$} as follows:
\begin{itemize}
\item[-] \vspace{-2mm} Sort the vertices {$u\in V\setminus V_h\setminus\{v\}$} with positive $W(v,u)$ as $v_1,v_2,\ldots,v_t$ such that $W(v,v_1)\ge W(v, v_2)\ge \cdots\ge W(v,v_t)>0$
\item[-] \vspace{-1mm} $P^v\leftarrow\{v_1,\ldots, v_{\min\{t,k/2-1\}}\}$
\item[-] \vspace{-1mm} {$B^v\leftarrow $ a set of $\min\{d_{G_\ell}(v),k/2\}$ neighbors of $v$ in $G_\ell$ such that {the number of edges between $B^v$ and $P^v$ is maximized}.}
\item[-] \vspace{-1mm} $C^v\leftarrow$ {the component of $ G_\ell[\{v\}\cup B^v\cup P^v]$ that contains $v$}
\item[-]\vspace{-1mm} Expand $C^v$ to be a connected $k$-subgraph of $G$
\end{itemize}
\item \vspace{-2mm} Output {\sc Alg\ref{pro5}}$(G)\leftarrow$ the densest $C^v$ for $v\in V\setminus V_h$
\end{enumerate}
\vspace{-1mm}\hrule
\medskip
In the above algorithm, $B^v$ can be {found in $O(m+n\log n)$} time, and $v$ is the ``hub'' vertex ensuring that $C^v$ is connected.
Hence the algorithm is correct, and runs in $O(mn+n^2\log n)$ time, {where Step 2 finishes in $O(n^2\log n)$ time}. The key point here is that $C^v$ contains all edges between $B^v$ and $P^v$, {where $B^v$ and $P^v$ are not necessarily disjoint.}
Using a similar analysis to that in \cite{fkp01}, we obtain the following.
\begin{lemma}\label{per5}
If {$k\!\le \!\frac23n$, then} $\sigma( \text{\sc Alg\ref{pro5}}(G))\ge \frac{({\sigma}^*_{k}(G)-2\bar\sigma)^2}{2\max\{k, 2d_h\}}\!\cdot\!\frac{k-2}{k}\geq \frac{({\sigma}^*_{k}(G)-2\bar\sigma)^2}{6\max\{k, 2d_h\}}$.~
\end{lemma}
\begin{proof} From Lemma 3.3 of \cite{fkp01} we know that $G_\ell$ contains a $k$-subgraph, denoted as $H$, with average degree at least $\sigma_k^*(G)-2\bar\sigma$. Note that the number of length-2 walks within $H$ is at least $k(\sigma_k^*(G)-2\bar\sigma)^2$. This is because each $v\in V( H)$ contributes $(d_H(v))^2$ to this number, and $\sum_{v\in V(H)}(d_H(v))^2\geq k(\sigma_k^*(G)-2\bar\sigma)^2$ by convexity. It follows that there is a vertex $v\in V(H)$ which is the endpoint of at least {a number $(\sigma_k^*(G)-2\bar\sigma)^2$ of} length-2 walks in $H$. By the construction of $P^v$, there are at least $(\sigma_k^*(G)-2\bar\sigma)^2\cdot\frac{(k/2-1)}{k}$ walks of length 2 between this {vertex} $v$ and vertices in {$P^v$}. Therefore, the number of edges between $B^v$ and $P^v$ is at least $\frac{(\sigma_k^*(G)-2\bar\sigma)^2(k-2)}{2k}$ if $d_{G_\ell}(v)\le k/2$, and at least $\frac{(\sigma_k^*(G)-2\bar\sigma)^2 (k-2)}{2k}\cdot\frac{k/2}{d_{G_\ell}(v)}$ edges {otherwise}. Since we do not require $P^v$ and $B^v$ to be disjoint, each edge may have been counted twice. Notice from the definition of $d_h$ that $d_{G_{\ell}}(v)\le d_G(v)\le d_h$. Since $C^v$ contains all edges between $B^v$ and $P^v$, it contains at least $\min\{\frac{(\sigma_k^*(G)-2\bar\sigma)^2(k-2)}{4k}, \frac{(\sigma_k^*(G)-2\bar\sigma)^2(k-2)}{8d_h}\}$ edges.
This guarantees $\sigma( \text{\sc Alg\ref{pro5}}(G))\ge \frac{({\sigma}^*_{k}(G)-2\bar\sigma)^2}{2\max\{k, 2d_h\}}\cdot\frac{k-2}{k}$. {Since $k\ge3$, the lemma follows.}
\end{proof}
\iffalse
\begin{proof}
\redcomment{It has been shown in \cite{fkp01} that
$\max\{\sigma(C^v\setminus v): {v\in V(G_\ell)}\}\ge \frac{({\sigma}_{k-1}(G)-(k-1)d_h/n)^2}{2\max\{k-1, 2d_h\}}$. Therefore
\[\sigma( \text{\sc Alg\ref{pro5}}(G))\ge\frac{({\sigma}_{k-1}(G)-(k-1)d_h/n)^2}{2\max\{k-1, 2d_h\}}\cdot\frac{k-1}{k}\ge \frac{({\sigma}_{k-1}(G)-kd_h/n)^2}{4\max\{k, 2d_h\}}
\] as desired}
\end{proof}
\fi
We are now ready to prove that the four algorithms given above jointly guarantees an $O(n^{2/5})$-approximation.
\begin{theorem}
A connected $k$-subgraph $C$ of $G$ can be found in {$O(mn\log n)$} time such that $\sigma^*_k(G)/\sigma(C)\le O(n^{2/5})$.
\end{theorem}
\begin{proof}
Let $C$ be the densest connected $k$-subgraph of $G$ among the outputs of Algorithms \ref{alg1} -- \ref{pro5}. As mentioned at the beginning of Section \ref{2/5}, it suffices to consider the case of $k<n^{4/5}$. The connectivity of $C$ gives $\sigma(C)\ge1$. {Clearly, we may assume $n\ge8$, which along with $k<n^{4/5}$ implies $k\le2n/3$.} By Lemmas \ref{per3} -- \ref{per5}, we may assume that
\[{\sigma}(C)\ge \max\left\{1, \text{ }\frac{k{\sigma}^*(G)}{4n},\text{ }\frac{\bar\sigma}{\sqrt{k}},\text{ } \frac{\sqrt{k}d_h}{2n}, \text{ } \frac{({\sigma}_k(G)-2\bar\sigma)^2}{6\max\{k, 2d_h\}} \right\}.\]
If $k\ge n^{3/5}$, then ${\sigma}(C)\ge k\cdot{\sigma}^*(G)/(4n)\ge {\sigma}^*(G)/(4n^{2/5})\ge {\sigma}^*_k(G)/(4n^{2/5})$.
If $k\le n^{2/5}$, then ${\sigma}(C)\ge 1\ge {\sigma}^*_k(G)/k\ge {\sigma}^*_k(G)/n^{2/5}$. So we are only left with the case {of} $n^{2/5}\le k\le n^{3/5}$.
Since ${\sigma}(C)\ge \bar\sigma/\sqrt{k}\ge {\bar\sigma}/{n^{3/10}}\ge {\bar\sigma}/{n^{2/5}}$, we may assume $\bar\sigma<\sigma^*_k(G)/4$, and hence ${\sigma}^*_k(G)-2\bar\sigma\ge {\sigma}^*_k(G)/2$.
Next we use the geometric mean to prove the performance guarantee as claimed.
In case of $k\ge 2d_h$, since ${\sigma}^*(G)\ge {\sigma}^*_k(G)$, we have
\[{\sigma}(C)\ge\left(1\cdot\frac{k{\sigma}^*(G)}{4n}\cdot\frac{( {\sigma}^*_k(G)/2)^2}{6k}\right)^{1/3}\ge \frac{{\sigma}^*_k(G)}{5n^{2/5}},\]
In case of $k<2d_h$, we have
\[{\sigma}(C)\ge\left(1\cdot\frac{\sqrt{k}d_h}{2n}\cdot\frac{({\sigma}^*_k(G)/2)^2}{12d_h}\cdot\frac{\sqrt{k}d_h}{2n}\cdot\frac{({\sigma}^*_k(G)/2)^2}{12d_h}\right)^{1/5}\ge \frac{{\sigma}^*_k(G)}{7 n^{2/5}},\]
where the last inequality follows from the fact that $k\ge {\sigma}^*_k(G)$.
\end{proof}
\section{Conclusion}
In Section 3, we have given four strongly polynomial time algorithms that jointly guarantee an $O(\min\{n^{2/5},n^2/k^2\})$-approximation for the unweighted problem -- DC$k$SP. {The approximation ratio is compared with the maximum density of {\em all} $k$-subgraphs, and in this case no $O(n^{1/3-\varepsilon})$-approximation for any $\varepsilon>0$ can be {expected (recall $\Lambda\ge n^{ 1/3}/3$} in Example \ref{eg1}(a)).} When studying the weighted generalization -- HC$k$SP, we can extend the techniques developed in Section~\ref{k/n}, and obtain an $O(n^2/k^2)$-approximation for {the weighted case. Besides, \bluecomment{the following} simple greedy approach achieves a {$(k/2)$}-approximation}.
\begin{algorithm}\label{alg5}
{\em Input:} connected graph $G=(V,E)$ with $|G|\ge k$ and weight $w\in \mathbb Z_+^E$. \quad
{\em Output:} a connected $k$-subgraph of $G$, denoted as {\sc Alg\ref{alg5}}($G$).
\end{algorithm}
\vspace{-0mm}\hrule
\begin{enumerate}
\vspace{-2mm}
\item \vspace{0mm} For every $v\in V$, sort the neighbors of $v$ as $v_1, v_2,\ldots, v_t$ such that $w(vv_1)\ge w(vv_2)\ge \cdots\ge w(vv_t)$, where $t=\min\{d_G(v), k-1\}
\item \vspace{-2mm} $C^v\leftarrow G[ \{v,v_1,v_2,\ldots, v_t\}]$
\item \vspace{-2mm} {\bf If} $|C^v|<k$, {\bf then} expand it to be a connected $k$-subgraph
\item \vspace{-2mm} Output {\sc Alg\ref{alg5}}$(G)\leftarrow$ the heaviest $C^v$ for all $v\in V$
\end{enumerate}
\vspace{-2mm}\hrule
\vspace{2mm}
\newpage
Notice that the weighted degree of a vertex $v$ in any heaviest $k$-subgraph of $G$ is not greater than the weight of $C^v$ \bluecomment{constructed in Algorithm \ref{alg5}}. {It is easy to see that Algorithm \ref{alg5} outputs
a connected $k$-subgraph of $G$ whose weighted density is at least {$2/k$} of that of the heaviest $k$-subgraph of $G$ (which is not necessarily connected). The running time is bottlenecked by the sorting at Step 1 which takes $O(|d_G(v)|\cdot\log|d_G(v)|)$ time for each $v\in V$. Hence the algorithm runs in $O(\log n\cdot\sum_{v\in V} |d_G(v)|)=O(m\log n)$ time.}
As $\min\{n^2/k^2,k\}\le n^{2/3}$, we have the following result.
\begin{theorem}
For any connected graph $G=(V,E)$ with weight $w\in\mathbb Z_+^E$, a connected $k$-subgraph $H$ of $G$ can be found in $O(nm)$ time such that $\sigma^*_k(G,w)/\sigma(H,w)$ $\le O(\min\{n^{2/3},n^2/k^2,k\})$, where $\sigma(H,w)$ is the weighted density of $H$, and $\sigma^*_k(G,w)$ is the weighted density of a heaviest $k$-subgraph of $G$ (which is not necessarily connected).
\end{theorem}
Since the weighted density of a graph is not necessarily related to its number of edges or vertices, {a couple of} the results in the previous sections (such as Lemmas \ref{add}, \ref{expand} and \ref{per5}) do not hold for the general weighted case. Neither the techniques of extending unweighted case approximations to weighted cases in \cite{kp93,fkp01} apply to our setting due to the connectivity constraint. {An immediate question is whether an $O(n^{2/5})$-approximation algorithm exists for HC$k$SP. Note from {$\Lambda_w\ge n^{ 1/2}/2$} in Example \ref{eg1}(b) that no one can achieve an $O(n^{1/2-\varepsilon})$-approximation for any $\varepsilon>0$ if she/he compares the solution value with the maximum weighted density of {\em all} $k$-subgraphs. Among other algorithmic approaches, analyzing the properties of {densest/heaviest} {\em connected} $k$-subgraphs is an important and challenging task in obtaining improved approximation ratios for DC$k$SP and HC$k$SP.}
|
\section*{#1}}
\newcommand{\silentsubsec}[1]{\addtocontents{toc}{\SkipTocEntry}\subsection*{#1}}
\hyphenation{qua-si-ran-dom}
\hyphenation{Schutz-en-ber-ger}
\hyphenation{com-mut-at-iv-ity}
\newtheorem{theorem}{Theorem}[section]
\newtheorem{prop}[theorem]{Proposition}
\newtheorem{proc}[theorem]{Procedure}
\newtheorem{lemma}[theorem]{Lemma}
\newtheorem{lem}[theorem]{Lemma}
\newtheorem{corollary}[theorem]{Corollary}
\newtheorem{conjecture}[theorem]{Conjecture}
\newtheorem{open}[theorem]{Open problem}
\newtheorem{proposition}[theorem]{Proposition}
\theoremstyle{remark}
\newtheorem*{remark}{Remark}
\newtheorem{example}{Example}
\newtheorem{rem}{Remark}
\newcounter{counter}
\numberwithin{counter}{section}
\theoremstyle{definition}
\newtheorem{definition}[theorem]{Definition}
\def\mathop{\rm lim\,inf}\limits{\mathop{\rm lim\,inf}\limits}
\def\mbox{mod }{\mbox{mod }}
\def\K{K}
\def\simeq{\simeq}
\def\mbox{outdeg}{\mbox{outdeg}}
\def\mbox{indeg}{\mbox{indeg}}
\def\mbox{\scriptsize indeg}{\mbox{\scriptsize indeg}}
\def\mbox{\scriptsize vertex}{\mbox{\scriptsize vertex}}
\def\mbox{\scriptsize edge}{\mbox{\scriptsize edge}}
\defs{s}
\deft{t}
\defs{s}
\deft{t}
\def{\mbox{\scriptsize root}}{{\mbox{\scriptsize root}}}
\def{\scriptsize \mbox{ spans }}{{\scriptsize \mbox{ spans }}}
\def\mbox{Im}{\mbox{Im}}
\def\mbox{Id}{\mbox{Id}}
\def\mathbf{a}{\mathbf{a}}
\def\mathbf{b}{\mathbf{b}}
\def\mathbf{d}{\mathbf{d}}
\def\mathbf{m}{\mathbf{m}}
\def\mathbf{n}{\mathbf{n}}
\def\mathbf{q}{\mathbf{q}}
\def\mathbf{r}{\mathbf{r}}
\def\mathbf{s}{\mathbf{s}}
\def\mathbf{t}{\mathbf{t}}
\def\mathbf{u}{\mathbf{u}}
\def\mathbf{v}{\mathbf{v}}
\def\mathbf{w}{\mathbf{w}}
\def\mathbf{x}{\mathbf{x}}
\def\mathbf{y}{\mathbf{y}}
\def\mathbf{z}{\mathbf{z}}
\def\mathbf{0}{\mathbf{0}}
\def\mathbf{m}{\mathbf{m}}
\def\mathbf{k}{\mathbf{k}}
\def\widetilde{\widetilde}
\def\mathop{\mathrm{Var}}{\mathop{\mathrm{Var}}}
\def\Proc{\CMcal{P}}
\def\Net{{ \mathcal{N}}}
\defN{N}
\def\mathrm{Rec\,}{\mathrm{Rec\,}}
\def\mathrm{Crit\,}{\mathrm{Crit\,}}
\def\mathop{\triangleright}{\mathop{\triangleright}}
\def\acts \! \acts \,{\mathop{\hspace{1pt} \triangleright \hspace{0.5pt} \triangleright}}
\def\acts \hspace{0.5pt} \acts{\acts \! \acts \,}
\def\mathrm{End\, }{\mathrm{End\, }}
\def\mathbf{0}{\mathbf{0}}
\def\mathbf{1}{\mathbf{1}}
\def\mathcal{S}{\mathcal{S}}
\def\mathcal{SG}{\mathcal{SG}}
\defM{M}
\def1{1}
\def{\tt Rotor}{{\tt Rotor}}
\def{\tt Sand}{{\tt Sand}}
\def\mathbb{N}{\mathbb{N}}
\def\mathbb{Z}{\mathbb{Z}}
\def\mathbb{Q}{\mathbb{Q}}
\def\mathbb{R}{\mathbb{R}}
\def\mathbb{E}{\mathbb{E}}
\def\mathbb{P}{\mathbb{P}}
\def\epsilon{\epsilon}
\def\operatorname{Var}{\operatorname{Var}}
\def\operatorname{Aut}{\operatorname{Aut}}
\newcommand\norm[1]{\left\lVert#1\right\rVert}
\begin{document}
\title{The divisible sandpile at critical density}
\author[Levine, Murugan, Peres, Ugurcan]{Lionel Levine, Mathav Murugan, Yuval Peres and Baris Evren Ugurcan}
\address{Lionel Levine, Department of Mathematics, Cornell University, Ithaca, NY 14853, USA. {\tt \url{http://www.math.cornell.edu/~levine}}}
\address{Department of Mathematics, University of British Columbia and Pacific Institute for the Mathematical Sciences, Vancouver, BC V6T 1Z2, Canada. <EMAIL>}
\address{Yuval Peres, Microsoft Research, Redmond, WA 98052, USA. <EMAIL>}
\address{Baris Evren Ugurcan, Department of Mathematics, Cornell University, Ithaca, NY 14853, USA. <EMAIL>}
\date{July 28, 2015}
\keywords{divisible sandpile, stabilizability, bi-Laplacian Gaussian field}
\subjclass[2010]{
60J45,
60G15,
82C20,
82C26
}
\thanks{The first author was supported by \href{http://www.nsf.gov/awardsearch/showAward?AWD_ID=1243606}{NSF DMS-1243606} and a Sloan Fellowship.}
\begin{abstract}
The divisible sandpile starts with i.i.d.\ random variables (``masses'') at the vertices of an infinite, vertex-transitive graph, and redistributes mass by a local toppling rule in an attempt to make all masses $\leq 1$. The process stabilizes almost surely if $m<1$ and it almost surely does not stabilize if $m>1$, where $m$ is the mean mass per vertex.
The main result of this paper is that in the critical case $m=1$, if the initial masses have finite variance, then the process almost surely does not stabilize. To give quantitative estimates on a finite graph, we relate the number of topplings to a discrete bi-Laplacian Gaussian field.
\end{abstract}
\maketitle
\section{Introduction}
This paper is concerned with the dichotomy between \emph{stabilizing} and \emph{exploding} configurations in a model of mass redistribution, the \emph{divisible sandpile model}.
The main interest in this model is twofold.
First, it is a natural starting place for the analogous and more difficult dichotomy in the \emph{abelian sandpile model}. Second, the divisible sandpile itself leads to interesting questions in potential theory. For example, under what conditions must a random harmonic function be an almost sure constant? (Lemma~\ref{l.stationaryharmonic} gives some sufficient conditions.)
Both the motivation for this paper and many of the proof techniques are directly inspired by the work
of Fey, Meester and Redig \cite{FMR}.
By a \emph{graph} $G=(V,E)$ we will always mean a connected, locally finite and undirected graph with vertex set $V$ and edge set $E$. We write $x \sim y$ to mean that $(x,y) \in E$, and $\deg(x)$ for the number of $y$ such that $x \sim y$.
A \emph{divisible sandpile configuration} on $G$ is a function $s: V \to \mathbb{R}$. We refer to $s(x)$ as an amount of `mass' present at vertex $x$; a negative value of $s(x)$ can be imagined as a `hole' waiting to be filled by mass. A vertex $x \in V$ is called \emph{unstable} if $s(x)> 1$. An unstable vertex $x$ \emph{topples} by keeping mass $1$ for itself and distributing the excess $s(x) -1$ equally among its neighbors $y \sim x$.
At each discrete time step, all unstable vertices topple simultaneously. (This parallel toppling assumption is mainly for simplicity; in Section \ref{s.maindefs} we will relax it.)
The following trivial consequence of the toppling rule is worth emphasizing: if for a particular vertex $x$ the inequality $s(x) \geq 1$ holds at some time, then it holds at all later times.
Note that the entire system evolves deterministically once an initial condition $s$ is fixed.
The initial $s$ can be deterministic or random; below we will see one example of each type.
Write $u_n(x)$ for the total amount of mass emitted before time $n$ from $x$ to one of its neighbors. (By the symmetry of the toppling rule, it does not matter which neighbor.) This quantity increases with $n$, so $u_n \uparrow u$ as $n \uparrow \infty$ for a function $u : V \to [0,\infty]$. We call this function $u$ the \emph{odometer} of $s$. Note that if $u(x)=\infty$ for some $x$, then each neighbor of $x$ receives an infinite amount of mass from $x$, so $u(y) = \infty$ for all $y \sim x$. We therefore have the following dichotomy:
\begin{align*} \text{Either \; $u(x)<\infty$ for all $x \in V$,} \\ \text{or \; $u(x) = \infty$ for all $x \in V$.} \end{align*}
In the former case we say that $s$ \emph{stabilizes}, and in the latter case we say that $s$ \emph{explodes}.
The following theme repeats itself at several places: \emph{The question of whether $s$ stabilizes depends not only on $s$ itself but also on the underlying graph}. For instance, fixing a vertex $o$, we will see that the divisible sandpile
\[ s(x) = \begin{cases} 1 & x \neq o \\ 2 & x=o \end{cases} \]
stabilizes on $G$ if and only if the simple random walk on $G$ is transient (Lemma~\ref{l.diracplusone}).
Our main result treats the case of initial masses $s(x)$ that are independent and identically distributed (i.i.d.) random variables with finite variance. Write $\mathbb{E} s$ and $\mathop{\mathrm{Var}} s$ for the common mean and variance of the $s(x)$. The mean $\mathbb{E} s$ is sometimes called the \emph{density} (in the physical sense of the word, mass per unit volume). Because sites topple when their mass exceeds $1$, intuition suggests that the density should be the main determiner of whether or not $s$ stabilizes: the higher the density, the harder it is to stabilize. Indeed, we will see that $s$ stabilizes almost surely if $\mathbb{E} s <1$ (Lemma \ref{l.stable}) and explodes almost surely if $\mathbb{E} s >1$ (Lemma \ref{l.notstab}).
Our main result addresses the critical case $\mathbb{E} s =1$.
\begin{theorem} \label{main}
Let $s$ be an i.i.d.\ divisible sandpile on an infinite, vertex-transitive graph, with $\mathbb{E} s=1$ and $0 < \operatorname{Var} s < \infty$. Then $s$ almost surely does not stabilize.
\end{theorem}
Our theme that `stabilizability depends on the underlying graph' repeats again in the proof of Theorem \ref{main}.
The proof splits into three cases depending on the graph. The cases in increasing order of difficulty are
\begin{itemize}
\item recurrent (Lemma \ref{l.recurrent}). Examples: $\mathbb{Z}, \mathbb{Z}^2$.
\item transient with $\sum_{x \in V} g(o,x)^2 = \infty$ (Section \ref{s.strans}). Examples: $\mathbb{Z}^3,\mathbb{Z}^4$.
\item transient with $\sum_{x \in V} g(o,x)^2 <\infty$ (Section \ref{s.dtrans}). Examples: $\mathbb{Z}^d$ with $d\ge 5$.
\end{itemize}
Here $g$ denotes Green's function: $g(o,x)$ is the expected number of visits to $x$ by a simple random walk started at $o$. The reason for the order of difficulty is that `stabilization is harder in lower dimensions,' in a sense formalized by Theorem~\ref{p.torus} below.
\subsection{Potential theory of real-valued functions}
The \emph{graph Laplacian} $\Delta$ acts on functions $u: V \to \mathbb{R}$ by
\begin{equation} \label{e.thelaplacian} \Delta u(x) = \sum_{y \sim x} (u(y) - u(x)). \end{equation}
Using a `least action principle' (Proposition \ref{lap}), the question of whether a divisible sandpile stabilizes can be reformulated as a question in potential theory:
\begin{quote}
Given a function $s: V \to \mathbb{R}$,
does there exist a nonnegative function
$u : V \to \mathbb{R}$ such that $s + \Delta u \le 1$ pointwise?
\end{quote}
\subsection{Potential theory of integer-valued functions}
In the related \emph{abelian sandpile model},
configurations are integer-valued functions $s : V \to \mathbb{Z}$. We think of $s(x)$ as a number of particles present at $x$. A vertex $x \in V$ is \emph{unstable} if it
has at least $\deg(x)$ particles. An unstable site $x$ \emph{topples} by sending one particle to each of its $\deg(x)$ neighbors. This model also has a dichotomy between stabilizing ($u<\infty$) and exploding ($u \equiv \infty$), which can be reformulated as follows:
\begin{quote}
Given a function $s: V \to \mathbb{Z}$, does there exist a nonnegative
function $u: V \to \mathbb{Z}$ such that $s + \Delta u \le \deg - 1$ pointwise?
\end{quote}
The restriction that $u$ must be integer-valued introduces new difficulties that are not present in the divisible sandpile model. The first step in the proof of Theorem~\ref{main} is to argue that if $\mathbb{E} s =1$ and $s$ stabilizes, then it necessarily stabilizes to the all $1$ configuration.
This step fails for the abelian sandpile except in dimension~$1$.
Indeed, a result analogous to Theorem~\ref{main} does hold for the abelian sandpile when the underlying graph is $\mathbb{Z}$ \cite[Theorem~3.2]{FMR}, but no such result can hold in higher dimensions:
The density $\mathbb{E} s$ alone is not enough to determine whether an abelian sandpile $s$ on $\mathbb{Z}^d$ stabilizes, if $d < \mathbb{E} s < 2d-1$ (see \cite[Section 5]{FR}, \cite[Theorem 3.1]{FMR} and \cite[Proposition 1.4]{FLP}; the essential idea in these arguments arose first in bootstrap percolation \cite{Ent,Sch}).
We would like to highlight an open problem: Given a probability distribution $\mu$ on $\mathbb{Z}$ (say, supported on $\{0,1,2,3,4\}$ with rational probabilities) is it algorithmically decidable whether the i.i.d.\ abelian sandpile on $\mathbb{Z}^2$ with marginal $\mu$ stabilizes almost surely?
\subsection{Quantitative estimates and bi-Laplacian field}
For a \emph{finite} connected graph $G=(V,E)$, the divisible sandpile $s : V \to \mathbb{R}$ stabilizes if and only if $ \sum_{x \in V} s(x) \le \abs{V}$.
Our next result gives the order of the odometer in a critical case when this sum is exactly $|V|$.
Specifically, to formalize the idea that `stabilization is harder in lower dimensions,' we take an identically distributed Gaussian initial condition on the discrete torus $\mathbb{Z}_n^d$, conditioned to have total mass $n^d$. The expected odometer can be taken as an indication of difficulty to stabilize: How much mass must each site emit on average? According to equation \eqref{e.phi} below, the expected odometer tends to $\infty$ with $n$ in all dimensions (reflecting the failure to stabilize on the infinite lattice $\mathbb{Z}^d$) but it
decreases with dimension.
\begin{theorem} \label{p.torus}
Let $( \sigma(\mathbf x))_{\mathbf x \in \mathbb{Z}^d_n}$ be i.i.d.\ $N(0,1)$, and consider the divisible sandpile
\[ s_{d,n} (\mathbf x) = 1 + \sigma(\mathbf x) - \frac{1}{n^d} \sum_{\mathbf y \in \mathbb{Z}^d_n} \sigma(\mathbf y). \]
Then $s_{d,n}:\mathbb{Z}^d_n \to \mathbb{R}$ stabilizes to the all $1$ configuration, and there exists a constant $C_d$ such that the odometer $u_{d,n}$ satisfies
\[
C_d^{-1} \phi_d(n) \le \mathbb{E} u_{d,n}(x) \le C_d \phi_d(n)
\]
for all $n \ge 2$, where $\phi_d$ is defined by
\begin{equation} \label{e.phi}
\phi_d(n):= \begin{cases}
n^{3/2}, & d=1 \\
n, & d=2 \\
n^{1/2}, & d=3 \\
\log n, & d=4 \\
(\log n)^{1/2}, & d \ge 5. \end{cases}
\end{equation}
\end{theorem}
The first step in computing these orders is proving an equality in law between the odometer $u_{d,n}$ and a certain `discrete bi-Laplacian Gaussian field' shifted to have minimum value $0$. This equality in law actually holds for any finite connected graph, as detailed in the next proposition.
\begin{proposition} \label{t.finite}
Let $G= (V, E)$ be a finite connected graph.
Let $( \sigma(x))_{x \in V}$ be i.i.d.\ $N(0,1)$, and consider the divisible sandpile
\[ s(x) = 1 + \sigma(x) - \frac{1}{\abs{V}} \sum_{y \in V} \sigma(y). \]
Then $s$ stabilizes to the all $1$ configuration, and the distribution of its odometer
$u : V \to [0,\infty)$ is
\[ (u(x))_{x \in V} \eqindist \left( \eta(x) - \min \eta \right)_{x \in V} \]
where the $\eta(x)$ are jointly Gaussian with mean zero and covariance
\[ \mathbb{E}[\eta(x) \eta(y)] = \frac{1}{\deg(x)\deg(y)}\sum_{z \in V} g(z,x) g(z,y) \]
where $g$ is defined by $g(x,y) = \frac{1}{\abs{V}} \sum_{z \in V} g^z(x,y)$ and $g^z(x,y)$ is the expected number of visits to $y$ by the simple random walk started at $x$ before hitting $z$.
\end{proposition}
Proposition~\ref{t.finite} suggests the possibility of a central limit theorem for the divisible sandpile odometer on $\mathbb{Z}_n^d$: We believe that if $\sigma$ is identically distributed with zero mean and finite variance, then the odometer, after a suitable shift and rescaling,
converges weakly as $n \to \infty$ to the bi-Laplacian Gaussian field on $\mathbb{R}^d$.
\subsection{Proof ideas}
By conservation of density (Proposition~\ref{p.cden}) the assumption $\mathbb{E} s = 1$ implies that if $s$ stabilizes then it must stabilize to the all $1$ configuration, so that the odometer $u$ satisfies
\begin{equation} \label{e.allones} \Delta u = 1-s. \end{equation}
where $\Delta$ is the Laplacian \eqref{e.thelaplacian}. This relation leads to a contradiction in one of three ways:
\begin{itemize}
\item If $G$ is recurrent (examples: $\mathbb{Z},\mathbb{Z}^2$), then $1+\beta \delta_o$ does not stabilize (Lemma~\ref{l.diracplusone}). By resampling the random variable $s(o)$ we derive a contradiction from \eqref{e.allones}.
\item If $G$ is simply transient (examples: $\mathbb{Z}^3, \mathbb{Z}^4$) then we can attempt to solve \eqref{e.allones} for $u$, writing
\begin{equation} \label{e.greenseries} u(y) = \sum_x g(x,y) (s(x)-1) \end{equation}
where $g$ is Green's function. The sum on the right side diverges a.s.\ if taken over all $x \in V$ (since $g(\cdot,y)$ is not square-summable) but we can stabilize $s$ in nested finite subsets $V_n \uparrow V$ instead. The corresponding finite sums, suitably normalized, tend in distribution to a mean zero Gaussian by the Lindeberg central limit theorem, contradicting the nonnegativity of $u(y)$.
\item If $G$ is doubly transient (example $\mathbb{Z}^d$ for $d \geq 5$) then the right side of \eqref{e.greenseries} converges a.s.. The difference between the left and right sides is then a random harmonic function with automorphism-invariant law. The proof is completed by showing that under mild moment assumptions any such function is an almost sure constant (Lemma~\ref{l.stationaryharmonic}).
\end{itemize}
\subsection{Related work}
The divisible sandpile was introduced in \cite{LP09,LP10} to study the scaling limits of two growth models,
rotor aggregation and internal DLA. The divisible sandpile has also been used as a device for proving an exact mean value property for discrete harmonic functions \cite[Lemma 2.2]{JLS13}.
These works focused on sandpiles with finite total mass on an infinite graph, in which case exploding is not a possibility.
In the present paper we expand the focus to sandpiles with infinite total mass.
The abelian sandpile has a much longer history: it arose in statistical physics as a model of `self-organized criticality' (SOC) \cite{BTW87,Dha90}. The dichotomy between stabilizing and exploding configurations arose in the course of a debate about whether SOC does or does not involve tuning a parameter to a critical value \cite{FR,MQ}.
Without reopening that particular debate, we view the stabilizing/exploding dichotomy as a topic with its own intrinsic mathematical interest. An example of its importance can be seen in the partial differential equation for the scaling limit of the abelian sandpile on $\mathbb{Z}^2$, which relies on a classification of certain `quadratic' sandpiles according to whether they are stabilizing or exploding \cite{LPS12}.
The Gaussian vector $\eta$ in Proposition~\ref{t.finite} can be interpreted as a discrete bi-Laplacian field. In $\mathbb{Z}^d$ for dimensions $d\geq 5$, Sun and Wu construct another discrete model for the bi-Laplacian field by assigning random signs to each component of the uniform spanning forest \cite{SW13}.
\section{Toppling procedures and stabilization} \label{s.maindefs}
In this section $G=(V,E)$ is a locally finite, connected, undirected graph. Denote by $\mathcal{X}= \mathbb{R}^{V}$ the set of divisible sandpile configurations on $G$.
\begin{definition} \label{dtoppling}
Let $T \subset [0,\infty)$ be a well-ordered set of \emph{toppling times} such that $0 \in T$ and $T$ is a closed subset of $[0,\infty)$.
A \emph{toppling procedure} is a function
\begin{align*} T \times V &\to [0,\infty) \\
(t,x) &\mapsto u_t(x) \end{align*}
satisfying for all $x \in V$
\begin{enumerate}
\item $u_0(x)=0$.
\item $u_{t_1}(x) \leq u_{t_2}(x)$ for all $t_1 \leq t_2$.
\item If $t_n \uparrow t$, then $u_{t_n}(x) \uparrow u_t(x)$.
\end{enumerate}
\end{definition}
In the more general toppling procedures considered by Fey, Meester and Redig \cite{FMR}, the assumption that $T$ is well-ordered becomes a ``no infinite backward chain'' condition, but we will not need that level of generality. See Examples \ref{e.parallel}-\ref{e.stages} below for the three specific toppling procedures we will use.
The interpretation of a toppling procedure is that starting from an initial configuration $s \in \mathcal{X}$, the total mass emitted by a site $x \in V$ to each of its neighbors during the time interval $[0,t]$ is $u_t(x)$, so that the resulting configuration at time $t$ is
\[ s_t = s + \Delta u_t \]
where $\Delta$ is the graph Laplacian \eqref{e.thelaplacian}.
For $a \in \mathbb{R}$ write $a^+ = \max(a,0)$. For $t \in T$ write $t^- := \sup \{r \in T \,:\, r<t \}$. Note that $t^- \in T$ since $T$ is closed.
\begin{definition}
A toppling proceedure $u$ is called \emph{legal} for initial configuration $s$ if
\[ u_t(x) - u_{t^-}(x) \leq \frac{(s_{t^-}(x)-1)^+}{ \operatorname{deg}(x)}\]
for all $x \in V$ and all $t \in T \setminus \{ 0 \}$.
\end{definition}
Thus, in a legal toppling procedure, a site with mass $\leq 1$ cannot emit any mass, while a site with mass $>1$ must keep at least mass $1$ for itself.
\begin{definition}
A toppling procedure $u$ is called \emph{finite} if for all $x \in V$ we have
\begin{equation*}
u_\infty(x) := \lim_{t \to \sup T} u_t(x)
< \infty
\end{equation*}
and \emph{infinite} otherwise.
The limit exists in $[0,\infty]$ since $u_t(x)$ is nondecreasing in $t$.
\end{definition}
Note that if $u$ is a finite toppling procedure, then the limit
\[ s_\infty := \lim_{t \to \sup T} s_t = s + \lim_{t \to \sup T} \Delta u_t \]
exists and equals $s + \Delta u_\infty$.
\begin{definition} \label{defstab}
Let $s \in \mathcal{X}$.
A toppling procedure $u$ is called \emph{stabilizing} for $s$ if $u$ is finite and $s_\infty \leq 1$ pointwise. We say that $s$ \emph{stabilizes} if there exists a stabilizing toppling procedure for $s$.
\end{definition}
Throughout this paper, all inequalities between functions hold pointwise, and we will usually omit the word ``pointwise.''
A basic question arises: For which $s \in \mathcal{X}$ does there exist a stabilizing toppling procedure? For instance, one might expect (correctly) that there is no such procedure for $s \equiv 2$. We can rephrase this question in terms of the set of functions
\[ \mathcal{F}_s := \{ f: V \to \mathbb{R} \,|\, f \geq 0 \text{ and } s + \Delta f \leq 1 \}. \]
If $u$ is a stabilizing toppling procedure for $s$, then $u_\infty \in \mathcal{F}_s$. Conversely, any $f \in \mathcal{F}_s$ arises from a stabilizing toppling procedure for $s$ simply by setting $T = \{0,1\}$ and $u_1 = f$. Therefore $s$ stabilizes if and only if $\mathcal{F}_s$ is nonempty.
\begin{prop}\label{lap} \moniker{Least action principle and abelian property}
Let $s \in \mathcal{X}$, and let $\ell$ be a legal toppling procedure for $s$.
\begin{enumerate}[\em (i)]
\item For all $f \in \mathcal{F}_s$,
\[ \ell_\infty \leq f. \]
\item If $u$ is any stabilizing toppling procedure for $s$, then \[ \ell_\infty \leq u_\infty. \]
\item If $u$ is any legal stabilizing toppling procedure for $s$, then for all $x \in V$,
\begin{equation} \label{e.LAP} u_\infty(x) = \inf \{f(x) \,|\, f \in \mathcal{F}_s \} \end{equation}
In particular, $u_\infty$ and the final configuration \[ s_\infty = s + \Delta u_\infty \] do not depend on the choice of legal stabilizing toppling procedure $u$.
\end{enumerate}
\end{prop}
\begin{proof}
(i) For $y \in V$ let $\tau_y = \inf \{ t \in T \,:\, \ell_t(y)>f(y) \}$, and suppose for a contradiction that $\tau_y < \infty$ for some $y \in V$.
Since $T$ is well-ordered, the infimum is attained. Moreover, $\tau_y^- := \sup \{t \in T \,:\, t < \tau_y\}<\tau_y$ by assumption (3) of Definition~\ref{dtoppling}.
Let $\tau = \inf_{y \in V} \tau_y$. Since $T$ is well-ordered, $\tau = \tau_y$ for some $y \in V$. Now at time $\tau^-$, since $u$ is legal for $s$,
\[ s_{\tau^-}(y) \geq 1 + \operatorname{deg}(y)\left(\ell_\tau(y) - \ell_{\tau^-}(y)\right) > 1 +\operatorname{deg}(y) w(y) \]
where $w = f - \ell_{\tau^-}$. On the other hand,
\begin{align*} s_{\tau^-}(y) &= (s + \Delta \ell_{\tau^-})(y) \\
&= (s + \Delta f)(y) - \Delta w (y) \\
&\leq 1 + \deg(y)w(y) - \sum_{x \sim y} w(x).
\end{align*}
It follows that $\sum_{x \sim y} w(x) < 0$. But for all $x \in V$ we have $\tau^- < \tau \leq \tau_x$, so $w(x) \geq 0$, which yields the required contradiction.
Part (ii) follows from (i), using $f = u_\infty$.
Part (iii) also follows from (i): the function $u_\infty$ simultaneously attains the infimum for all $x\in V$.
\end{proof}
Whenever we need to fix a particular toppling procedure, we will choose one of the following three.
\begin{example} \label{e.parallel}
\dmoniker{Toppling in parallel}
Let $T=\mathbb{N}$. At each time $t\in \mathbb{N}$, all unstable sites of $s_{t-1}$ topple all of their excess mass simultaneously: For all $x \in V$,
\[ u_t(x) - u_{t-1}(x) = \frac{(s_{t-1}(x)-1)^+}{\operatorname{deg}(x)}. \]
This $u$ is a legal toppling procedure. Two further observations about $u$ will be useful in the proof of Lemma~\ref{c.lap} below. First, if $s$ stabilizes, then $u$ is finite by Proposition~\ref{lap}(ii).
Second, whenever $u$ is finite, $u$ is stabilizing for $s$: indeed, if $u_\infty(x) = \frac{1}{\deg(x)} \sum_{t \in \mathbb{N}} (s_t(x)-1)^+ < \infty$, we have $(s_t(x)-1)^+ \to 0$ as $t \to \infty$ and hence $s_\infty(x) \le 1$.
\end{example}
\begin{example} \label{e.nest}
\dmoniker{Toppling in nested volumes}
Let $V_1 \subset V_2 \subset \ldots$ be finite sets with $\bigcup_{n \geq 1} V_n = V$. Between times $n-1$ and $n$ we topple in parallel to stabilize all sites in $V_n$: Formally, we take $T$ to be the set of all rationals of the form $n-\frac{1}{k}$ for positive integers $n$ and $k$.
For $n \geq 1$ and $k \geq 1$ we set
\[ u_{n-\frac{1}{k}}(x) - u_{n-\frac{1}{k-1}}(x) = \frac{(s_{n-\frac{1}{k-1}}(x)-1)^+ \cdot \mathbf{1}_{x \in V_n}}{\operatorname{deg}(x)} \]
and
\[ u_{n}(x) = \lim_{k \to \infty} u_{n-\frac{1}{k}}(x). \]
\end{example}
\begin{example} \label{e.stages}
\dmoniker{Toppling in two stages}
In this procedure we are given a decomposition of the initial configuration into two pieces
\[ s = s^1 + s^2 \]
where $s^1$ stabilizes and $s^2 \geq 0$. In the first stage we ignore the extra mass $s^2$ and stabilize the $s^1$ piece by toppling in parallel at times $1-\frac1k$ for positive integers $k$, obtaining
\[ s_1 = s + \Delta u^1_\infty = s^1_\infty + s^2. \]
The condition that $s^1$ stabilizes ensures $u^1_\infty < \infty$, and the condition $s^2 \geq 0$ ensures that all topplings that are legal for $s^1$ are also legal for $s$. Now we topple $s^1_\infty +s^2$ in parallel at times $2-\frac1k$ for positive integers $k$.
\end{example}
Now we come to a central definition of this paper. Let $s \in \mathcal{X}$.
\begin{definition}
The \emph{odometer} of $s$ is the function $u_\infty : V \to [0,\infty]$ of \eqref{e.LAP}. If $s$ stabilizes, then its \emph{stabilization} is the configuration
\[ s_\infty = s + \Delta u_\infty. \]
\end{definition}
If $s$ stabilizes, then its odometer $u_\infty(x)$ is the total amount of mass sent from $x$ to one of its neighbors, in any legal stabilizing toppling procedure for $s$.
If $s$ does not stabilize, then $u_\infty \equiv \infty$:
The odometer is defined as a pointwise infimum, with the usual convention that the infimum of the empty set is $\infty$. Next we observe that the odometer can also be expressed as a pointwise supremum.
\begin{lemma}\label{c.lap}
Let $u_\infty$ be the odometer of $s \in \mathcal{X}$. Then for all $x \in V$,
\begin{equation} \label{e.legalsup} u_\infty(x) = \sup \{ \ell_\infty(x) \,|\, \ell \mbox{ is a legal toppling procedure for $s$} \}. \end{equation}
\end{lemma}
\begin{proof}
Denote the right side of \eqref{e.legalsup} by $L(x)$. By Proposition \ref{lap}(i), $L \le u_\infty$. To prove the reverse inequality we will use a particular legal $\ell$, the parallel toppling procedure of Example~\ref{e.parallel}. There are two cases: First, if this $\ell$ is finite, then $\ell$ is stabilizing as well as legal, so $L \geq \ell_\infty = u_\infty$ by Proposition \ref{lap}(iii).
Second, if $\ell$ is not finite, then $\ell_\infty(o)=\infty$ for some $o \in V$. Then for any neighbor $x \sim o$, we have
\[
\ell_{t+1}(x) \ge \ell_{t}(o) + s(x)-1 \]
and the right side tends to $\infty$ with $t$, so $\ell_\infty(x)=\infty$.
Since the graph $G$ is connected it follows that $\ell_\infty \equiv \infty$. In this case, both $L$ and $u_\infty$ are identically $\infty$.
\end{proof}
We pause to record several equivalent conditions for $s$ stabilizing.
\begin{corollary}\label{c.equiv}
Let $s \in \mathcal{X}$ have odometer $u_\infty$. The following are equivalent.
\begin{enumerate}
\item There exists a legal stabilizing toppling procedure for $s$.
\item There exists a stabilizing toppling procedure for $s$.
\item $\mathcal{F}_s \neq \emptyset$.
\item $u_\infty(x) < \infty$ for all $x \in V$.
\item Every legal toppling procedure for $s$ is finite.
\item The parallel toppling procedure for $s$ is finite.
\end{enumerate}
\end{corollary}
\begin{proof}
The implications $(1) \Rightarrow (2) \Rightarrow (3) \Rightarrow (4)$ and $(5) \Rightarrow (6)$ are immediate from the definitions. The implication $(4) \Rightarrow (5)$ follows from Lemma~\ref{c.lap}. Finally, let $u$ be the parallel toppling procedure for $s$. If $u$ is finite, then $u$ is both legal and stabilizing for $s$, which shows the remaining implication $(6) \Rightarrow (1)$.
\end{proof}
Denote by $\mathbb{P}_o$ the law of the discrete time simple random walk $(X_j)_{j \in \mathbb{N}}$ on $G$ started at $X_0 = o$. Writing $p_j(x) = \mathbb{P}_o(X_j = x)/ \deg(x)$, observe that the Laplacian \eqref{e.thelaplacian} of $p_j$ satisfies
\begin{align} \Delta p_j(x) &= \sum_{y \sim x} \left(\frac{\mathbb{P}_o(X_j=y)}{\deg(y)} - \frac{\mathbb{P}_o(X_j=x)}{\deg(x)} \right) \nonumber \\
&= \sum_{y \sim x} \mathbb{P}_o(X_j=y, X_{j+1}=x) - \mathbb{P}_o(X_j=x) \nonumber \\
&= \deg(x) (p_{j+1}(x) - p_j(x)). \label{e.ptlaplacian}
\end{align}
\old{
\begin{align} \frac{\Delta p_t(x)}{\deg(x)} &= \frac{1}{\deg(x)} \sum_{y \sim x} (p_t(y) - p_t(x)) \nonumber \\
&= \sum_{y \sim x} \mathbb{P}_o(X_t=y, X_{t+1}=x) \nonumber \\
&= p_{t+1}(x) - p_t(x). \label{e.ptlaplacian}
\end{align}
}
\old{
\begin{align*} \Delta p_t(x)
&= \deg(x) \left( -p_t(x) + \sum_{y \sim x} \mathbb{P}_o(X_t=y, X_{t+1}=x) \right) \\
&= \deg(x) (p_{t+1}(x) - p_t(x)).
\end{align*}
}
The next lemma will play an important role in the proof of the recurrent case of Theorem~\ref{main}. It relates the stabilizability of a particular configuration $s$ to the transience of the simple random walk.
\begin{lemma} \label{l.diracplusone}
Fix $o \in V$ and $\beta>0$. The divisible sandpile
\[ s(x) = \begin{cases} 1, & x \neq o \\ 1+\beta, & x=o \end{cases} \]
stabilizes on $G$ if and only if the simple random walk on $G$ is transient.
\end{lemma}
\begin{proof}
We compute the parallel toppling procedure $u_t$ and the configuration $s_t$ at time $t$ of Example \ref{e.parallel}.
Setting $p_t(x) = \mathbb{P}_o(X_t=x)/\deg(x)$, let us show by induction on $t$ that
\begin{equation} \label{e.odomt} u_t(x)= \beta \sum_{j=0}^{t-1} p_j(x) \end{equation}
and
\begin{equation} \label{e.sandt} s_t(x)= 1 + \beta \deg(x) p_t(x) \end{equation}
for all $x \in V$ and all $t \in \mathbb{N}$. Indeed, if \eqref{e.odomt} holds at time $t$, then summing \eqref{e.ptlaplacian} we obtain
\[ s_{t} = s + \Delta u_{t} = s + \beta \deg(x) (p_t - p_0) = 1 + \beta \deg(x) p_t \]
so \eqref{e.sandt} holds at time $t$, whence
\[ u_{t+1}(x) - u_t(x) = \frac{(s_t(x)-1)^+}{\deg(x)} = \beta p_t(x) \]
so \eqref{e.odomt} holds at time $t+1$, completing the inductive step.
Taking $t \uparrow \infty$ in \eqref{e.odomt} yields $u_\infty(x)= \beta g(o,x)/\deg(x)$ where
\begin{equation} \label{e.gdef} g(o,x)= \sum_{j=0}^{\infty} \mathbb{P}_{o}(X_j=x) \end{equation}
is the Green function of $G$, which is finite if and only if $G$ is transient.
If $G$ is transient, then the parallel toppling procedure $u$ is finite and $s + \Delta u_\infty \equiv 1$, so $s$ stabilizes. If $G$ is recurrent, then the parallel toppling procedure is infinite, so $s$ does not stabilize by Corollary~\ref{c.equiv}.
\end{proof}
\section{Conservation of density}
\label{s.conservation}
In this section we assume that $G =(V,E)$ is vertex-transitive, and we fix a subgroup $\Gamma$ of $\operatorname{Aut}(G)$ that acts transitively: for any $x,y \in V$ there is an automorphism $\alpha \in \Gamma$ such that $\alpha x=y$.
For the rest of the paper, we assume $o \in V$ be an arbitrary fixed vertex.
Write $\mathcal{X}= \mathbb{R}^V$ (viewed as a measurable space with the Borel $\sigma$-field)
and $T_\alpha: \mathcal{X} \to \mathcal{X}$ for the shift $(T_\alpha f)(x)= f(\alpha^{-1} x)$.
Let $\mathbb{P}$ be a probability measure on $\mathcal{X}$
satisfying $\mathbb{E} |s(o)| <\infty$. We assume that $\mathbb{P}$ is \textbf{$\Gamma$-invariant}; that is, if $s$ has distribution $\mathbb{P}$ then $T_\alpha s$ has distribution $\mathbb{P}$ for all $\alpha \in \Gamma$.
\begin{prop}
\moniker{Conservation of Density}
\label{p.cden}
If $\mathbb{P}$ is $\Gamma$-invariant and $\mathbb{P} \{ s \text{ stabilizes} \} = 1$, then the stabilization $s_\infty$ satisfies \[ \mathbb{E} s_\infty(o) = \mathbb{E} s (o). \]
\end{prop}
Fey, Meester and Redig \cite{FMR} used an ergodic theory argument to prove the conservation of density for the abelian sandpile on $\mathbb{Z}^d$ when $\Gamma$ is the group of translations. They considered only nonnegative initial conditions: $s \geq 0$. We will give an elementary proof which starts in the same way, by choosing an automorphism-invariant toppling procedure (toppling in parallel) and uses the following observation about averages of uniformly integrable random variables.
\begin{lemma}
\label{l.unifint}
If $X$ is a random variable with $\mathbb{E} |X|<\infty$, and $X_1, X_2, \ldots$ is any sequence of random variables such that $X_i \eqindist X$ for all $i$, then the family
\[ \mathcal{S} = \{ a_1 X_1 + \ldots + a_k X_k \,|\, k \geq 1, a_i \geq 0, \sum a_i = 1 \} \]
is uniformly integrable.
\end{lemma}
\begin{proof}
We may assume $X \geq 0$.
Given $\epsilon>0$, we must show that there is a $\delta>0$ so that for any set $A$ with $\mathbb{P}(A)<\delta$ we have $\mathbb{E}(Y \mathbf{1}_A)<\epsilon$ for all $Y \in \mathcal{S}$.
Since $\mathbb{E} X < \infty$, there is such $\delta$ for X itself: choose $M$ so that $\mathbb{E}(X \mathbf{1}_{\{X>M\}}) <\epsilon$ and then set $\delta=\mathbb{P}(X>M)$.
Now for any set $A$ with $\mathbb{P}(A)<\delta$ and any $Y = \sum a_i X_i \in \mathcal{S}$
\[ \mathbb{E} [Y \mathbf{1}_A] = \sum_{i=1}^k a_i \mathbb{E} [X_i \mathbf{1}_A] < \epsilon \]
so the same $\delta$ works for all $Y \in \mathcal{S}$.
\end{proof}
\begin{proof}[Proof of Proposition~\ref{p.cden}]
We topple in parallel: at each time step $t=0,1,\ldots$, each site $x \in V$ distributes all of its excess mass $\sigma_t(x) = (s_t(x)-1)^+$ equally among its $r$ neighbors where $r$ is the common degree of all vertices in $G$.
The resulting configuration after $t$ time steps is \[ s_t = s_0 + \Delta u_t. \]
where $u_t = r^{-1}(\sigma_0 + \ldots + \sigma_{t-1})$. Since $\mathbb{P} \{ s \text{ stabilizes} \}= 1$ we have $s_t(o) \to s_\infty(o)$, a.s..
We will show that the random variables $\sigma_t(o)$ for $t \in \mathbb{N}$ are uniformly integrable. To finish the proof from there, note that the law of $u_t$ is $\Gamma$-invariant, so $\mathbb{E} u_t(x) = \mathbb{E} u_t(y)$ for all $x,y \in V$. In particular, $\mathbb{E} \Delta u_t(o)=0$ and hence $\mathbb{E} s_t(o) = \mathbb{E} s_0(o)$ for all $t<\infty$.
Since $s_t \geq \min(s_0,1)$ the uniform integrability of $\sigma_t(o)$ implies that of $s_t(o)$, so we conclude $\mathbb{E} s_\infty(o)=\mathbb{E} s_0(o)$.
At time step $t$ the origin retains mass $\leq 1$ and receives mass $\sigma_{t-1}(y)/r$ from each neighbor $y$, so
\[ s_{t}(o) \leq 1 + \sum_{x \sim o} \frac{\sigma_{t-1}(x)}{r} \]
hence
\[ \sigma_{t}(o) \leq \frac{1}{r} \sum_{x \sim o} \sigma_{t-1}(x). \]
Inducting on $t$ we find that
\[ \sigma_t(o) \leq Y_t := \sum_{x \in V} a_t(x) \sigma_0(x) \]
where $a_t(x)$ is the probability that a $t$-step simple random walk started at $o \in V$ ends at $x$. By Lemma~\ref{l.unifint} the random variables $Y_t$ are uniformly integrable, which completes the proof.
\end{proof}
We remark that the above proof also applies to the abelian sandpile to show \cite[Lemma~2.10]{FMR}, by taking $\sigma_t(x) = \floor{s_t(x)^+/r}$, the number of times $x$ topples at time $t$.
\section{Behavior at critical density}
In this section $G=(V,E)$ is an infinite vertex-transitive graph, $o \in V$ is a fixed vertex, $\Gamma \subset \operatorname{Aut}(G)$ is a group of automorphisms that acts transitively $V$, and $\mathbb{P}$ is a $\Gamma$-invariant and ergodic probability measure on $\mathcal{X} =\mathbb{R}^V$ with $\mathbb{E} |s(o)|<\infty$.
The event that $s$ stabilizes is $\Gamma$-invariant, so it has probability $0$ or $1$ by ergodicity.
\begin{lem} \label{l.notstab}
If $\mathbb{E} s(o) >1$, then $\mathbb{P} \{ s \text{ stabilizes} \} =0$.
\end{lem}
\begin{proof}
If $s$ stabilizes, then by conservation of density (Proposition~\ref{p.cden}), we have $\mathbb{E} s_\infty(o) =\mathbb{E} s(o)$.
Since the final configuration $s_\infty$ is stable we have $s_\infty(o) \leq 1$, so $\mathbb{E} s(o) \leq 1$.
\end{proof}
\begin{lem} \label{l.stable}
If $\mathbb{E} s(o) <1$ then $\mathbb{P} \{ s \text{ stabilizes} \} =1$.
\end{lem}
\begin{proof}
We will show the contrapositive. We will use the observation made in the introduction that if $s_t(o) \geq 1$ for some time $t$ then $s_T(o) \geq 1$ for all $T \geq t$.
Consider first the case that $s$ is bounded below: $\mathbb{P}(s(o) \ge M)=1$ for some $M \in (-\infty,0]$.
Define $u_t$ and $s_t$ by toppling in parallel as in \textsection\ref{s.conservation}.
Supposing that $\mathbb{P} \{s \text{ stabilizes}\} =0$, we have $u_t(o)>0$ for some sufficiently large $t$, a.s.; this fact is contained in the proof of Lemma~\ref{c.lap}.
Since $u_t(o)>0$ implies $s_t(o) \geq 1$ and hence $s_T(o) \geq 1$ for all $T \geq t$, we have
\[ \mathbb{P} \left\{ \mathop{\rm lim\,inf}\limits_{t \to \infty} s_t(o) \ge 1 \right \} = 1. \]
Since $s_t \ge \min(s,1) \ge M$, by Fatou's lemma $\mathbb{E} \left(\mathop{\rm lim\,inf}\limits_{t\to \infty} s_t(o) \right) \le \mathbb{E} s(o)$, which shows $\mathbb{E} s(o) \geq 1$ as desired.
Now we use a truncation argument to reduce the general case to case where $s$ is bounded below. Choose sufficiently small $M \in (-\infty,0]$ such that
$\mathbb{E} s(o) \mathbf{1}_{s(o) \ge M} < 1$. By the previous case the configuration $s \mathbf{1}_{s \ge M}$ stabilizes almost surely. Since $s \le s \mathbf{1}_{s \ge M}$, we have that
$s$ stabilizes almost surely.
\end{proof}
If $s$ stabilizes then the \textbf{odometer} of $s$ is the function $u_\infty : V \to [0,\infty)$, where $u$ is any legal stabilizing toppling procedure for $s$.
\begin{lem}
\label{l.allones}
If $s$ has $\Gamma$-invariant law $\mathbb{P}$ and $\mathbb{P} \{s \text{ stabilizes} \}=1$, then the odometer $u_\infty$ has $\Gamma$-invariant law. Moreover, if $\mathbb{E} s(o) = 1$ then $s_\infty \equiv 1$ and $\Delta u_\infty = 1 - s$.
\end{lem}
\begin{proof}
We use \eqref{e.LAP} along with the fact that $\Delta$ commutes with $T_\alpha$: if $s + \Delta f \leq 1$ then $T_\alpha s + \Delta (T_\alpha f) = T_\alpha (s + \Delta f) \leq 1$, so if $u_\infty$ is the odometer for $s$ then $T_\alpha u_\infty$ is the odometer for $T_\alpha s$.
By conservation of density (Proposition~\ref{p.cden}), $\mathbb{E} s_\infty(o)=1$. Since $s_\infty \leq 1$ it follows that $s_\infty \equiv 1$, and hence $\Delta u_\infty = 1-s$.
\end{proof}
In the case that $G$ is recurrent we can prove our main theorem with no moment assumption, using an ``extra head'' construction.
\begin{lem} \label{l.recurrent}
Suppose that $G$ is an infinite, recurrent, vertex-transitive graph and $\{ s(x)\}_{x \in V}$ are i.i.d.\ with $\mathbb{E} s(x) = 1$ and $\mathbb{P} \{s(x)=1 \} \neq 1$. Then \[ \mathbb{P} \{ s \mbox{ stabilizes} \} =0.\]
\end{lem}
\begin{proof}
Let $S \subset \mathcal{X}$ denote the set of configurations that stabilize.
By ergodicity, $\mathbb{P}(S) \in \{0,1\}$.
We will show that if $\mathbb{P}(S)=1$, then the graph $G$ must be transient. Let $s:V \to \mathbb{R}$ denote an i.i.d.\ configuration with the given distribution such that $\mathbb{E} s=1$, and fix a vertex $o \in V$.
We create a new i.i.d.\ configuration $s'$ with the same law as $s$, by independently resampling $s(o)$ from the same distribution.
Then $s'= s + \beta \delta_o$, where $\delta_o(x) = \mathbf{1} \{x = o\}$ and $\beta$ is a mean zero random variable. Since $\mathop{\mathrm{Var}} s >0$, we have that $\mathbb{P}(\beta >0) >0$. Using $\mathbb{P}(S ) =1$ along with Lemma~\ref{l.allones}, there exist
$s:V \to \mathbb{R}$ and $\beta>0$ such that $s$ and $s+\beta \delta_o$ both stabilize to the all $1$ configuration.
By toppling $s+\beta \delta_o$ in two stages (Example~\ref{e.stages}), it follows that $1+\beta \delta_o$ stabilizes to $1$, so $G$ is transient by Lemma \ref{l.diracplusone}.
\end{proof}
In the preceding lemma the hypothesis that $s$ is i.i.d.\ can be substantially weakened: The proof uses only the fact that with positive probability, the conditional distribution of $s(o)$ given $\{s(x)\}_{x \neq o}$ is not a single atom.
\section{Proof of Theorem \ref{main}}
\subsection{Singly transient case} \label{s.strans}
Recall Green's function \eqref{e.gdef}. In this section we assume that $g(o,y)<\infty$ for all $y \in V$ but
\begin{equation} \label{e.singlytransient} \sum_{y \in V} g(o,y)^2 = \infty. \end{equation}
Define $V_n= \{ x \in V : d(x,o) \le n \}$ where $d$ is the graph distance.
Then $V_1 \subset V_2 \subset \ldots$ are finite sets with $\bigcup_{n \geq 1} V_n = V$.
Let $g_n(x,y)$ be the expected number of visits to $y$ by simple random walk started at $x$ and killed on exiting $V_n$.
By the monotone convergence theorem, for fixed $x,y \in V$ we have
\[ g_n(x,y) \uparrow g(x,y) \]
as $n \to \infty$.
In particular, setting
\[ \nu_n := \left( \sum_{y \in V_n} g_n(o,x)^2 \right)^{1/2} \]
we have $\nu_n \uparrow \infty$ as $n \to \infty$ by \eqref{e.singlytransient}.
The proof of the following lemma is inspired by
\cite[Theorem 3.1]{FR} and \cite[Theorem 3.5]{FMR}.
\begin{lemma} \label{l.nondeg} Let $G=(V,E)$ be singly transient \eqref{e.singlytransient}, vertex transitive graph.
Let $\{ s(x)\}_{x \in V}$ be i.i.d.\ with $\mathbb{E} s = 1$ and $\mathop{\mathrm{Var}} s < \infty$. If \[ \frac{1}{\nu_n} \sum_{x \in V_n} g_n(o,x) (s(x) - 1) \]
converges in distribution as $n \to \infty$ to a nondegenerate normal random variable~$Z$, then $\mathbb{P} \{ s \text{\emph{ stabilizes}} \} = 0$.\end{lemma}
\begin{proof}
Assume to the contrary that $s$ stabilizes a.s.. Let $u$ be the nested volume toppling procedure (Example \ref{e.nest}). Then for each $n \in \mathbb{N}$,
\begin{equation}\label{e.finitevolume}
s + \Delta u_n = \xi_n \qquad \text{on } V_n
\end{equation}
with $\xi_n \leq 1$ on $V_n$.
Equation \eqref{e.finitevolume} can be rewritten as
\[
u_n (y) = r^{-1}\sum_{x \in V_n} g_n(x,y) (s(x) -1) + r^{-1}\sum_{x \in V_n} g_n(x,y) (1 - \xi_n(x))
\]
where $r$ is the common degree
(both sides have Laplacian $\xi_n - s$ in $V_n$ and vanish on $V_n^c$).
Observe that the second term is a nonnegative random variable. Therefore, for any $\epsilon>0$
\begin{equation} \label{e.decomp}
\mathbb{P} \{ u_n (o) > \epsilon \nu_n \} \geq \mathbb{P} \left\{ \frac{1}{\nu_n} \sum_{x \in V_n} g_n(o,x) (s(x) - 1) >r \epsilon \right\}.
\end{equation}
Since $u$ is a legal toppling procedure and we have assumed that $s$ stabilizes a.s., we have $u_\infty(o) < \infty$, a.s..
Now since
$u_n(o) \uparrow u_\infty(o)$ and $\nu_n \uparrow \infty$, the left side of \eqref{e.decomp} tends to zero as $n \rightarrow \infty$.
However, the right side tends to a positive limit $P(Z>r\epsilon)>0$, which gives the desired contradiction.
\end{proof}
To complete the proof of Theorem~\ref{main} in the singly transient case, it remains to show that $\frac{1}{\nu_n} \sum_{x \in V_n} g_n(o,x) (s(x) - 1)$ converges in distribution to a nondegenerate normal random variable. We show this using Lindeberg-Feller central limit theorem as follows.
\begin{lemma}\label{l.clt}
Let $\{X_{n1}: n \ge 1; i=1,\ldots,k_n\}$ be a triangular array of identically distributed random variables such that for each $n \in \mathbb{N}$, $\{X_{ni}: i=1,\ldots,k_n\}$ is independent.
Assume that $E[X_{11}]=0$ and $E[X_{11}^2]=1$. Let $a_{nk}>0$ be such that $\sum_{k=1}^{k_n} a_{nk}^2= 1$ and $\lim_{n \to \infty} b_n=0$ where $b_n = \max_{1\le k \le k_n} a_{nk}$. Then the sequence
$Y_n= \sum_{k=1}^{k_n} a_{nk}X_{nk}$ converges in distribution to standard normal random variable as $n \to \infty$.
\end{lemma}
\begin{proof}
By Lindeberg-Feller central limit theorem \cite[Theorem 27.2] {Bil} it suffices to check the Lindeberg condition:
\[
\lim_{n \to \infty} \sum_{k=1}^{k_n} \mathbb{E}[ a_{nk}^2 X_{nk}^2 \mathbf{1}_{\abs{a_{nk} X_{nk}} > \epsilon} ]= 0
\]
for all $\epsilon>0$. Since
$ \mathbb{E}[ X_{nk}^2 1_{\abs{a_{nk} X_{nk}} > \epsilon} ] \le \mathbb{E}[ X_{11}^2 1_{\abs{ X_{11}} > \epsilon/b_n} ] $ and $\sum_k a_{nk}^2=1$,
we have
\[
\lim_{n \to \infty} \sum_{k=1}^{k_n} \mathbb{E}[ a_{nk}^2 X_{nk}^2 \mathbf{1}_{\abs{a_{nk} X_{nk}} > \epsilon} ] \le \lim_{n \to \infty} \mathbb{E}[ X_{11}^2 \mathbf{1}_{\abs{ X_{11}} > \epsilon/b_n} ] =0
\]
because $\lim_{n \to \infty} b_n=0$ and $\mathbb{E}[X_{11}^2]=1$.
\end{proof}
\begin{lemma}\label{l.singlytransient}
Let $G=(V,E)$ be a singly transient \eqref{e.singlytransient}, vertex transitive graph. Suppose $\{ s(x)\}_{x \in V}$ are i.i.d.\ random variables with $\mathbb{E} s = 1$ and $ 0< \operatorname{Var} s < \infty$. Then $\mathbb{P} \{s \text{ stabilizes} \} = 0$.
\end{lemma}
\begin{proof}
Since $g(o,o)< \infty$ and $\nu_n \uparrow \infty$, we have
\[
\lim_{n \to \infty} \max_{x \in V_n} \frac{g_n(o,x)}{\nu_n} = \lim_{n \to \infty} \frac{g_n(o,o)}{\nu_n} \le \lim_{n \to \infty} \frac{g(o,o)}{\nu_n} =0.
\]
For $n \in \mathbb{N}$ and $x \in V_n$, define $X_{nx}= (s(x)-1)/\sqrt{\operatorname{Var}{s}}$ and $a_{nx}= \frac{g_n(o,x)}{ \nu_n}$ and
$b_{n}= \max_{x \in V_n} a_{nx}$. By Lemma~\ref{l.clt}, we have that
\[ \frac{1}{\nu_n} \sum_{x \in V_n} g_n(o,x) (s(x) - 1) \] converges in distribution to a normal random variable with mean 0 and variance $\operatorname{Var}(s)$. Lemma~\ref{l.nondeg} implies the desired conclusion.
\end{proof}
\subsection{Doubly transient case} \label{s.dtrans}
We start by outlining our proof strategy. Recall from Lemma~\ref{l.allones} that if $s$ stabilizes with $\mathbb{E} s(o)=1$ then it must stabilize to the constant configuration $s_\infty \equiv 1$. In particular, the odometer $u_\infty$ satisfies $\Delta u_\infty = 1-s$. In Lemma~\ref{l.vfun} below, by convolving $s-1$ with Green's function we can build an explicit function $v$ with Laplacian $1-s$; the convolution is defined almost surely provided that
\begin{equation} \label{e.greensumofsquares} \sum_{y \in V} g(o,y)^2 < \infty. \end{equation}
(This condition says that the expected number of collisions of two independent random walks started at $o$ is finite. The essential feature of Green's function here is of course that
\begin{equation} \label{e.laplacianofg} \Delta g(\cdot,y) = -r\delta_y \end{equation}
where $r$ is the common degree of all vertices of $G$, and $\delta_y(x) = \mathbf{1}\{x = y\}$.)
Having built the function $v$, the difference $v-u_\infty$ is then a random harmonic function with $\Gamma$-invariant law, which must be an almost sure constant by the following lemma.
\begin{lemma}
\label{l.stationaryharmonic}
\moniker{Harmonic Functions With Invariant Law}
Let $\Gamma$ be a group of automorphisms of $G$ that acts transitively on the vertex set $V$.
Suppose that $h: V \to \mathbb{R}$ has $\Gamma$-invariant law and $\Delta h \equiv 0$.
If $h$ can be expressed as a difference of two functions $h=v-u$ where $u \geq 0$ and $\sup_{x \in V} \mathbb{E} v(x)^+ <\infty$, then $h$ is almost surely constant.
\end{lemma}
\begin{proof}
Let $(X_n)_{n \geq 0}$ be simple random walk on $G$ started at $X_0=o$. Although $h$ is harmonic, $h(X_n)$ need not be a martingale since it need not have finite expectation. But for any $a \in \mathbb{R}$, since $h = v-u$ and $u \geq 0$, the truncation
\begin{align*} M_n &:= a + (h(X_n)-a)^+ \\
&\ghost{:}\leq a + (v(X_n) - a)^+ \end{align*}
has finite expectation. Since the function $t \mapsto (t-a)^+$ is convex, $x \mapsto (h(x)-a)^+$ is subharmonic, so $M_n$ is a submartingale:
\[ \mathbb{E}[M_{n+1} | h, X_0, \ldots, X_n] = \frac1r \sum_{w \sim X_n} (a + (h(w)-a)^+) \geq M_n. \]
A submartingale bounded in $L^1$ converges almost surely \cite[11.5]{Williams}. Since this holds for any $a \in \mathbb{R}$ it follows that $h(X_n)$ converges almost surely.
But since $h$ has $\Gamma$-invariant law, $h(X_n)$ is a stationary sequence,
so the only way it can converge a.s.\ is if $h(X_n)=h(X_0)$ for all $n$. Since $G$ is connected, for any vertex $x \in V$ there exists $n$ with $\mathbb{P}(X_n=x)>0$, so $h(x)=h(o)$.
\old
Thus the usual proof of upcrossing lemma works. Since $u \geq 0$ we have $\sup \mathbb{E}h(X_n)^+ \le \mathbb{E} v(X_n)^+ < \infty$, and
we can modify to the proof of martingale convergence theorem (Theorem 5.2.8 in \cite{Dur}) to show that $h(X_n)$ converges almost surely.
}
\end{proof}
\begin{lem} \label{l.vfun} Let $\Gamma$ be a countable subgroup of $\operatorname{Aut}(G)$ which acts transitively on the vertices of $V$.
Suppose $\{ \sigma(x) \}_{x \in V}$ are i.i.d.\ random variables with $\mathbb{E} \sigma(x) = 0$ and $0 < \operatorname{Var} \sigma(\mathbf{0}) < \infty$ Let $y_1,y_2,\ldots$ be an enumeration of the vertex set of $G$. For $\alpha \in \Gamma$, let
\begin{equation} \label{e.vz} v_\alpha(x) := \frac{1}{r} \sum_{i=1}^\infty g(x,\alpha y_i) \sigma(\alpha y_i). \end{equation}
If $G$ is doubly transient \eqref{e.greensumofsquares}, then the following hold almost surely.
\begin{enumerate}[{\em (a)}]
\item The series defining $v_\alpha(x)$ converges for all $x \in V, \alpha \in \Gamma$.
\item $v_\alpha = v_\beta$ \ for all $\alpha, \beta \in \Gamma$.
\item $\Delta v_e = -\sigma$ where $e$ denotes the identity automorphism.
\item $v_e$ has $\Gamma$-invariant law, that is $T_\alpha v_e \eqindist v_e$ for all $\alpha \in \Gamma$.
\item $v_e$ is unbounded above and below.
\end{enumerate}
\end{lem}
\begin{proof}
(a) Each term in the series \eqref{e.vz} is independent, and the $i$-th term has variance $g(x,\alpha y_i)^2 U$ where $U = r^{-2}\mathbb{E} \sigma(\mathbf{0})^2$. By \eqref{e.greensumofsquares} the sum of these variances is finite, which implies that the series converges a.s..
(b) Note that there is something to check here because the series defining $v_\alpha$ is only conditionally convergent.
Fix $x \in V$ and denote the partial sum $r^{-1}\sum_{i=1}^n g(x,\alpha y_i) \sigma(\alpha y_i)$ by $v_{\alpha,n}$.
We compare $v_{\alpha,n}$ and $v_{e,n}$ where $e$ is the identity element of $\Gamma$.
Given $\epsilon > 0$, choose $n_1$ such that
\[ U \sum_{i=n_1}^\infty g(x, y_i)^2 < \epsilon^3. \]
There exists $N_1$ depending on $n_1$ and $\alpha$ such that
\[ \{ y_i: i=1,2,\ldots,n_1 \} \subset \{ \alpha y_i: i=1,2,\ldots,N_1 \}. \]
This implies that for any $N \ge N_1$ we have $\operatorname{Var} \left( v_{\alpha,N}- v_{e,N} \right)
< 2\epsilon^3$. By Chebyshev's inequality,
\begin{eqnarray*}
\mathbb{P} \left( \left| v_{\alpha,N}- v_{e,N} \right| > \epsilon \right) \le \epsilon^{-2} \operatorname{Var} \left( v_{\alpha,N} - v_{e,N} \right)
< 2 \epsilon
\end{eqnarray*}
for all $N \ge N_1$.
By part (a) there a.s.\ exists $N_2 \ge N_1$
such that $\max(|v_{\alpha,n}- v_\alpha(x)|,|v_{e,n}- v_e(x)|)< \epsilon/3$ for all $n \ge N_2$. By the triangle inequality it follows that
\begin{eqnarray*}
\mathbb{P} \left( \left| v_{\alpha}(x)- v_{e}(x)\right| > \frac{\epsilon}{3} \right) < 2 \epsilon.
\end{eqnarray*}
Since $\epsilon$ was arbitrary we obtain $v_\alpha(x) = v_e(x)$, a.s.. By taking countable intersections we have that $v_\alpha = v_e$, a.s..
(c) Using \eqref{e.laplacianofg}, we compute
\begin{eqnarray*}
\Delta v_e (x) &=& \frac{1}{r} \sum_{w\sim x} (v_e(w) - v_e(x)) \\
&=& \frac{1}{r} \sum_{w \sim x} \left( \sum_{i=1}^\infty g(w,y_i) \sigma(y_i) - \sum_{i=1}^\infty g(x,y_i) \sigma(y_i) \right) \\
&=& \sum_{i=1}^\infty \sigma(y_i) \frac{1}{r} \sum_{w \sim x} \left( g(w,y_i) - g(x,y_i) \right) \\
&=& - \sum_{i=1}^\infty \sigma(y_i) \mathbf{1} \{y_i =x \} \\
&=& - \sigma(x).
\end{eqnarray*}
(d) To show that $v_e$ has $\Gamma$-invariant law, write $v_e = v_e^\sigma$ to make the dependence on the initial configuration $\sigma$ explicit. We have for all $\alpha \in \Gamma$
\begin{align*}
T_\alpha v_e^\sigma (x)
&= \sum_{i=1}^\infty g(\alpha^{-1}x, y_i) \sigma(y_i) \\
&= \sum_{i=1}^\infty g(x, \alpha y_i) \sigma(y_i) \\
&= \sum_{i=1}^\infty g(x, \alpha y_i) (T_\alpha \sigma)(\alpha y_i) \\
&= v_\alpha^{T_\alpha \sigma}(x) \\
&= v_e^{T_\alpha \sigma}(x)
\end{align*}
where in the last equality we have used part (b). Hence $T_\alpha v_e^\sigma = v_e^{T_\alpha \sigma} \eqindist v_e^\sigma$ since $\sigma$ has $\Gamma$-invariant law.
(e)
To show that $v_e$ is almost surely unbounded below, we use the assumption that $\sigma(o)$ has zero mean and positive variance, which implies that
\[ \mathbb{P}(\sigma(o) < -\delta)>p \]
for some $p,\delta>0$. Since $\sum_{y \in V} g(o,y) = \infty$ and $\sum_{y \in V} g(o,y)^2<\infty$, we can choose $N$ large enough so that
\[ \delta r^{-1}\sum_{i=1}^N g(o,y_i) > 2M, \qquad U \sum_{i=N+1}^\infty g(o,y_i)^2 < 1. \]
By Chebyshev's inequality
\[ \mathbb{P} ( r^{-1}\sum_{i=N+1}^\infty g(x,y_i) \sigma(y_i) \geq M) \leq \frac{1}{M^2}. \]
On the event that $\sigma(y_i)<-\delta$ for all $1 \leq i \leq N$ and $r^{-1}\sum_{i=N+1}^\infty g(x,y_i) \sigma(y_i) \geq M$ we have $v_e(o) < -2M + M$. Since the random variables $\sigma(y_i)$ are i.i.d., we obtain
\[ \mathbb{P} ( v_e(o) < -M) \geq p^N (1 - \frac{1}{M^2}) > 0. \]
Since $v_e$ is stationary, we have $\mathbb{P}( \inf v_e < -M ) \in \{0,1\}$ by ergodicity. Since this probability is $>0$, it must be $1$. Since $M$ was arbitrary, $v_e$ is a.s.\ unbounded below.
A similar argument shows that $v_e$ is also a.s.\ unbounded above.
\end{proof}
\begin{lem}
Let $G=(V,E)$ be a doubly transient \eqref{e.greensumofsquares}, vertex transitive graph and let $\{ s(x)\}_{x \in V}$ be i.i.d.\ random variables with $\mathbb{E} s = 1$ and $0 < \mathop{\mathrm{Var}} s < \infty$. Then $\mathbb{P} \{ $s$ \text{ stabilizes} \} = 0$.
\end{lem}
\begin{proof} Assume to the contrary that $s$ stabilizes a.s.\ with odometer $u_\infty$.
Let $\Gamma$ be a countable subgroup of $\operatorname{Aut}(G)$ that acts transitively on the vertices of $G$.
Let $v : V \to \mathbb{R}$ be given by equation \eqref{e.vz} with $\alpha = e$ and $\sigma = s-1$. By Lemma \ref{l.vfun}, $v$ has $\Gamma$-invariant law and $\Delta v = 1-s$. Therefore
\[ h := v-u_\infty \]
has $\Gamma$-invariant law and $h$ is harmonic: $\Delta h \equiv 0$ on $V$.
Further, by Fatou's lemma, $\mathbb{E} v(x)^2 \leq \deg(x)^{-2} \operatorname{Var} (s) \sum_{i=1}^\infty g(x,y_i)^2 < \infty$. Lemma~\ref{l.stationaryharmonic} now implies that $h$ is almost surely constant.
This contradicts Lemma \ref{l.vfun} because $u_\infty \ge 0$ and $v$ is almost surely unbounded below.
\end{proof}
\section{Stabilizability of Cones}
\label{s.cones}
Until now we have mainly been concerned with the stabilizability of random initial configurations. In this section we examine stabilizability of a few deterministic configurations on the square grid $\mathbb{Z}^2$. We present two examples, one of which stabilizes.
\begin{lem}
Define $C_1 = \{ (x,y) \in \mathbb{Z}^2 : x \ge 0, \abs{y} \le x \}$. Then the configuration
$s_0= (1+ \alpha) \mathbf{1}_{C_1}$ does not stabilize for $\alpha>0$.
\end{lem}
\begin{proof}
By least action principle (Proposition \ref{lap}), it suffices to show the existence of an infinite legal toppling procedure.
Let $u_k$ denote the parallel toppling procedure of Example \ref{e.parallel} where $k \in \mathbb{N}$.
Let $C= \{ (x,y)\in \mathbb{Z}^2: x > 0, \abs{y}< x \}$. Let $(X_n,Y_n)$ denote the simple random walk on $\mathbb{Z}^2$ and let $N$ denote the stopping time $N= \min \{ n \ge 0: X_n= \abs{Y_n} \}$.
As in the proof of Lemma \ref{l.diracplusone}, we keep track of the mass from each $(x,y) \in C$ to obtain
\[
u_k(1,0) \ge \frac{\alpha}{4} \sum_{(x,y) \in C} \sum_{l=0}^k \mathbb{P}_{(x,y)} ( (X_{N},Y_{ N}) = (0,0), N=l ).
\]
for all $k \in \mathbb{N}$.
To see this note that $N$ is the exit time of the set $C$ and the only way to exit $C$ at $(0,0)$ is from $(1,0)$.
As a result, we have
\[ u_\infty(1,0) \ge \frac{\alpha}{4} \sum_{(x,y) \in C} p(x,y) \]
where $p(x,y)=\mathbb{P}^{(x,y)}\left( (X_N,Y_N)=(0,0) \right)$.
Consider the function $q:\mathbb{Z}^2 \to [0,1]$ given by
\begin{equation*}
q(x,y)=
\begin{cases} p(\abs{x},y) & \text{if $\abs{y}< \abs{x}$,}
\\
-p(\abs{y},x) &\text{if $\abs{x} < \abs{y}$,}
\\
0 & \text{ otherwise.}
\end{cases}
\end{equation*}
Note that $\Delta q = \delta_{(0,1)} + \delta_{(0,-1)} -\delta_{(1,0)} -\delta_{(-1,0)}$. Let $g(x,y)$ denote the potential kernel in $\mathbb{Z}^2$ defined by
\[
g(x,y)= \sum_{n=0}^\infty [\mathbb{P}_{(0,0)} (X_n = (x,y)) - \mathbb{P}_{(0,0)}(X_n = (0,0))]
\]
where $(X_n)_{n \in \mathbb{N}}$ denotes the simple random walk on $\mathbb{Z}^2$. Although the simple random walk on $\mathbb{Z}^2$ is transient, it turns out that the sum defining
$g$ is absolutely convergent.
By standard estimates on $g$ (See \cite[Chapter 1]{Law}), we know that $g$ has sub-linear (logarithmic) growth.
This combined with \cite[Theorem 6.1]{HS} implies that
\[ q(x,y) - \frac{1}{4} \left( g(x+1,y)+g(x-1,y)-g(x,y-1)-g(x,y+1) \right) \] is identically zero because it is harmonic with sub-linear growth and attains the value $0$ at $(0,0)$.
Therefore there exists $c_1>0$ such that for all $(x,y) \in C$, we have
\begin{eqnarray*}
q(x,y)&=& \frac{1}{4} \left( g(x+1,y)+g(x-1,y)-g(x,y-1)-g(x,y+1) \right) + \frac{1}{4} \Delta g(x,y)\\\
&=& \frac{1}{2} \left( g(x+1,y)+g(x-1,y)-2g(x,y) \right) \\
& > & c_1 \frac{x^2 - y^2}{(x^2+y^2)^2}
\end{eqnarray*}
The first line above follows from the fact that $\Delta g = 0$ for all points except $(0,0)$.
The last line above follows from \cite[Theorem 1.6.5 (b)]{Law}.
Therefore
$ u_\infty(1,0) \ge \frac{\alpha}{4} \sum_{(x,y) \in C} q(x,y) > \frac{\alpha c_1}{4} \sum_{(x,y) \in B} \frac{x^2-y^2}{ (x^2 +y^2)^2} = \infty$. Hence $s_0$ does not stabilize.
\end{proof}
We need the following technical lemma for the next example:
\begin{lemma} \label{l.hillhole}
Let $\sigma: V \to \mathbb{R}$ be a configuration in $G=(V,E)$. Assume that $H:= \{x : \sigma(x) > 1 \}$ satisfies $\abs{ H } < \infty $ and
\[ \sum_{x \in H} (\sigma(x)- 1)^+ < \infty.
\]
Let $F \subset V$ be such that $\abs{F} < \infty$ and
\[
\sum_{x \in V} (\sigma(x)- 1)^+ \le \sum_{x \in F} (1-\sigma(x))^+.
\]
Then $\sigma$ stabilizes.
\end{lemma}
\begin{proof}
Let $a: H \times F \to [0,\infty)$ be a non-negative function such that
$\sum_{y \in F} a(x,y) = \sigma(x) -1 $ for all $x \in H$ and
$ \sum_{x \in H} a(x,y) \le 1 -\sigma(y)$
for all $y \in F$. The function $a$ encodes how to redistribute the excess mass from $H$ to $F$.
Let \[ f_{x,z}(y) = \frac{g_z(x,y)}{\deg(y)} = \frac{ \mathbb{E}^x ( \mbox{number of vists to $y$ before being killed at $z$})}{\deg(y)}\] denote the Green's function normalized with degree.
Observe that $\Delta f_{x,z} = \delta_z - \delta_x$.
Therefore the function $u:= \sum_{x \in H, y \in F} a(x,y) f_{ x, y}$ satisfies $\sigma + \Delta u \le 1$ and $u \ge 0$.
This in turn implies that $\sigma$ stabilizes.
\end{proof}
\begin{remark}
The condition that $\abs{H}, \abs{F} < \infty$ in the above lemma is necessary. The following example illustrates this: Consider a probability measure $\mu$ on $\mathbb{N}^* = \{ 1,2,3,\ldots\}$
and consider the function $\sigma_\mu = 1 + \delta_0 - \mu(x)$ where $\mu(x) = \mu ( \{ x\})$. Then it can be shown that $\sigma_\mu$ stabilizes if and only if $\sum_{x \in \mathbb{N}^*} x \mu(x) < \infty$.
\end{remark}
\begin{lem}
\label{l.conestab}
Define $C_a= \{ (x,y) \in \mathbb{Z}^2: x \ge 0, \abs{y} \le ax \}$.
Then the configuration\ $s_a= m \mathbf{1}_{C_a}$ stabilizes if $\frac{2ma}{1+a^2} \le 1$ and $a \in (0,1]$. Moreover $s_0(x,y) = x \mathbf{1}_{\{x > 0, \, y=0\}} $ stabilizes.
\end{lem}
\begin{proof}
The case $a=1$ is trivial. Define for $a \in (0,1]$ \[u_a(x,y) = \frac{ \left(ax - \abs{y}\right)^2 }{2(1+a^2) } \mathbf{1}_{C_a} .\]
To see that $s_0$ stabilizes, we check that $s_0 + \Delta u_1 \le 1$. This follows immediately from the computation of $\Delta u_1$ as
\begin{eqnarray*}
\Delta u_1(x,y)&=& \begin{cases} 1-x & \mbox{if } x > 0 ,y =0 \\
1 & \mbox{if }\abs{y} < x, x>0 \\
\frac{1}{2}& \mbox{if }\abs{y} = x, x>0 \\
\frac{1}{4} & \mbox{if } x=y=0 \\
0 & \mbox{otherwise.}
\end{cases}
\end{eqnarray*}
\old{ In $C_a$ we have
\begin{eqnarray*}
(1+a^2) (u_1 - m u_a) &=& (1+a^2)(x- \abs{y})^2 - 2m (ax - \abs{y})^2 \\
& \ge & 2m\left( a(x- \abs{y})^2 -(ax - \abs{y})^2 \right)\\
& = & 2m(1-a) \left( a x^2 - \abs{y}^2 \right)\\
& \ge & 2m(1-a) \left( a^2 x^2 - \abs{y}^2 \right) \\
& \ge 0
\end{eqnarray*}
Outside $C_a$ it is clear that $u_1 - m u_a \ge 0$. }
A direct computation yields $\Delta u_a$ in different regions:
$\Delta u_a(0,0)= \frac{a^2}{2(1+a^2)} \le \frac{1}{4}$.
If $y=0, x > 0$ and all neighbors of $(x,0)$ are in $C_a$ (\emph{i.e.} $x \ge \lceil 1/a \rceil$), then
\begin{eqnarray*}
\Delta u_a(x,0) &=& 1- \frac{2a}{1+a^2} x.
\end{eqnarray*}
If $y=0, x > 0$ and $(x,\pm 1) \notin C_a$, then
\begin{eqnarray*}
\Delta u_a(x,0) &=& \frac{a^2(1-x^2)}{1+a^2}.
\end{eqnarray*}
If all neighbors of $(x,y)$ are in $C_a$ with $y \neq 0$, then
\begin{eqnarray*}
\Delta u_a(x,y) &=& 1.
\end{eqnarray*}
If $(x,y) \in C_a$ with $y \neq 0$ and one of the neighbors is not in $C_a$, then
\[
0 < \frac{a^2}{1+a^2} \le \Delta u_a(x,y) \le 1.
\]
If $(x,y) \notin C_a$ and if all neighbors of $(x,y)$ are not in $C_a$, then
\begin{eqnarray*}
\Delta u_a(x,y) &=& 0.
\end{eqnarray*}
If $(x,y) \notin C_a$ and one of the neighbors in $C_a$, then
at-least 2 of the neighbors are not in $C_a$ and
\[
0 \le \Delta u_a(x,y) < 1.
\]
Consider $v_a(x,y) = u_1(x+ \lceil 1/a \rceil,y) - m u_a(x+ \lceil 1/a \rceil,y) $. It is easy to check that
$v_a \ge 0$ if $\frac{2ma}{1+a^2} \le 1$.
If $x>0$ we have
\begin{eqnarray*}
s_a(x,0)+ \Delta v_a(x,0) &=& m + \left(1- \lceil 1/a \rceil - x\right) - m \left(1 - a\left( x+ \lceil 1/a \rceil\right) \right) \\
&\le & 1+ \left( \frac{2am}{1+a^2} - 1 \right) \left(x+ \lceil 1/a \rceil\right) \\
& \le & 1.
\end{eqnarray*}
If $x>0,y\neq 0, (x,y) \in C_a$, then
\begin{eqnarray*}
s_a(x,y)+ \Delta v_a(x,y) &=& m + 1 - m = 1.
\end{eqnarray*}
If $y\neq 0, (x,y) \notin C_a$, then
\begin{eqnarray*}
s_a(x,y)+ \Delta v_a(x,y) &\le& \Delta u_1 \le 1.
\end{eqnarray*}
If $x< - \lceil 1/a \rceil - 1$, then
\begin{eqnarray*}
s_a(x,y)+ \Delta v_a(x,y) &=& 0.
\end{eqnarray*}
Therefore $s_a+ \Delta v_a \le 1$ for all points except on the finite set $\{ (x , 0) : - \lceil 1/a \rceil - 1 \le x \le 0 \}$ and
$s_a+ \Delta v_a = 0$ for all $x <-1 - \lceil 1/a \rceil$. Lemma \ref{l.hillhole} implies that $s_a+ \Delta v_a $ stabilizes, and therefore
$s_a$ stabilizes.
\end{proof}
We conjecture that the bound in Lemma~\ref{l.conestab} is sharp.
\begin{conjecture}
Define for $a \in (0,1]$, $C_a= \{ (x,y) \in \mathbb{Z}^2: x \ge 0, \abs{y} \le ax \}$.
Then the divisible sandpile\ $s_a= m \mathbf{1}_{C_a}$ stabilizes if and only if $\frac{2ma}{1+a^2} \le 1$. Furthermore, the divisible sandpile $s_0 = k x \mathbf{1}_{\{x>0,\, y=0\}} $ stabilizes if and only if $k\le 1$.
\end{conjecture}
More generally, we have the following problem.
\begin{open}[Tests for stabilizability] Given $s:\mathbb{Z}^d \to \mathbb{R}$, find series tests or other criteria that can distinguish between stabilizing and exploding $s$.
\end{open}
\section{Finite graphs}
Let $G= (V, E)$ be a finite connected graph with $|V| = n$. For a finite connected graph, all harmonic functions are constant: the kernel of $\Delta$ is $1$-dimensional spanned by the constant function $1$.
\begin{lemma}
\label{l.massn}
Let $s: V \to \mathbb{R}$ be a divisible sandpile with $\sum_{x \in V} s(x) = n$. Then $s$ stabilizes to the all $1$ configuration, and the odometer of $s$ is the unique function $u$ satisfying $s + \Delta u = 1$ and $\min u =0$.
\end{lemma}
\begin{proof}
Since $\Delta$ has rank $n-1$ and $\sum_{x \in V} (s(x) -1) = 0$, we have $s-1 = \Delta v$ for some $v$. Letting $w = v-\min v$, we have $w \geq 0$ and $s + \Delta w = 1$, so $s$ stabilizes.
Now if $u$ is any function satisfying $s+ \Delta u \leq 1$, then
\[ \sum_{x\in V} (s+\Delta u)(x) = n \]
so in fact $s+\Delta u = 1$. This shows that $s$ stabilizes to the all $1$ configuration, and moreover any two functions $u$ satisfying $s+\Delta u \leq 1$ differ by an additive constant. By the least action principle (Proposition~\ref{lap}), among these functions the odometer is the smallest nonnegative one, so its minimum is $0$.
\end{proof}
Fix $x,z \in V$ and let $f(y) = \frac{g^{z}(x,y)}{\deg(y)}$ be the function satisfying $f(z)=0$ and $\Delta f = \delta_z - \delta_x$.
(Here $g^z(x,y)$ is the expected number of visits to $y$ by a random walk started at $x$ before hitting $z$).
With a slight abuse of notation, we define $g(x,y): = \sum_{z \in V} \frac{1}{n} g^{z}(x,y)$.
\begin{proof}[Proof of Proposition~\ref{t.finite}]
Observe that $\sum_{x \in V} s(x) = n$. Therefore $s$ stabilizes to the all $1$ configuration by Lemma~\ref{l.massn}, and the odometer $u$ satisfies
\[ s+ \Delta u = 1 \]
and $\min u = 0$.
Since $\Delta \frac{g^{z}(x,\cdot)}{\deg(\cdot)} = \delta_{z} - \delta_x$, the function
\begin{eqnarray*}
v^{z}(y) &:=& \frac{1}{\deg(y)}\sum_{x \in V} g^{z}(x,y) (s(x)-1)
\end{eqnarray*}
has $\Delta v^{z}(y) = 1-s(y)$ for $y \neq z$ and
\[ \Delta v^{z}(z) = \sum_{x \neq z} (s(x)-1) = 1-s(z). \]
Thus $u-v^z$ is harmonic on $V$ and hence is a (random) constant.
Let $v = \frac{1}{n} \sum_{z \in V} v^{z}$. Since $u-v^z$ is constant for all $z$, the difference $u-v$ is also constant.
Recalling that $g = \frac{1}{n} \sum_{z \in V} g^{z}$, we have $v(y) = \frac{1}{\deg(y)}\sum_{x \in V} g(x,y)(s(x) - 1)$. To compute the covariance of the Gaussian vector $v$, note that
\[ \mathbb{E} [(s(z)-1)(s(w)-1)] = 1_{\{z=w\}} - \frac1n \]
hence
\begin{align*}
\lefteqn{\mathbb{E}[v(x) v(y)] } \\&= \frac{1}{\deg(x)\deg(y)} \sum_{z,w \in V} g(z, x) g(w,y) \mathbb{E}[(s(z) - 1)(s(w) - 1)] \\
& = \frac{1}{\deg(x)\deg(y)} \left(\sum_{z \in V} g(z,x) g(z,y) - \frac{1}{n} \left( \sum_{z \in V} g(z,x)\right) \left( \sum_{w \in V} g(w,y) \right)\right).
\end{align*}
The function $K(y) :=\frac{1}{\deg(y)}\sum_{w \in V} g(w,y)$ has $\Delta K = \sum_{z,w \in V} \frac{1}{n} \left( \delta_z - \delta_w \right) = 0$, so $K$ is a constant. The second term on the right is just $\frac{K^2}{n}$.
Letting $C$ be a $N(0,\frac{K^2}{n})$ random variable independent of $v$, the Gaussian vectors $\eta$ and $(v(x)+C)_{x \in V}$ have the same covariance matrix, so
\[ \eta \eqindist v + C. \]
Since $u-v$ is constant and $\min u = 0$ we conclude that \[ u = v - \min v \eqindist \eta - \min \eta. \qedhere \]
\end{proof}
\begin{table}
\begin{center}
\begin{tabular}{|l|l|l|l|l|l|}
\hline
& $d=1$ & $d=2$ & $d=3$ & $d=4$ & $d\geq 5$ \\[3pt]
\hline
$\mathbb{E}(\eta_\mathbf 0 - \eta_\mathbf x)^2 \lesssim \psi_d(n,r) := $ & $n r^2$ & $r^2 \log \left( \frac{n}{r} \right)$ & $r$ & $\log (1+r) $ & $ 1 $ \\[3pt]
\hline
$\mathbb{E} \max \{ \eta_\mathbf{x} \,:\, \mathbf{x} \in \mathbb{Z}_n^{d} \} \asymp$ & $n^{3/2}$ & $n$ & $n^{1/2}$ & $\log n$ & $(\log n)^{1/2}$ \\[3pt]
\hline
\end{tabular}
\end{center}
\caption{\label{table:orders} Statistics of the bi-Laplacian Gaussian field $\eta$ on the discrete torus $\mathbb{Z}_n^d$. In the first line, $r=\norm{\mathbf x}_2$ and the symbol $\lesssim$ means there is a dimension-dependent constant $C_d$ such that $\mathbb{E} (\eta_\mathbf{0} - \eta_\mathbf{x})^2 \leq C_d \psi_d(n,r)$ for all $\mathbf{x} \in \mathbb{Z}_n^d$. The second line gives the order of the expected value of the maximum of the field
up to a dimension-dependent constant factor.
}
\end{table}
\section{Green function and bi-Laplacian field on $\mathbb{Z}_n^d$}
The rest of the paper is devoted to the proof of Theorem~\ref{p.torus}. Taking Proposition~\ref{t.finite} as a starting point, the expected odometer equals the expected maximum of the bi-Laplacian Gaussian field $\eta$, since
\[ \mathbb{E} u(x) = \mathbb{E} (\eta_x - \min \eta) = - \mathbb{E} \min \eta = \mathbb{E} \max \eta \]
where we have used that $\mathbb{E} \eta_x=0$. From the covariance matrix for $\eta$ we see that
\[ \mathbb{E} (\eta_x - \eta_y)^2 =\sum_{z \in V} \left(\frac{g(z,x)}{\deg(x)}-\frac{g(z,y)}{\deg(y)}\right)^2. \]
We will use asymptotics for the Green function $g$ of the discrete torus $\mathbb{Z}_n^d$ to estimate the right side.
This will enable us to use Talagrand's majorizing measure theorem to determine the order of $\mathbb{E} \max \{\eta_\mathbf{x} \,:\, \mathbf{x} \in \mathbb{Z}_n^d\} $ up to a dimension-dependent constant factor.
These calculations are carried out below and summarized in Table~\ref{table:orders}. The table entries give bounds up to a constant factor depending only on the dimension $d$. For example, the $d=3$ column means that there is a positive constant $C$ such that $\mathbb{E} (\eta_0 - \eta_\mathbf{x})^2 \leq C\norm{\mathbf{x}}_2$ for all $\mathbf{x} \in \mathbb{Z}_n^3$ and $C^{-1} n^{1/2} \leq \mathbb{E} \max \eta \leq C n^{1/2}$.
\begin{rem}\label{r-notation}
For the rest of our work, we identify the discrete torus $\mathbb{Z}_n^d$ with $\left( \mathbb{Z} \cap (-n/2, n/2] \right)^d$ which in turn is viewed as a subset of $\mathbb{R}^d$. For $\mathbf{x} \in \mathbb{Z}_n^d$ and $1 \le p \le \infty$, we denote by $\norm{\mathbf{x}}_p$ the $p$-norm
under the above identification. Note that for standard graph distance $d_G$ on $\mathbb{Z}^d_n$, we have $d_G(\mathbf{0}, \mathbf{x}) = \norm{\mathbf{x}}_1$.
\end{rem}
\subsection{Fourier analysis on the discrete torus}
In this section, we derive a formula for $\mathbb{E}(\eta_{\mathbf{0}} - \eta_{\mathbf{x}})^2$. We begin by recalling some basic facts about Fourier analysis on the discrete torus and the spectral theory of the Laplacian.
We equip the torus $\mathbb{Z}_n^d$ with normalized Haar measure $\mu$ (in other words the uniform probability measure).
Consider the Hilbert space $\mathcal{H}=L^2( \mathbb{Z}_n^d, \mu)$ of complex valued functions on torus with inner product
\[
\langle f ,g \rangle = \int_{\mathbb{Z}_n^d} f\bar{ g} \, d\mu = \frac{1}{n^d} \sum_{x \in \mathbb{Z}_n^d} f(x) \overline{g(x)}.
\]
We identify the Pontryagin dual group $\widehat{\mathbb{Z}_n^d}$ with $\mathbb{Z}_n^d$ as follows. For any $\mathbf{a} \in \mathbb{Z}_n^d$, the map $ \mathbf{x} \mapsto \exp\left( i 2 \pi \mathbf{x}\cdot\mathbf{a}/n \right)$
gives the corresponding element in $\widehat{\mathbb{Z}_n^d}$ (The dot product is the usual Euclidean dot product in $\mathbb{R}^n$). We denote this character by $\chi_\mathbf{a}$.
Recall that $\{ \chi_\mathbf{a} : \mathbf{a} \in \mathbb{Z}_n^d \}$ forms an orthonormal basis for $\mathcal{H}$. Moreover each $\chi_\mathbf{a}$ is an eigenfunction for the Laplacian $\Delta$
with eigenvalue
\[
\lambda_\mathbf{a}= -4 \sum_{i=1}^d \sin^2 \left(\frac{\pi a_i}{n} \right).
\]
Thus the Laplacian $\Delta:\mathcal{H} \to \mathcal{H}$ is a non-positive, bounded operator.
Moreover $\lambda_\mathbf{a} = 0$ if and only if $\mathbf{a}= \mathbf{0}$.
Laplacian $\Delta$ is a self-adjoint operator, that is
\begin{equation} \label{sa}
\langle f_1 , \Delta f_2 \rangle= \langle \Delta f_1, f_2 \rangle
\end{equation}
for all $f_1,f_2 \in \mathcal{H}$ (See Remark \ref{int-p}).
We denote by $g_\mathbf{x}(\mathbf{y})= g(\mathbf{y},\mathbf{x})$. Recall that
\begin{equation} \label{green}
\Delta g_\mathbf{x} = 2d \left( \frac{1}{n^d} \chi_\mathbf{0}- \delta_\mathbf{x}\right).
\end{equation}
Denote by $\widehat{g_\mathbf{x}}(\mathbf{a})$, the Fourier coefficient $\langle g_\mathbf{x}, \chi_\mathbf{a} \rangle$.
Since the function $\mathbf{x} \mapsto \sum_{\mathbf{y} } g_\mathbf{x}(\mathbf{y})$ is harmonic, it is constant. This implies that there exists $L \ge 0$ such that
\begin{equation}\label{ghat0}
\widehat{g_\mathbf{x}}(\mathbf{0})= n^{-d} \sum_{\mathbf{y} \in \mathbb{Z}^d_n} g_\mathbf x(\mathbf y)= L
\end{equation}
for all $\mathbf{x} \in \mathbb{Z}^d_n$ .
For $\mathbf{a} \neq \mathbf{0}$, we have
\begin{equation}\label{ghata}
\lambda_\mathbf{a} \widehat{g_\mathbf{x}}(\mathbf{a})= \lambda_\mathbf{a} \langle g_\mathbf{x}, \chi_\mathbf{a} \rangle = \langle g_\mathbf{x}, \Delta \chi_\mathbf{a} \rangle= \langle \Delta g_\mathbf{x}, \chi_\mathbf{a} \rangle = -2d \langle \delta_\mathbf{x}, \chi_\mathbf{a} \rangle= - 2d n^{-d} \chi_{-\mathbf{a}}(\mathbf x).
\end{equation}
For the above equation, we used $\Delta \chi_\mathbf{a}= \lambda_\mathbf{a} \chi_\mathbf{a}$, equations \eqref{sa}, \eqref{green} and $\langle \chi_\mathbf{0}, \chi_\mathbf{a}\rangle=0$.
By Parseval's theorem and equations \eqref{ghat0}, \eqref{ghata},
\begin{eqnarray}
\nonumber \mathbb{E} (\eta_\mathbf{0}- \eta_\mathbf{x})^2 &=& (2d)^{-2} \sum_{\mathbf{z} \in \mathbb{Z}^d_n} \left(g(\mathbf{z}, \mathbf{0} )- g(\mathbf{z},\mathbf{x})\right)^2 \\
\nonumber &=& (2d)^{-2} n^d \langle g_\mathbf{0} - g_\mathbf{x},g_\mathbf{0} - g_\mathbf{x} \rangle\\
\nonumber & = & (2d)^{-2}n^d \sum_{\mathbf{z} \in \mathbb{Z}^d_n} \abs{\widehat{g_\mathbf{0}}(\mathbf{z}) - \widehat{g_\mathbf{x}}(\mathbf{z}) }^2\\
\label{e.four-form} &=& \frac{1}{4} F_{n,d} (\mathbf x)
\end{eqnarray}
where
\begin{equation} \label{Fnd}
F_{n,d} (\mathbf x) := n^{-d} \sum_{\mathbf z \in \mathbb{Z}^d_n \setminus \{ \mathbf 0 \} } \frac{ \sin^2 \left( \frac{ \pi \mathbf{x}. \mathbf{z}}{n} \right) } { \left( \sum_{i=1}^d \sin^2 \left( \frac{\pi z_i}{n}\right) \right)^2}
\end{equation}
\begin{rem}\label{int-p}
In $\mathbb{R}^d$, we have the Green's second identity \[\int f_1 \Delta f_2 \, dx = -\int \nabla f_1.\nabla f_2 \, dx = \int f_1 \Delta f_2 \, dx \] for all $f_1,f_2 \in C_c^\infty (\mathbb{R}^d)$. Similarly, in our discrete setting we have
\[
\langle f_1 , \Delta f_2 \rangle= - \frac{1}{2 n^d} \sum_{\mathbf x, \mathbf y \in \mathbb{Z}^d_n} (f_1(\mathbf x)-f_2(\mathbf y))(\overline{f_2(\mathbf x)}-\overline{f_2(\mathbf y)})k(\mathbf x,\mathbf y) = \langle \Delta f_1, f_2 \rangle
\]
where $k(\mathbf x, \mathbf y)= \mathbf 1_{x \sim y}$ and $\mathbf x \sim \mathbf y$ if $\mathbf x$ and $\mathbf y$ are neighbors in $\mathbb{Z}^d_n$. See for instance, \cite[Lemma 2.1.2]{S-C} or \cite[Lemma 13.11]{LPW} for a proof.
\end{rem}
Our task now is to estimate the expression $F_{n,d}(\mathbf x)$. Henceforth we assume that $\mathbf{x} \neq \mathbf{0}$.
To study the quantity $\max _{\mathbf{x},\mathbf{y}\in \mathbb{Z}^d_n} (\eta_\mathbf{x} - \eta_\mathbf{y} )$ as $n$ goes to $\infty$,
we want to estimate $\mathbb{E}(\eta_{\mathbf{0}} - \eta_{\mathbf{x}})^2$ with $d$ fixed and $n$ large for different values of $\mathbf{x}$.
We approximate $F_{n,d}(\mathbf x)$ by an integral of a function over $\mathbb{R}^d$.
For $\mathbf w \in \mathbb{R}^d$ and $r >0$, we denote by $B_\infty(\mathbf w,r)$ the open ball with center $\mathbf w$ and radius $r$ under supremum norm, that is
\[
B_\infty(\mathbf w,r) = \{ \mathbf y \in \mathbb{R}^d: \norm{\mathbf y - \mathbf w}_\infty <r \}.
\]
We denote the indicator function of the ball $B_\infty(\mathbf z/n , 1/(2n))$ by $I_{\mathbf z, n}:\mathbb{R}^d \to \{0,1\}$, that is $I_{\mathbf z, n} = \mathbf{1}_{B_\infty(\mathbf z/n , 1/(2n))}$.
Define the function $G_{n,d,\mathbf x} : \mathbb{R}^d \to \mathbb{R}$
\[
G_{n,d,\mathbf x}= \sum_{\mathbf z \in \mathbb{Z}^d_n \setminus \{ \mathbf 0 \} } \frac{ \sin^2 \left( \frac{ \pi \mathbf{x}\cdot \mathbf{z}}{n} \right) } { \left( \sum_{i=1}^d \sin^2 \left( \frac{\pi z_i}{n}\right) \right)^2} I_{\mathbf z, n} .
\]
Since the cubes $B_\infty(\mathbf z/n , 1/(2n))$ are disjoint with volume $n^{-d}$ , we have
\begin{equation} \label{e.feqintg}
F_{n,d}(\mathbf x)= \int_{\mathbb{R}^d} G_{n,d,\mathbf x} (\mathbf y) \,d \mathbf y .
\end{equation}
By triangle inequality, we have
\begin{equation} \label{e.triangle}
(1+ \sqrt{d})^{-1} \norm{\mathbf{z}/n}_2 \le \norm{\mathbf y}_2 \le (1 + \sqrt{d}) \norm{\mathbf z/n}_2
\end{equation}
for all $\mathbf{z} \in \mathbb{Z}^d_n \setminus \{ \mathbf{0} \}$ and for all $\mathbf y \in B_\infty(\mathbf z/n , 1/(2n))$ under the usual identification from Remark~\ref{r-notation}.
We will estimate the function $G_{n,d,\mathbf x}$ using the function $H_{n,d,\mathbf x} : \mathbb{R}^d \to \mathbb{R}$ defined by
\begin{equation*} \label{defH}
H_{n,d,\mathbf x}(\mathbf y) = \sum_{\mathbf z \in \mathbb{Z}^d_n \setminus \{ \mathbf 0 \} } \frac{ \sin^2 \left( \frac{ \pi \mathbf{x} \cdot \mathbf{z}}{n} \right) } { \norm{\mathbf y}_2^4 } I_{\mathbf z, n}(\mathbf y) .
\end{equation*}
More precisely, we have the following lemma.
\begin{lem} \label{compLemma}
Fix $ d \in \mathbb{N}^*$. There exist positive reals $c_1,C_1$ such that
\begin{equation} \label{compGH}
c_1 H_{n,d,\mathbf x}(\mathbf y) \le G_{n,d,\mathbf x}( \mathbf y)\le C_1 H_{n,d,\mathbf x} (\mathbf y)
\end{equation}
for all $n \in \mathbb{N}^*$, for all $\mathbf x \in \mathbb{Z}^d_n \setminus \{ \mathbf 0 \}$ and for all $\mathbf y \in \mathbb{R}^d$.
\end{lem}
\begin{rem} We will use $C_i$ for large constants and $c_i$ for small constants. Here and in what follows, all constants are allowed to depend on $d$ but not on $\mathbf{x} \in \mathbb{Z}^d_n \setminus \{ \mathbf 0 \}$ or
$n$.
\end{rem}
\begin{proof}
The idea is to use the estimate
\[
\frac{2}{\pi} \abs{t} \le \abs{\sin t} \le \abs{t}
\]
for all $t \in [-\pi/2,\pi/2]$. Thus there exists a constant $C_2>0$ such that
\begin{equation} \label{e-cl1}
C_2^{-1} \norm{\mathbf z/n}_2^4 \le \left( \sum_{i=1}^d \sin^2 \left( \frac{\pi z_i}{n}\right) \right)^2 \le C_2 \norm{\mathbf z/n}_2^4
\end{equation}
for all $\mathbf{z} \in \mathbb{Z}^d_n$ and for all $n \in \mathbb{N}^*$. By \eqref{e.triangle}, we have
\begin{equation} \label{e-cl2}
(1+\sqrt{d})^{-4} \frac{I_{\mathbf z, n}(\mathbf y)}{ \norm{ \mathbf z/ n }_2^4} \le \frac{I_{\mathbf z, n}(\mathbf y)}{ \norm{\mathbf y }_2^4} \le (1+\sqrt{d})^4 \frac{I_{\mathbf z, n}(\mathbf y)}{ \norm{ \mathbf z/ n }_2^4}
\end{equation}
for all $\mathbf{z} \in \mathbb{Z}^d_n \setminus \{ \mathbf 0 \}$, for all $\mathbf y \in \mathbb{R}^d$ and for all $n \in \mathbb{N}^*$. Combining equations \eqref{e-cl1} and \eqref{e-cl2} gives \eqref{compGH}.
\end{proof}
By \eqref{e.four-form}, \eqref{e.feqintg} along with integration of \eqref{compGH} over the variable $\mathbf y$, there exists $c_1,C_1>0$ such that
\begin{equation} \label{compFeta}
c_1 d^2 \int_{\mathbb R^d }H_{n,d,\mathbf x}(\mathbf y) \, d\mathbf y \le \mathbb{E} (\eta_\mathbf 0 -\eta_\mathbf x)^2 \le C_1 d^2 \int_{\mathbb R^d} H_{n,d,\mathbf x} (\mathbf y) \, d \mathbf y
\end{equation}
for all $n \in \mathbb{N}^*$ and for all $\mathbf x \in \mathbb{Z}^d_n \setminus \{ \mathbf 0 \}$.
By \eqref{compFeta}, it suffices to estimate $\int_{\mathbb{R}^d}H_{n,d,\mathbf{x}} (\mathbf y) \,d\mathbf y$. Observe that the support of $H_{n,d,\mathbf{x}}$ satisfies
\[
\operatorname{Support}( H_{n,d,\mathbf{x}}) \subseteq B_2(\mathbf 0 , \sqrt d) \setminus B_2( \mathbf 0, 1/(2n))
\]
for all $d,n \in \mathbb{N}^*$ and for all $\mathbf x \in \mathbb{Z}^d_n$,
where $B_2$ denotes open ball with respect to Euclidean norm in $\mathbb{R}^d$.
\subsection{Upper bounds}
Define $\psi_d$ by
\begin{equation} \label{e.psi}
\psi_d(n,r):= \begin{cases}
n r^2 & \mbox{if } d=1 \\
r^2 \log \left( \frac{n}{r } \right) & \mbox{if } d=2 \\
r & \mbox{if } d=3 \\
\log (1+r) & \mbox{if } d=4 \\
1 & \mbox{if } d \ge 5. \end{cases}
\end{equation}
for all $n \in \mathbb{N}^*$ and all $r>0$ along with $\psi_d(n ,0):=0$ for all $d,n \in \mathbb{N}^*$.
The upper bounds for $ \mathbb{E}(\eta_{\mathbf{0}} - \eta_{\mathbf{x}})^2$ is summarized in the following Proposition.
\begin{proposition}\label{p.r1ub}
For each $d \in \mathbb{N}^*$, there exists $C_d >0$ such that
\begin{equation} \label{e.r1ub}
\mathbb{E}(\eta_{\mathbf{0}} - \eta_{\mathbf{x}})^2 \le C_d \psi_d(n, \norm{\mathbf x}_2)
\end{equation}
for all $n \in \mathbb{N}^*$ and for all $\mathbf x \in \mathbb{Z}_n^d$ , where $\psi_d$ is defined by \eqref{e.psi}.
\end{proposition}
\begin{proof}
By \eqref{compFeta} it suffices to find upper bounds for $\int_{\mathbb{R}^d} H_{n,d,\mathbf x}(\mathbf y) \, d \mathbf y$.
The strategy to establish upper bounds for $\int_{\mathbb{R}^d} H_{n,d,\mathbf x}(\mathbf y) \, d \mathbf y$ is to split it into two integrals as
$ \int_{\mathbb{R}^d} = \int_{ 1/(2n) \le \norm{\mathbf y}_2 \le \sqrt{d}/ \norm{\mathbf x}_2} +\int_{ \sqrt{d}/\norm{\mathbf x}_2 < \norm{\mathbf y}_2 \le \sqrt{d} }$.
Note that both the integrals are over non-empty annuli since $1/(2n) \le \sqrt{d}/(4\norm{\mathbf x}_2) \le \sqrt{d}/\norm{\mathbf x}_2 \le \sqrt{d}$ for all $\mathbf{x} \in \mathbb{Z}^d_n \setminus \{ \mathbf{0} \}$ with the identification from
Remark~\ref{r-notation}.
Using Cauchy-Schwarz inequality and the bound $\abs{\sin t } \le t$, we have $\abs{\sin (\pi \mathbf{x}. \mathbf{z/n})} \le \pi \norm{\mathbf x}_2 \norm{\mathbf z/n}_2 $.
This bound competes with the trivial bound $\abs{\sin (\pi \mathbf{x}. \mathbf{z/n}) } \le 1$. It will become clear that up to constants, the first bound is better for the first term
and the trivial bound $\abs{\sin (\pi \mathbf{x}. \mathbf{z/n}) } \le 1$ is better for the second term.
For the first integral we use the bound $\abs{\sin t} \le \abs{t}$ and Cauchy-Schwarz inequality to obtain
\[
H_{n,d,\mathbf x}(\mathbf y) \le \sum_{\mathbf z \in \mathbb{Z}^d_n \setminus \{ \mathbf 0 \} } \frac{\pi^2 \norm{\mathbf{x}}_2^2 \norm{ \mathbf{z}/n}_2^2 } { \norm{\mathbf y}_2^4 } I_{\mathbf z, n}(\mathbf y).
\]
By \eqref{e.triangle}, we have $\norm{\mathbf z/n}_2^2 I_{\mathbf z , n} (\mathbf y) \le (1 +\sqrt{d})^2 \norm{\mathbf y}_2^2 I_{\mathbf z, n}(\mathbf y)$. Therefore, we obtain
\begin{equation} \label{1d-ub}
H_{n,d,\mathbf x}(\mathbf y) \le \sum_{\mathbf z \in \mathbb{Z}^d_n \setminus \{ \mathbf 0 \} } \frac{(1 +\sqrt{d})^2 \pi^2 \norm{\mathbf{x}}_2^2 } { \norm{\mathbf y}_2^2 } I_{\mathbf z, n}(\mathbf y).
\end{equation}
for all $n,d \in \mathbb{N}^*$, for all $\mathbf y \in \mathbb{R}^d$ and for all $\mathbf x \in \mathbb{Z}^d_n \setminus \{ \mathbf 0 \}$.
Hence, we have
\begin{equation} \label{1t-ub}
I_1 := \int_{ 1/(2n) \le \norm{\mathbf y}_2 \le \sqrt{d}/ \norm{\mathbf x}_2} H_{n,d,\mathbf x}(\mathbf y) \, d \mathbf{y} \le (1 +\sqrt{d})^2 \pi^2 \norm{\mathbf x}_2^2 \omega_{d-1} \int_{1/(2n)}^{\sqrt{d}/ \norm{\mathbf x}_2} \frac{r^{d-1}}{r^2} \, dr
\end{equation}
where $\omega_{d-1}=\frac{2 \pi^{d/2} }{\Gamma (d/2)} $ is the $(d-1)$-dimensional surface measure of unit sphere $\mathbb{S}^{d-1}$ in $\mathbb{R}^d$.
For the second integral we use the bound $\abs{\sin t} \le 1$, to obtain
\begin{equation} \label{2t-ub}
I_2:=\int_{ \sqrt{d}/\norm{\mathbf x}_2 < \norm{\mathbf y}_2 \le \sqrt{d} } H_{n,d,\mathbf x}(\mathbf y) \, d \mathbf{y} \le \omega_{d-1} \int_{\sqrt{d}/ \norm{\mathbf x}_2}^ {\sqrt{d}} \frac{r^{d-1}}{r^4} \, dr.
\end{equation}
Combining equations \eqref{1t-ub} and \eqref{2t-ub}, we obtain
\begin{equation} \label{234ub}
\int_{\mathbb{R}^d} H_{n,d,\mathbf x}(\mathbf y) \, d \mathbf y \le (1 +\sqrt{d})^2 \pi^2 \norm{\mathbf x}_2^2 \omega_{d-1} \int_{1/(2n)}^{\sqrt{d}/ \norm{\mathbf x}_2} r^{d-3} \, dr+ \omega_{d-1}\int_{\sqrt{d}/ \norm{\mathbf x}_2}^ {\sqrt{d}} r^{d-5} \, dr.
\end{equation}
The desired upper bounds on $\mathbb{E}(\eta_{\mathbf{0}} - \eta_{\mathbf{x}})^2$ for all dimensions follow from \eqref{234ub} along with \eqref{compFeta}.
\end{proof}
\old{
$\mathbf{ d=1}$.
There exists $C_1>0$ such that
\begin{equation*} \label{ub1}
\mathbb{E}(\eta_{\mathbf{0}} - \eta_{\mathbf{x}})^2 \le C_1 n \norm{\mathbf x}_2^2 .
\end{equation*}
for all $n \in \mathbb{N}^*$ and for all $x \in \mathbb{Z}_n$. \\
$\mathbf{ d=2}$.\\
There exists $C_2>0$ such that
\begin{equation*} \label{ub2}
\mathbb{E}(\eta_{\mathbf{0}} - \eta_{\mathbf{x}})^2 \le C_2 \norm{\mathbf x}_2^2 \log \left( \frac{n}{\norm{\mathbf x}_2 }\right)
\end{equation*}
for all $n \in \mathbb{N}^*$ and for all $\mathbf x \in \mathbb{Z}_n^2$. \\
$\mathbf{ d=3}$.\\
There exists $C_3>0$ such that
\begin{equation*} \label{ub3}
\mathbb{E}(\eta_{\mathbf{0}} - \eta_{\mathbf{x}})^2 \le C_3 \norm{\mathbf x}_2
\end{equation*}
for all $n \in \mathbb{N}^*$ and for all $\mathbf x \in \mathbb{Z}_n^3$. \\
$\mathbf{ d=4}$.\\
There exists $C_4>0$ such that
\begin{equation*} \label{ub4}
\mathbb{E}(\eta_{\mathbf{0}} - \eta_{\mathbf{x}})^2 \le C_4 \left( 1+ \log \norm{\mathbf x}_2 \right)
\end{equation*}
for all $n \in \mathbb{N}^*$ and for all $\mathbf x \in \mathbb{Z}_n^4$. \\
$\mathbf{ d \ge 5}$.\\
There exists $C_5>0$ such that
\begin{equation*} \label{ub5}
\mathbb{E}(\eta_{\mathbf{0}} - \eta_{\mathbf{x}})^2 \le C_5 \omega_{d-1} d^{d/2}.
\end{equation*}
for all $d \in \mathbb{N}^*$ with $d \ge 5$ , for all $n \in \mathbb{N}^*$ and for all $\mathbf x \in \mathbb{Z}_n^d$
}
\begin{rem}
The terms $I_1$ and $I_2$ correspond to the energy (square of 2-norm) of the low and high frequency oscillations of the function $g_\mathbf 0 - g_\mathbf x$ respectively.
For $d=1,2$, the term $I_1$ dominates $I_2$.
For $d=3$, both $I_1$ and $I_2$ are of the same order. For $d \ge 4$, the term $I_2$ dominates $I_1$.
Hence our approach to obtain matching lower bounds in the next subsection is as follows: For $d=1,2$, we obtain lower bounds on lower frequency terms and for $d \ge 3$ we obtain lower bounds on higher frequency terms.
\end{rem}
It is well know that the Gaussian field $\eta$ induces a Hilbert space on $\mathbb{Z}^d_n$ given by the distance metric
\[
d_\eta(\mathbf{x},\mathbf{y}) := \left( \mathbb{E} (\eta_\mathbf{x} - \eta_\mathbf{y})^2 \right)^{1/2}.
\]
The upper bounds on $d_\eta$ provided by Proposition \ref{p.r1ub} transfers to upper bounds on $\mathbb{E} \sup_{\mathbf x \in \mathbb{Z}^d_n} \eta_{\mathbf x}$. The main tool to transfer bounds is
Dudley's bound \cite[Proposition 1.2.1]{Tal} described below. There exists $L>0$ such that
\begin{equation} \label{e.dud}
\mathbb{E} \sup_{\mathbf x \in \mathbb{Z}^d_n} \le L \sum_{k=0}^\infty 2^{k/2} e_k
\end{equation}
where $e_k = \inf \sup_{\mathbf t \in \mathbb{Z}^d_n} d_\eta(\mathbf t,T_k)$ and the infimum is taken over all subsets $T_k \subseteq \mathbb{Z}^d_n$ with $\abs{T_k} \le 2^{2^{k}}$.
\begin{proposition}\label{p.r2ub}
For each $d \in \mathbb{N}^*$, there exists $C_d >0$ such that
\begin{equation} \label{e.r2ub}
\mathbb{E} \sup_{\mathbf x \in \mathbb{Z}^d_n} \eta_{\mathbf x} \le C_d \phi_d(n)
\end{equation}
for all $n \in \mathbb{N}^*$, where $\phi_d$ is defined by \eqref{e.phi}.
\end{proposition}
\begin{proof}
Let $d_G$ denote the standard graph distance on the torus $\mathbb{Z}^d_n$.
By choosing a submesh of appropriate cardinality the following statement is clear:
For any $d \in \mathbb{N}^*$, there exist $C_{d,1} >0$ such that for any $n \ge 2$ and for any $ 2 \le m < n^d$, there exists a set $S_m \subset \mathbb{Z}^d_n$ with $\abs{S_m}=m$ such that
\begin{equation} \label{e.sep}
\sup_{\mathbf t \in \mathbb{Z}_n^d} d_G(t,S_m)= \sup_{\mathbf t \in \mathbb{Z}_n^d} \inf_{\mathbf s \in S_m} d_G(\mathbf t,\mathbf s) \le C_{d,1} \frac{n}{m^{1/d}}.
\end{equation}
For each $d \in \mathbb{N}^*$ by Dudley's bound \eqref{e.dud}, \eqref{e.sep} and Proposition \ref{p.r1ub}, there exists $C_{d,2}, C_{d,3} >0$
\begin{equation} \label{e.ubm}
\mathbb{E} \max_{\mathbf{x} \in \mathbb{Z}^d_n} \eta_x \le C_{d,2} \sum_{k=0}^{\lfloor \log \log n^d \rfloor} 2^{k/2} \left[ \psi_d\left(n, \frac{C_{d,1} n}{2^{(2^k/d)}}\right) \right]^{1/2} \le C_{d,3} \phi_d(n)
\end{equation}
The second inequality above follows from a straightforward case by case calculation.
\end{proof}
\subsection{Lower bounds}
Next, we prove matching lower bounds on $\mathbb{E} (\eta_{\mathbf{0}} - \eta_{\mathbf{x}} )^2$.
For dimensions $d=1,2$, we estimate $F_{n,d}$ directly. \\
\paragraph{$\mathbf{ d=1}$:}
We use the bound $\abs{t} \ge \abs{ \sin t } \ge \frac{2}{\pi} \abs{t}$ for all $\abs{t} \le \pi/2$ to obtain
\[
F_{n,1} (\mathbf x) \ge n^{-1} \frac{\sin^2 \frac{\pi \mathbf x}{n}}{\sin^4 \frac{\pi}{n}} \ge 4 \pi^{-4} n\norm{ \mathbf x}_2^2
\]
for all $n \in \mathbb{N}^*$ and for all $\mathbf x \in \mathbb{Z}_n \setminus \{ \mathbf 0 \}$. Hence there exists
$c_1>0$ such that
\begin{equation} \label{lb1}
\mathbb{E}(\eta_{\mathbf{0}} - \eta_{\mathbf{x}})^2 \ge c_1 n \norm{\mathbf x}_2^2
\end{equation}
for all $n \in \mathbb{N}^*$ and for all $\mathbf x \in \mathbb{Z}_n \setminus \{ \mathbf 0 \}$. \\
\paragraph{$\mathbf{ d=2}$:}
Let $S_{k} \subset \mathbb{Z}^2$ denote the sphere with center $\mathbf 0$ and radius $k$ in the supremum norm, that is $S_{k} = \{ \mathbf y \in \mathbb{Z}^2:\norm{\mathbf y}_\infty =k \}$.
For $x \in \mathbb{R}^2$, we define $H_{\mathbf x}= \{ \mathbf y \in \mathbb{Z}^2: \abs{\mathbf x \cdot \mathbf y} \ge \norm{\mathbf x}_2 \norm{\mathbf y}_2/\sqrt{2} \}$.
It is easy to check that $\abs{S_k} =4k$ and $\abs{S_k \cap H_\mathbf x }\ge 2k$ for all $k \in \mathbb{N}^*$ and for all $\mathbf x \in \mathbb{R}^2$. Let $\alpha \in (1, \sqrt{2})$.
If $\norm{\mathbf z}_2 \le \frac{n}{\alpha \norm{\mathbf x}_2} $ by Cauchy-Schwarz inequality we have $\abs{\mathbf x \cdot \mathbf z/n} \le \alpha^{-1}$.
We need the inequality $\pi \abs{t} \ge \abs{\sin \pi t} \ge \beta \abs{t}$ for all $\abs{t} \le \alpha ^{-1}$ where $\beta= \alpha \sin (\pi \alpha^{-1})$. Putting together the above pieces, we obtain
\begin{eqnarray*}
F_{n,2}(\mathbf x) &\ge& \sum_{k=1}^{\left\lfloor \frac{n}{\alpha \norm{\mathbf x}_2} \right\rfloor \wedge \lfloor n/4 \rfloor} \sum_{\mathbf z \in S_k } \beta^2 \pi^{-4} \norm{\mathbf z}_2^{-4} \abs{\mathbf x \cdot \mathbf z}^2 \\
& \ge & \frac{\beta^2}{2 \pi^4} \norm{\mathbf x}_2^2 \sum_{k=1}^{\left\lfloor \frac{n}{\alpha \norm{\mathbf x}_2} \right\rfloor \wedge \lfloor n/4 \rfloor} \sum_{\mathbf z \in S_k \cap H_\mathbf x}\norm{\mathbf z}_2^{-2} \\
& \ge & \frac{\beta^2}{4 \pi^4} \norm{\mathbf x}_2^2 \sum_{k=1}^{\left\lfloor \frac{n}{\alpha \norm{\mathbf x}_2} \right\rfloor\wedge \lfloor n/4 \rfloor} \sum_{\mathbf z \in S_k \cap H_\mathbf x} k^{-2} \\
& \ge & \frac{\beta^2}{2 \pi^4} \norm{\mathbf x}_2^2 \sum_{k=1}^{\left\lfloor \frac{n}{\alpha \norm{\mathbf x}_2} \right\rfloor\wedge \lfloor n/4 \rfloor} k^{-1} \\
& \ge & \frac{\beta^2}{2 \pi^4} \norm{\mathbf x}_2^2\log \left( \left( \left\lfloor \frac{n}{\alpha \norm{\mathbf x}_2} \right\rfloor \wedge \lfloor n/4 \rfloor \right)+1 \right) \\
& \ge & \frac{\beta^2}{2 \pi^4} \norm{\mathbf x}_2^2\log \left( \frac{n}{\alpha \norm{\mathbf x}_2} \wedge \frac{n}{4} \right).
\end{eqnarray*}
Since $\norm{\mathbf x}_2 \le n/\sqrt{2}$ for all $\mathbf x \in \mathbb{Z}^2_n$, we have the desired lower bound by using $\alpha < \sqrt{2}$. That is, there exists
$c_2>0$ such that
\begin{equation} \label{lb2}
\mathbb{E}(\eta_{\mathbf{0}} - \eta_{\mathbf{x}})^2 \ge c_2 \norm{\mathbf x}_2^2\log \left( \ \frac{n}{ \norm{\mathbf x}_2} \right)
\end{equation}
for all $n \in \mathbb{N}^*$ with $n \ge 4$ and for all $\mathbf x \in \mathbb{Z}^2_n \setminus \{ \mathbf 0 \}$.
\\
\paragraph{$\mathbf{d \geq 3}$:} For dimensions $d \ge 3$, we will approximate $H_{n,d,\mathbf x}$ by its almost everywhere point-wise limit $H_{\infty,d,\mathbf x}$ defined by
\[
H_{\infty,d,\mathbf x} (\mathbf y) = \frac{ \sin^2( \pi \mathbf x \cdot \mathbf y) }{ \norm{\mathbf y}_2^4} \mathbf 1_{B_\infty( \mathbf 0,1/2) }(\mathbf y).
\]
for $ y \neq \mathbf 0$ and $0$ otherwise.
Therefore we would like to estimate integrals of the form
\begin{equation} \label{e-hd1}
\int_{r_1 \le \norm{\mathbf y}_2 \le r_2} \frac{ \sin^2( \pi \mathbf x \cdot \mathbf y) }{ \norm{\mathbf y}_2^4} \, d\mathbf y= \int_{r_1}^{r_2} r^{d-5} s_d(r \norm{\mathbf x}_2 ) \, d r
\end{equation}
where $s_d(t) = \int_{\mathbb{S}^{d-1}} \sin^2(\pi t y_1 ) \, \nu_{d-1}(d\mathbf y)$ and $\nu_{d-1}$ denotes the surface measure in $\mathbb{S}^{d-1}$. We will need the following lower bound for $s_d$.
\begin{lem}\label{lem-sine} Fix $ d\in \mathbb{N}^*$ with $d \ge 3$. Then
for all $\epsilon >0$, there exists $\delta>0$ such that $s_d(t) \ge \delta$ for all $t \ge \epsilon$.
\end{lem}
\begin{proof}
Since $s_d:\mathbb{R} \to \mathbb{R}$ is a continuous function with $s_d(t) > 0$ for all $t \neq 0$, it suffices to show that $\mathop{\rm lim\,inf}\limits_{t \to \infty} s_d(t) >0$.
By \cite[Corollary 4]{BGMN} (See Remark \ref{cor-4}(b)),
\[
s_d(t) = c_d \int_{-1}^1 (1-x^2)^{(d-3)/2} \sin^2 (\pi t w) \, dw \ge c_d 2^{(3-d)/2} \int_{-1/2}^{1/2} \sin^2 (\pi t w) \, dw
\]
where $c_d$ is a constant that depends on $d$.
Since $\lim_{t \to \infty } \int_{-1/2}^{1/2} \sin^2 (\pi t w) \, dw = 1/2$, the conclusion follows.
\end{proof}
\begin{rem} \label{cor-4}
\begin{itemize}
\item[(a)] Recall that we used the point-wise bound $\abs{\sin t} \le 1$ to obtain upper bounds on $I_2$. We want to somehow reverse that inequality to obtain corresponding lower bounds. Although the reverse inequality $\abs{\sin t} > \delta$ is not true for any $\delta>0$ in a pointwise sense,
it is true in an average sense. That is the content of Lemma \ref{lem-sine}.
\item[(b)] \cite[Corollary 4]{BGMN} implies the following striking result in geometric probability: Let $d \ge 3$. For a uniformly distributed random vector $\mathbf y=(y_1,y_2,\ldots,y_d)$ in the $(d-1)$-dimensional unit sphere $\mathbb{S}^{d-1}$ in $\mathbb{R}^d$, the projection
$(y_1,y_2, \ldots, y_{d-2})$ is uniformly distributed in the $(d-2)$-dimensional unit ball $\mathbb{B}^{d-2}= B_2(\mathbf 0 , 1)$ in $\mathbb{R}^{d-2}$.
\item[(c)] Since $\lim_{n \to \infty} H_{n,d,\mathbf x}= H_{\infty , d, \mathbf x}$ almost everywhere, one might wonder if we can prove matching lower bounds for $I_2$ using dominated convergence theorem. This approach gives a lower
bound as $n$ goes to $\infty$ but with both $d$ and $\mathbf x$ fixed. However we want lower bounds with fixed $d$ and with both $n$ and $\mathbf x$ varying. Hence there is a need to quantify this convergence as both $n$ and $\mathbf x$ varies.
We fulfill this need in Lemma \ref{lem-conv}.
\end{itemize}
\end{rem}
One can easily check that $\lim_{n \to \infty} H_{n,d,\mathbf x}= H_{\infty , d, \mathbf x}$ almost everywhere. We need the following quantitative version of this convergence.
\begin{lem}\label{lem-conv}
Fix $d \in \mathbb{N}^*$.
For any $\epsilon >0$, there exists positive reals $\delta, N $ such that
\[
\abs{H_{n,d,\mathbf{x}}(\mathbf y) - \frac{ \sin^2( \pi \mathbf x \cdot \mathbf y) }{ \norm{\mathbf y}_2^4}} \le \frac{\epsilon}{\norm{ \mathbf y}_2^4}
\]
for all $n \ge N$, for all $\mathbf x \in \mathbb{Z}^d_n \setminus \{ \mathbf 0 \}$ with $\norm{\mathbf x}_2< \delta n$ and for almost every $y \in B_2(\mathbf 0,1/4) \setminus B_2 (\mathbf 0, 1/(8 \norm{\mathbf x}_2))$.
\end{lem}
\begin{proof}
Note that, we have the inclusion
\[
B_\infty (\mathbf 0,1/(2n)) \subset B_2( \mathbf 0, 1/(8 \norm{\mathbf x}_2)) \subset B_2(\mathbf 0 , 1/4)
\]
for all $n \ge 4 \sqrt{d}$ and for all $\mathbf x \in \mathbb{Z}^d_n \setminus \{ \mathbf 0 \}$ with $\norm{ \mathbf x}_2 \le n/ 4 \sqrt{d}$.
We use $\norm{\mathbf x}_2 \ge 1$ and the comparison of norms $\norm{\mathbf w}_2 \le \sqrt{d} \norm{\mathbf w}_\infty$ to prove the above inclusions.
The function $\mathbf y \mapsto \sin^2(\pi \mathbf x \cdot \mathbf y)$ has gradient bounded uniformly in 2-norm by $\pi \norm{\mathbf x}_2$. Hence we have
\[
\abs{H_{n,d,\mathbf{x}}(\mathbf y) - \frac{ \sin^2( \pi \mathbf x \cdot \mathbf y) }{ \norm{\mathbf y}_2^4}} \le \pi \norm{\mathbf x}_2 \frac{\sqrt{d}}{n} \norm{\mathbf y}_2^{-4}
\]
for all $n \ge 4 \sqrt{d}$ and for all $\mathbf x \in \mathbb{Z}^d_n \setminus \{ \mathbf 0 \}$ with $\norm{ \mathbf x}_2 \le n/ 4 \sqrt{d}$ and for almost every $y \in B_2(\mathbf 0,1/4) \setminus B_2 (\mathbf 0, 1/(8 \norm{\mathbf x}_2))$.
The choice $\delta = \min\left( \frac{1}{4 \sqrt{d}}, \frac{\epsilon}{ \pi \sqrt{d}}\right)$ and $N= 4 \sqrt{d}$ satisfies all the requirements.
\end{proof}
We put together the above pieces to obtain the following lower bound for $d \ge 3$.
\begin{lem}\label{lem-highd-lb}
Fix $d \ge 3$. There exists positive reals $\delta,N, c_d$ such that
\[
\int_{\mathbb{R}^d} H_{n,d,\mathbf x} (\mathbf y) \, d\mathbf y \ge c_d \int_{1/(8 \norm{\mathbf{x}}_2)}^{1/4} r^{d-5} \, dr
\]
for all $n \ge N$ and for all $\mathbf x \in \mathbb{Z}^d_n \setminus \{ \mathbf 0 \}$ with $\norm{\mathbf x}_2< \delta n$.
\end{lem}
\begin{proof}
By Lemma \ref{lem-sine}, there exists $\epsilon_1>0$ such that $s_d(t) \ge 2 \epsilon_1$ for all $t \ge 1/8$. By Lemma \ref{lem-conv}, there exists positive reals $\delta, N $ such that
\[
\abs{H_{n,d,\mathbf{x}}(\mathbf y) - \frac{ \sin^2( \pi \mathbf x \cdot \mathbf y) }{ \norm{\mathbf y}_2^4}} \le \frac{\epsilon_1 }{\omega_{d-1} \norm{ \mathbf y}_2^4}
\]
for all $n \ge N$, for all $\mathbf x \in \mathbb{Z}^d_n \setminus \{ \mathbf 0 \}$ with $\norm{\mathbf x}_2< \delta n$ and for almost every $y \in B_2(\mathbf 0,1/4) \setminus B_2 (\mathbf 0, 1/(8 \norm{\mathbf x}_2))$.
Combining the above observations, we have for all $n \ge N$ and for all $\mathbf x \in \mathbb{Z}^d_n \setminus \{ \mathbf 0 \}$ with $\norm{\mathbf x}_2< \delta n$
\begin{eqnarray*}
\int_{\mathbb{R}^d} H_{n,d,\mathbf x} (\mathbf y) \, d\mathbf y & \ge & \int_{1/(8\norm{\mathbf{x})}_2 \le \norm{\mathbf y}_2 \le 1/4} H_{n,d,\mathbf x} (\mathbf y) \, d\mathbf y \\
& \ge & \int_{1/(8\norm{\mathbf{x})}_2 \le \norm{\mathbf y}_2 \le 1/4} ( H_{\infty,d,\mathbf x} (\mathbf y) - \epsilon_1 \omega_{d-1}^{-1} \norm{\mathbf y}_2^{-4})\, d\mathbf y \\
& = & \int_{1/(8 \norm{\mathbf{x}}_2)}^{1/4} r^{d-5} (s_d(r \norm{\mathbf x}_2) - \epsilon_1) \, dr \\
& \ge & \epsilon_1 \int_{1/(8 \norm{\mathbf{x}}_2)}^{1/4} r^{d-5} \, dr . \qedhere
\end{eqnarray*}
\end{proof}
We now establish the following lower bounds corresponding to the upper bounds in Proposition \ref{p.r1ub}.
\begin{proposition} \label{p.r1lb}
For each $d \in \mathbb{N}^*$, there exists positive reals $\delta_d,N_d,c_d$ such that
\[
\mathbb{E}(\eta_{\mathbf{0}} - \eta_{\mathbf{x}})^2 \ge c_d \psi_d(n, \norm{\mathbf x}_2)
\]
for all $n \ge N_d$ and for all $\mathbf x \in \mathbb{Z}^d_n \setminus \{ \mathbf 0 \}$ with $\norm{\mathbf x}_2< \delta_d n$, where $\psi_d$ is defined by \eqref{e.psi}.
\end{proposition}
\begin{proof}
The cases $d=1,2$ follow from \eqref{lb1} and \eqref{lb2} respectively. The case $d \geq 3$ follows from
Lemma \ref{lem-highd-lb} along with \eqref{compFeta}.
\end{proof}
\begin{rem}\label{rem-sketch}
The condition $\norm{\mathbf x}_2 < \delta_d n$ that appears in the lower bound for the case $d \ge 3$ is somewhat unsatisfactory.
We believe that the lower bound is true without any such an additional condition.
However the lower bounds in the present form are good enough for our main application.
\old{\item[(b)]
Recall that the reason behind the estimation of $\mathbb{E} (\eta_\mathbf 0 - \eta_\mathbf x)^2 $ is to estimate the quantity
$M_{d}(n)=\mathbb E \max_{\mathbf x} \eta_{\mathbf x}$. Upper bounds on $\mathbb{E} (\eta_\mathbf 0 - \eta_\mathbf x)^2 $
translate to upper bounds on $M_d(n)$ by Talagrand's equation 1.12. Suppose there is a method to give lower bounds $M_{d}(n) \ge l_d(n)$ (for some function $l_d$) using \emph{strong} lower bounds on $\mathbb{E} (\eta_\mathbf 0 - \eta_\mathbf x)^2 $ without
the $\norm{x}_2 < \delta n $ condition. Then the same method can be used to give lower bounds on $M_{d}(n)$ using our \emph{weaker} lower bounds. We simply use the inequality
\[
M_{d}(n) \ge \mathbb E \max_{\mathbf x,\mathbf y \in \mathbb{Z}^d_n \cap B_\infty( \mathbf 0 , \frac{\delta}{2 \sqrt{d} } n)} \eta_{\mathbf x} \ge l_d (\delta n/2 \sqrt d).
\]
This will give the same lower bounds assuming $l_d(n)$ and $l_d (\delta n/2 \sqrt d)$ are of the same order as a function of $n$.
\end{itemize} }
\end{rem}
Next, we obtain lower bounds matching the upper bounds in Proposition \ref{p.r2ub}. We start by recalling notation and setup for Talagrand's majorizing measure \cite[Theorem 2.1.1]{Tal}.
We consider centered multivariate Gaussian random variables $(\eta_t)_{t \in T}$ indexed by a set $T$ with cardinality $\abs{T}$. An \emph{admissible sequence} $\{ \mathcal{A}_k \}$ is an increasing sequence of partitions of $T$ such that $|\mathcal{A}_k| \leq 2^{2^{k}}$.
Here ``increasing sequence" refers to the fact that every set in $\mathcal{A}_{n+1}$ is contained in a set in $\mathcal{A}_n$.
We denote by $A_k(t)$ the unique element of $\mathcal{A}_n$ that contains $t \in T$. Recall that $d_\eta(t_1,t_2) = \left( \mathbb{E} (\eta_{t_1}-\eta_{t_2})^2 \right)^{1/2}$ denotes
the Hilbert space metric induced by $(\eta_t)_{t \in T}$. We define the function
\[ \gamma_2 (T,d_\eta) = \inf \sup_{t \in T} \sum_{k=0}^\infty \operatorname{diam}_\eta(A_n(t))\]
where $\operatorname{diam}_\eta$ denotes the diameter in the $d_\eta$ metric and the infimum is taken over all admissible sequences.
The majorizing measure theorem \cite[Theorem 2.1.1]{Tal} states that there is
some universal constant $L$ for which
\begin{equation} \label{e.majormeas}
\frac{1}{L} \gamma_2(T, d) \leq \mathbb{E} \sup_{t \in T} \eta_t \leq L \gamma_2(T, d).
\end{equation}
\begin{proposition} \label{p.r2lb}
For each $d \in \mathbb{N}^*$, there exists $c_d >0$ such that
\begin{equation} \label{e.r2lb}
\mathbb{E} \sup_{\mathbf x \in \mathbb{Z}^d_n} \eta_{\mathbf x} \ge c_d \phi_d(n)
\end{equation}
for all $n \in \mathbb{N}^*$, where $\phi_d$ is defined by \eqref{e.phi}.
\end{proposition}
\begin{proof}
The strategy is to use the lower bound for $d_\eta$ given by Proposition \ref{p.r1lb} along with \eqref{e.majormeas} where $T= \{ \mathbf x \in \mathbb{Z}^d_n: \norm{\mathbf{x}}_2 < \delta_d n \} $ and
$\mathbb{E} \sup_{t \in \mathbb{Z}^d_n} \eta_t \ge \mathbb{E} \sup_{t \in T} \eta_t $. Note that it suffices to show \eqref{e.r2lb} for large enough $n$, \emph{i.e.} $n > N_d$ for some fixed $N_d$.
For $d=1,2,3$, by \eqref{e.majormeas} we have
\begin{equation} \label{e.llb0}
\mathbb{E} \sup_{t \in \mathbb{Z}^d_n} \eta_t \ge L^{-1} \inf \sup_{t \in T} \operatorname{diam}_\eta (A_0(t)).
\end{equation}
Since $\abs{ \mathcal{A}_0} \le 2$, we have $ \sup_{t \in T}\operatorname{diam}_G (A_0(t)) \ge c_0 n$ for some $c_0>0$. Therefore by Proposition \ref{p.r1lb} along with \eqref{e.llb0} we obtain the desired result.
For $d=4$, by \eqref{e.majormeas} we have
\begin{equation}\label{e.llb1}
\mathbb{E} \sup_{t \in \mathbb{Z}^d_n} \eta_t \ge L^{-1} \inf \sup_{t \in T} 2^{k/2} \operatorname{diam}_\eta (A_k(t)).
\end{equation}
where $k = \left\lfloor\log_2 \log_2 \abs{T}\right\rfloor-1$. This gives $2^{k/2} \ge c_0 \sqrt{\log n}$ for some $c_0 >0$. Moreover, $k = \left\lfloor\log_2 \log_2 \abs{T}\right\rfloor-1$ and $\abs{\mathcal{A}_k} \le 2^{2^{k}}$
implies that at least one of the sets $A_k(t)$ has cardinality greater than or equal to $\sqrt{\abs{T}}$, which in turn implies $\operatorname{diam}_G(A_k(t)) \ge c_1 \sqrt{n}$ for some $c_1>0$.
By Proposition \ref{p.r1lb}, we obtain $\operatorname{diam}_G(A_k(t)) \ge c_2 \sqrt{\log n}$ for some $c_2>0$ and for large enough $n$. The conclusion for $d=4$ then follows from \eqref{e.llb1}.
The case $d \ge 5$ is a direct consequence of Sudakov minoration (\cite[Lemma 2.1.2]{Tal}).
\end{proof}
\begin{proof}[Proof of Proposition \ref{p.torus}]
The upper and lower bounds follow from Propositions \ref{p.r2ub} and \ref{p.r2lb}.
\end{proof}
\section*{Acknowledgment}
We thank Daqian Sun for proofreading part of an early draft, and the referee for suggestions that improved the paper.
|
\section{Introduction}
The exploration of underlying mechanism for anomalous and normal heat conduction in low dimensional systems represents a huge challenge in the area of statistical physics~\cite{Lepri1997prl,Lepri2003pr,Dhar2008ap,Liu2013epjb}. After enormous efforts for more than one decade from numerical simulations\cite{Hu1998pre,Tong1999prb,Hatano1999pre,Tsironis1999pre,Sarmiento1999pre,Dhar1999prl,Alonso1999prl,Hu2000pre,Aoki2000pla,Li2001prl,Dhar2001prl,Aoki2001prl,Zhang2002pre,
Li2002prl,Saito2003epl,Savin2003pre,Lepri2003pre,Segal2003jcp,Gendelman2004prl,Li2005chaos,Zhang2005jcp,Zhao2005prl,
Zhao2006prl,He2008pre,Shao2008pre,Dubi2009pre,Henry2009prb,Saito2010prl,Wang2010prl,Yang2010nt,Wang2011epl,Wang2012pre,Xiong2012pre,
Landi2013pre,Xiong2014pre,Das2014pre,Mendl2014pre,Savin2014pre,Liu2014prb,Wang2015pre}, theoretical predictions\cite{Lepri1998epl,
Lepri1998pre,Narayan2002prl,Wang2004prl,Cipriani2005prl,Pereira2006prl,Delfini2006pre,Basile2006prl,Beijeren2012prl,
Mendl2013prl,Pereira2013pre,Spohn2014jsp,Liu2014prl} and experimental observations\cite{Chang2008prl,Xu2014nc,Meier2014prl}, there is still no consensus for the exact physical mechanism causing anomalous heat conduction. It is believed that momentum conservation plays an important role in determining the actual heat conduction behavior. In general, 1D non-integrable lattices with momentum conserving property should have anomalous heat conduction where the thermal conductivity $\kappa$ diverges with the lattice size $N$ as $\kappa\propto N^{\alpha}$ where $0<\alpha<1$~\cite{Lepri1997prl,Lepri2003pr,Dhar2008ap,Liu2013epjb}. However, the 1D coupled rotator lattice is a well known exception. It exhibits normal heat conduction behavior despite its momentum conserving nature~\cite{Giardina2000prl,Gendelman2000prl}.
The normal heat conduction in 1D coupled rotator lattice was discovered via numerical simulations by two groups independently~\cite{Giardina2000prl,Gendelman2000prl}. In order to understand the underlying mechanism, the temperature dependence of thermal conductivity has been studied in detail in both works. In Ref. \cite{Giardina2000prl}, the temperature dependence of thermal conductivity of coupled rotator lattice has been found to be like $\kappa(T)\propto e^{\Delta V/T}$ where $\Delta V$ is a positive constant. However, in Ref. \cite{Gendelman2000prl}, a temperature dependence of thermal conductivity of $\kappa(T)\propto e^{-T/A}$ was claimed where $A$ is a positive constant. There it was also argued that a possible phase transition at temperature around $T\sim 0.2-0.3$ where heat conduction is normal above this temperature while anomalous below this temperature exists~\cite{Gendelman2000prl}. Although this phase transition statement was challenged as a finite size effect~\cite{Yang2005prl}, a later work supported this phase transition conjecture after deriving a similar temperature dependent thermal conductivity as $\kappa(T)\propto e^{-T/A}$~\cite{Pereira2006prl}.
Most recently, the 1D coupled rotator lattice re-attracts much attention due to the new finding of simultaneously existing normal diffusion of momentum as well as heat energy~\cite{Li2014arxiv}. It is argued that the normal behavior might be due to the reduced number of conserved quantities which is unique for periodic interaction potentials~\cite{Das2014arxiv,Spohn2014arxiv}. But the new debate is that whether the stretch is conserved or not in 1D coupled rotator lattice. The investigation of 1D coupled rotator lattice will be the key to unravel the true mechanism behind the connection between momentum conservation and normal or anomalous heat conduction. Therefore, it is the right time to revisit the temperature dependence of thermal conductivity of 1D coupled rotator lattice as a first step.
In this work, we will revisit the temperature dependence of thermal conductivities for the 1D coupled rotator lattice. We find that the temperature dependence is neither $\kappa(T)\propto e^{\Delta V/T}$ nor $\kappa(T)\propto e^{-T/A}$ as previously claimed~\cite{Giardina2000prl,Gendelman2000prl}. The actual temperature dependence is a power-law dependence as $\kappa(T)\propto T^{-3.2}$. The possible connection with the momentum diffusion of single kicked rotator or the Chirikov standard map has also been discussed. In order to determine whether there is a phase transition, we also present the temperature dependent thermal conductance at different system sizes. All the thermal conductances for different sizes collapse to the same value at low temperatures while approach to the power-law behavior as $\kappa(T)\propto T^{-3.3}$ at high temperatures. However, the crossover temperature decreases as the system size increases. This fact indicates that there is no phase transition between normal and anomalous heat conduction. In thermodynamical limit, the heat conduction is normal for all temperatures except the trivial zero temperature point. In Sec. II the lattice model will be introduced. Numerical results and discussions will be presented in Sec. III and the results will be summarized in Sec. IV.
\section{Model}
The Hamiltonian for the 1D coupled rotator lattice is defined as the following~\cite{Giardina2000prl,Gendelman2000prl}:
\begin{equation}\label{ham}
H=\sum^{N}_{i=1}\left[\frac{p^2_i}{2}+K[1-\cos{(q_{i+1}-q_i)}]\right]
\end{equation}
where $q_i$ and $p_i$ denote the displacement from equilibrium and momentum for $i$-th atom, respectively. The mass of the atom $m$ and the Boltzmann constant $k_B$ has been set into unity. The parameter $K$ with energy dimension represents the coupling strength of the inter-atom potential. Therefore, the system temperature $T$ can be rescaled by $T/K$ and we can also set $K=1$ for simplicity~\cite{Li2012rmp}. The equations of motion for $i$-th atom are
\begin{eqnarray}\label{equation-lattice}
\dot{q}_i&=&p_i \nonumber\\
\dot{p}_i&=&K\sin{(q_{i+1}-q_i)}-K\sin{(q_i-q_{i-1})}
\end{eqnarray}
At low temperature limit, the displacements are small values so that $|q_{i+1}-q_i|\ll 1$. The Hamiltonian can be expanded into the Taylor series as:
\begin{eqnarray}\label{rotator-low-T}
H&=&\sum^{N}_{i=1}\left[\frac{p^2_i}{2}+\frac{(q_{i+1}-q_i)^2}{2}\right.\nonumber\\
&-&\left.\frac{(q_{i+1}-q_i)^4}{4!}+\frac{(q_{i+1}-q_i)^6}{6!}-...\right]
\end{eqnarray}
This Hamiltonian will approach to the integrable Harmonic lattice only at zero temperature $T=0$. The first stable nonlinear potential term will be the sextic potential.
On the other hand, at infinite high temperature limit $T=\infty$, the rotator lattice will approach to another integrable system consisting of $N$ independent and isolated free particles as
\begin{equation}
H=\sum^{N}_{i=1}\frac{p^2_i}{2}
\end{equation}
This is because the kinetic energy is proportional to the temperature as $\left<p^2_i\right>=T$ and the potential energy in Eq. (\ref{ham}) is confined by the cosine function.
Without getting into numerics, we can get a qualitative picture of the thermal conductivity of 1D coupled rotator lattice as the function of temperature. The thermal conductivity $\kappa(T)$ will diverge in the low temperature limit approaching to the harmonic limit and will decay to zero in the high temperature limit as approaching to the unconnected $N$ free particles. As a general picture, the thermal conductivity $\kappa(T)$ will decrease as the temperature increases.
\begin{figure}
\includegraphics[width=\columnwidth]{T-profile.eps}
\vspace{-0.5cm} \caption{\label{fig:T-profile}
(color online). Temperature profiles for the 1D coupled rotator lattice at (a) $T=0.1$ and (b) $T=0.8$. The lattice sizes are $N=50,100,200,400,800$ and $1600$ and the color rule of the lines are the same for both (a) and (b). The first and last atom are put into contact with a Langevin heat bath with temperature set as $T_{L/R}=T(1\pm\Delta)$ where $\Delta=0.1$ here.}
\end{figure}
In non-equilibrium numerical simulations, we put the first and last atom into contact with Langevin heat bath. The temperatures for left and right heat bath are set as $T_{L/R}=T(1\pm \Delta)$ where $T$ denotes the average temperature and $\Delta = 0.1$ gives rise to the temperature gradient along the lattice. The fixed boundary conditions with $q_0=q_{N+1}=0$ are also been applied.
\section{Results and Discussions}
Before we discuss the results of thermal conductivities, we first show the temperature profiles. In Fig. \ref{fig:T-profile}, the temperature profiles at two different temperatures $T=0.1$ and $T=0.8$ are plotted. The lattice sizes are chosen as $N=50,100,200,400,800$ and $1600$. For relative low temperature at $T=0.1$, the temperature jumps at two boundaries are obvious. However, the temperature jumps are reduced with the increase of lattice size $N$ as can be seen in Fig. \ref{fig:T-profile}(a). In Fig. \ref{fig:T-profile}(b) where the temperature is relatively high at $T=0.8$, all the temperature profiles collapse to the same straight line as the temperature jumps are very small for all lattice sizes.
In order to obtain the temperature dependence of thermal conductivities, we need first define the way how $\kappa(T)$ can be calculated numerically. We notice that the temperature profiles all collapse to the same straight line if the temperature jumps can be ignored at high temperatures or long lattice sizes. It is appropriate to define the thermal conductivity $\kappa(T)$ as:
\begin{equation}\label{kappa-def}
\kappa(T)=\frac{JN}{T_L-T_R}
\end{equation}
where $J=\left<J_i\right>$ is the average heat flux along the lattice in the stationary state and the local heat flux $J_i$ is defined as $J_i=-p_i K \sin{(q_{i+1}-q_i)}$ via the energy continuity equation.
\begin{figure}
\includegraphics[width=\columnwidth]{kappa.eps}
\vspace{-0.5cm} \caption{\label{fig:kappa}
(color online). Thermal conductivities $\kappa(T)$ as the function of temperature for different lattice sizes $N=50,100,200,400,800$ and $1600$. The straight line is proportional to $T^{-3.2}$ which describes the temperature behavior of $\kappa(T)$ at high temperature region very well.}
\end{figure}
In Fig. \ref{fig:kappa}, the thermal conductivities $\kappa(T)$ as the function of temperature are plotted for different lattice sizes $N=50,100,200,400,800$ and $1600$. At high temperature region, all the thermal conductivities $\kappa(T)$ for different lattice sizes collapse together indicating the saturation of thermal conductivities as the increase of lattice size. This is characteristic for lattices with normal heat conduction. As the temperature decreases, the thermal conductivities first increases and then becomes flat. This is because the phonon mean free paths are getting longer as the temperature reduces and the ballistic regime will be approached if the phonon mean free paths are longer than the lattice size. The boundary jumps will dominate the temperature profiles as in Fig. \ref{fig:T-profile}(a) and the definition of $\kappa(T)$ of Eq. (\ref{kappa-def}) will no longer reflect the actual thermal conductivities.
At high temperatures, it is clearly seen that the thermal conductivities follows a power-law dependence as $\kappa(T)\propto T^{-3.2}$. For the longest size we considered here as $N=1600$, this power-law behavior can be best fitted for more than two orders of magnitudes for the $\kappa(T)$ value as shown in Fig. \ref{fig:kappa-T_1600}. This also explains the poor fitting of the $\kappa(T)\propto e^{\Delta V/T}$ dependence in Ref. \cite{Giardina2000prl} and the $\kappa(T)\propto e^{-T/A}$ dependence in Ref. \cite{Gendelman2000prl}.
\begin{figure}
\includegraphics[width=\columnwidth]{kappa-T_1600.eps}
\vspace{-0.5cm} \caption{\label{fig:kappa-T_1600}
(color online). Thermal conductivities $\kappa(T)$ as the function of temperature for lattice size $N=1600$. The data are taken from Fig. \ref{fig:kappa} and can be best fitted to be a power-law dependence as $\kappa(T)\propto T^{-3.2}$.}
\end{figure}
This power-law dependence of $\kappa(T)\propto T^{-3.2}$ cannot be explained by the effective phonon theory which is able to predict the temperature dependent thermal conductivities for other typical 1D nonlinear lattices such as FPU-$\beta$ lattice and generalized nonlinear Klein-Gordon lattices~\cite{Li2006epl,Li2007epl,Li2007pre,Li2009jpsj,Li2010prl,Li2012aip,Li2013pre,Yang2014pre}. According to the effective phonon theory, the thermal conductivity for low temperature rotator lattice with Hamiltonian of Eq. (\ref{rotator-low-T}) can be derived as
\begin{equation}\label{kappa-ept}
\kappa(T)\propto \frac{1}{\epsilon} \propto T^{-2}
\end{equation}
where $\epsilon$ is the nonlinearity strength with the following temperature dependence
\begin{equation}
\epsilon\propto \frac{\left<(q_{i+1}-q_i)^6\right>}{\left<(q_{i+1}-q_i)^2\right>}\propto T^2
\end{equation}
at low temperature region. Here we consider the sextic potential term as the lowest nonlinear term because the dynamics governed by the negative quartic potential term is unstable. Therefore, the actual temperature behavior of $\kappa(T)\propto T^{-3.2}$ is steeper than the prediction of $\kappa(T)\propto T^{-2}$ from effective phonon theory.
Although it is difficult to unravel the exact physical mechanism behind the power-law dependence of thermal conductivities for coupled rotator lattice, it is very helpful to look into the transport properties of the single kicked rotator (the Chirikov standard map)~\cite{Chirikov1979pr}. As the name indicates, the coupled rotator lattice is a kind of connected kicked rotators. The equations of motion for the single kicked rotator are
\begin{eqnarray}\label{equation-single}
p_{n+1}-p_{n}&=& K \sin{(q_n)}\nonumber\\
q_{n+1}-q_{n}&=& p_{n+1}
\end{eqnarray}
where $q_n$ and $p_n$ denote the coordinate and momentum after $n$-th kick.
\begin{figure}
\includegraphics[width=\columnwidth]{Dp_T.eps}
\vspace{-0.5cm} \caption{\label{fig:Dp}
(color online). Momentum diffusion constant $D_P$ as the function of temperature for 1D coupled rotator lattice. The numerical data are obtained via equilibrium MD simulations as in Ref. \cite{Li2014arxiv}. The solid line of $T^{-3.2}$ is guided for the eye.}
\end{figure}
The variation of momentum $p$ is unbounded and can be characterized by a normal diffusion~\cite{Chirikov1979pr,MacKay1984pd}
\begin{equation}
\left<\Delta p^2(t)\right>\sim D t
\end{equation}
where $D$ is the diffusion constant. This is similar to the normal momentum diffusion for 1D coupled rotator lattice.
According to Ref. \cite{Chirikov1979pr,MacKay1984pd}, the diffusion constant $D$ depends on the coupling strength $K$ as
\begin{eqnarray}\label{dif-single}
D &\propto& (K-1.2)^3,\,\,\, 1.2<K<4 \\
D &\propto& K^2,\,\,\,K>4
\end{eqnarray}
where $1.2$ is the chaos threshold for the kicked rotator.
If one assume the transport properties of coupled rotator lattice are the same as that of single kicked rotator, one would expect that the energy diffusion constant $D_E$ as well as the momentum diffusion constant $D_P$ of coupled rotator lattice should also follow the same dependence as that of single kicked rotator as in Eq. (\ref{dif-single}). To translate the $K$ dependence into $T$ dependence, we notice that the parameter $K$ in single kicked rotator of Eq. (\ref{equation-single}) plays the same role as that in the coupled rotator lattice as in Eq. (\ref{equation-lattice}). As we have discussed above, the dynamics of coupled rotator lattice can be rescaled with $T/K$. The $K$ dependence should be inversely proportional to the $T$ dependence. Low $K$ value region corresponds to the high $T$ region and vice verse. And for normal heat conduction, the thermal conductivity $\kappa$ is proportional to the energy diffusion constant $D_E$. This will finally give rise to the prediction of temperature dependent thermal conductivity $\kappa(T)$ of coupled rotator lattice from the analogy with single kicked rotator:
\begin{eqnarray}
\kappa(T) &\propto& T^{-2},\,\,\,\mbox{low} \,\,T\\
\kappa(T) &\propto& T^{-3},\,\,\,\mbox{high} \,\,T
\end{eqnarray}
As a consistence check, we plot the momentum diffusion constant $D_P$ for the coupled rotator lattice as the function of temperature in Fig. \ref{fig:Dp}. The same temperature dependence of $D_P\propto T^{-3.2}$ has been obtained.
Therefore, the prediction of $\kappa(T) \propto T^{-3}$ with analogy to the single kicked rotator at high temperature or low $K$ region is close to the numerical observation of $\kappa(T) \propto T^{-3.2}$. This indicates there might be some deeper connection between the diffusion behavior of single kicked rotator and 1D coupled rotator lattice. In addition, the analogy analysis also predicts a $\kappa(T) \propto T^{-2}$ behavior at low temperature region which is the same as the prediction from the effective phonon theory of Eq. (\ref{kappa-ept}). The reason why this temperature behavior cannot be observed in numerical simulations might be due to the severe finite size effect at low temperature region as can be seen in Fig. \ref{fig:T-profile}(a).
\begin{figure}
\includegraphics[width=\columnwidth]{conductance.eps}
\vspace{-0.5cm} \caption{\label{fig:conductance}
(color online). Heat conductance $\sigma$ as the function of temperature $T$ for different lattice size $N=50,100,200,400,800$ and $1600$. All the other parameters are the same as in Fig. \ref{fig:kappa}. The left-pointing arrow represents the trend of decreasing crossover temperature as the lattice size increases.}
\end{figure}
In the final part we will discuss the issue about the possible phase transition between normal and anomalous heat conduction for 1D coupled rotator lattice. As we have shown in Fig. \ref{fig:kappa-T_1600}, the actual temperature dependence of $\kappa(T)$ is a power-law dependence as $\kappa(T) \propto T^{-3.2}$. This indicates the analytic derivation of $\kappa(T)\propto e^{-T/A}$ in Ref. \cite{Pereira2006prl} originally for a lattice with pinned on-site potential can not be extended to 1D coupled rotator lattice.
On the other hand, the claim that the transition temperature is around $(0.2-0.3)$ in Ref. \cite{Gendelman2000prl} has been challenged in Ref. \cite{Yang2005prl} with numerical simulations of longer sizes. The possible phase transition could be a finite size effect. Here we reconfirm the finite size effect by giving a more illustrative picture in Fig. \ref{fig:conductance}. The heat conductance $\sigma\equiv\kappa(T)/N$ as the function of temperature has been plotted for different lattice sizes. At high temperature regions, all the heat conductances follow the same temperature behavior of $\sigma\propto T^{-3.2}$. As the temperature decreases, the phonon mean free path will overcome the lattice size and the heat conductance will saturate to a value determined by the harmonic limit of coupled rotator lattice. As can be seen from Fig. \ref{fig:conductance}, the heat conductance for short lattice will first bend and become flat as the temperature decreases. This crossover temperature decreases as the lattice size increases smoothly and no sign of phase transition can be observed. It is also noticed that the crossover temperature happens to be around $(0.2-0.3)$ for lattice sizes of a few thousands.
\section{Summary}
We have systematically investigated the temperature dependence of thermal conductivities of 1D coupled rotator lattice. The actual temperature dependence is a power-law dependence of $\kappa(T)\propto T^{-3.2}$ which is different with the observations of previous studies. The possible connection with the single kicked rotator or the Chirikov standard map has been discussed where a $\kappa(T)\propto T^{-3}$ dependence can be implied. Our results also reconfirm that the previously claimed possible phase transition should be a finite size effect.
\section{acknowledgments}
The numerical calculations were carried out at Shanghai Supercomputer Center. This work has
been supported by the NSF China with grant No. 11334007 (Y.L., N.L., B.L.), the NSF China with grant No. 11347216 (Y.L), Tongji University under Grant No. 2013KJ025 (Y.L), the NSF China with Grant No. 11205114 (N.L.), the Program for New Century Excellent Talents
of the Ministry of Education of China with Grant No. NCET-12-0409 (N.L.) and the
Shanghai Rising-Star Program with grant No. 13QA1403600 (N.L.).
\bibliographystyle{apsrev4-1}
|
\section*{Abstract}
\baselineskip 18pt
\section{Introduction}
\label{sec:intro}
The production-inventory problem we study is that of matching supply and demand
over a planning horizon of $N$ periods, with the provision that a set of supply
decisions must be made before the first period.
For example, the production-inventory planning horizon might represent a
``selling season'' and the decisions to match supply with demand are made before
the season starts.
This is not merely a quantity decision --- how much
to stock up so as to fulfill demand over the entire season; rather,
it is more like a capacity planning decision --- how to set the
supply or production rates as a function of time over the horizon, vis-\`a-vis
the demand forecast.
We consider this production-inventory matching problem within the context of both
lost-sales and backorder inventory models.
More precisely, consider the following general supply and demand matching problem.
Suppose $D_1+\cdots +D_n$ denotes the cumulative demand up to period
$n=1,2,\dots, N$; and assume $\mbox{\sf Var} (D_n)=\sig^2$ for all $n$.
Further suppose the set of supply decisions before the first period takes
the form $\ex[D_1]+\cdots + \ex[D_n]+\kappa\sig n^\al$, where $\al$ and
$\kappa$ are decision variables such that $\al\in [0,1]$ while $\kappa$
can be either positive or negative.
This assumed form allows a fair amount of flexibility:
one can choose to over-supply (positive $\kappa$), thus
creating safety stock, if the lost-sales penalty is high;
or to under-supply (negative $\kappa$), if the production cost
or holding cost is high.
We are especially interested in asymptotic results.
When $n$ is large, demand will approach a normal distribution, with
a standard deviation equal to $\sig\sqrt{n}$.
Shall we match supply with demand by setting $\al=\frac{1}{2}$?
What is the corresponding lost-sales quantity?
What is the impact on lost sales if $\al$ takes on other values?
What is the right value for $\kappa$;
in particular, should it be positive or negative?
We note that this general supply and demand matching problem arises in a wide
variety of production-inventory systems.
As such, there is another way to interpret the asymptotics associated with
large values of $N$ in our model.
Instead of viewing this as a prolonged planning horizon, we can view the latter
as having a fixed length, but with the demand being scaled up.
In other words, the original total demand for the entire season, $D$, now
becomes $D_1+\cdots +D_N$, with the $D_n$ being independent copies of $D$.
The capacity problem continues to consist of making the set of supply decisions
for the entire horizon in advance, with the corresponding deliveries received
every period to satisfy the demand.
Indeed, this alternative view is closer to the class of production-inventory
applications we have in mind --- a high-volume, fast-moving supply and demand
context, which requires an understanding of the asymptotic behavior of key
system performance measures under various supply strategies, as well the
pre-planning of such strategies.
As a specific business analytics instance of such capacity problems often encountered within
the context of workforce management for large services providers, we note
that any significant workforce actions aimed at altering a planned resource
supply usually require long lead times, whereas customer demand can exhibit
volatility and uncertainty over relatively short periods of time.
Competition within the marketplace and the dynamics within these companies
create constant pressure for growth; under this pressure, workforce capacity
planning routinely examines resource requirements when the demand forecast
is scaled up in different proportions.
These considerations in part motivate the current problem setting.
Another application of a similar flavor concerns capacity provisioning in
cloud computing environments where high-frequency demand having considerable
uncertainty needs to be fulfilled with a predetermined supply of resources.
Any changes in these supply decisions are costly and should be avoided.
Now, let $S_n$ denote the net demand at period $n=1,\ldots,N$, namely the
difference between demand and supply (both in terms of the cumulative quantities)
up to period $n$.
In the backorder model, where unmet demand in any period will be backlogged
and supplied by future supplies, backorder and inventory are, respectively,
the positive and negative parts of $S_n$; as such, the model is quite tractable.
On the other hand, the lost-sales model, in which unmet demand in any period will be lost,
is much harder to solve because the cumulative lost-sales quantity, $L_n$, is the running
maximum of $S_n$: $L_n=\max_{0\le j\le n}\{S_j\}$ (with $S_0:=0$), which depends on $S_1, S_2,\ldots, S_n$.
When supply is linear, i.e., $\al=1$, then $S_n$ is a random walk and $L_n$ is
a well-studied object; in particular, $\ex[L_n]$ is well-known and very
accessible.
For other values of $\al$, however, $\ex[L_n]$ does not appear to have explicit
expressions.
The only possible exception is $\al=\frac{1}{2}$, in which case one can approximate
$S_n$ by a Brownian motion with square-root drift.
We derive such an analysis by relating this Brownian motion to the hitting
time of an Ornstein-Uhlenbeck process, for which the density function can be
obtained by inverting Laplace transforms.
Given this level of difficulty, our approach consists of constructing upper-
and lower-bounds for $\ex[L_N]$, and analyzing the asymptotic behavior of
these bounds for large $N$.
Our findings for such asymptotic behaviors are summarized as follows.
When $\kappa >0$ (the over-supply case), setting $\al$ within $(\frac{1}{2},1]$ will
result in a bounded $\ex[L_N]$, whereas $\al \in [0,\frac{1}{2}]$ will cause $\ex[L_N]$
to grow on the order of $\sqrt {N}$.
When $\kappa < 0$ (the under-supply case), $\ex[L_N]$ will grow on the order of
$N^\al$ upon setting $\al$ within $(\frac{1}{2},1]$; whereas it will grow on the order
of $\sqrt {N}$ when $\al \in [0,\frac{1}{2}]$.
In terms of minimizing the total cost over both production cost and lost-sales
penalty, our results imply that setting $\al=\frac{1}{2}$ is the only meaningful choice
leading to non-trivial results, independent of the values of the penalty and holding costs.
Hence, what remains is to find the best value of $\kappa$.
We first obtain an asymptotically optimal solution through a
Brownian approximation of the objective function for large $N$.
Then, for each finite $N$, we examine optimal solutions through the upper
and lower bounds of $\ex[L_N]$ as surrogates for the original problem.
Both bounds lead to simple optimal solutions, and our results include performance
guarantees for both problems as approximations to the original problem.
We also show that these results are readily extended to include holding costs.
In addition, we establish the equivalence between the lost-sales model and
the backorder model when both have the same penalty cost that goes to infinity.
\subsection{Related Work}
A brief review of the related literature is in order.
To the extent that our set of supply decisions has to be made before the
season starts, our model resembles the newsvendor model.
There are important distinctions, however.
To start with, the newsvendor decision is a stocking quantity decision, whereas
we are concerned with the supply {\it dynamics} over time --- all $N$ periods
of the planning horizon.
Consequently, the lost-sales quantity in the newsvendor model is a single number that
can be accounted for at the end of the horizon.
In our model, lost sales build up over the horizon, as demand and supply are
realized and evolve.
Granted, under our cost model which minimizes the total production cost and
lost-sales penalty over the horizon, the lost-sales quantity $\ex[L_N]$ is
associated with the end of the horizon, as in the newsvendor model.
However, we consider inventory holding costs as well the backorder alternative
to lost sales, features that are not present in the newsvendor model.
The lost sales problem is known to have a much more complicated structure than
the problem of backorder, and hence it is more difficult to analyze.
Regarding the optimal policy, the exact description is not known even in the
simplest case.
Karlin and Scarf~\cite{KarlinScarf} demonstrate that the base-stock policies
are not optimal even for systems with a lead time of one period.
Monotonicity properties of the optimal ordering policy are derived by
Morton~\cite{Morton}, Zipkin~\cite{zipkin} and most recently
Huh~\cite{HuhOR10}, each through different methods.
Huh {\it et al.}~\cite{HuhMS09} show that, asymptotically as the ratio of
unit penalty cost and unit holding cost grows to infinity, the difference
in performance between a particular base-stock policy and the optimal policy
will eventually vanish.
Thus, base-stock policies can be regarded as asymptotically optimal when there
is such a significant imbalance between the penalty cost and holding cost.
Levi {\it et al.}~\cite{levi} study production and lost sales over multiple
periods focusing on the performance of a simple heuristic policy that balances
the inventory holding cost and lost-sales penalty in each period, with the aim
of establishing its robustness relative to the optimal policy in a very general
context.
The setting of Levi {\it et al.} is dynamic programming: in each period a
different decision can be made and actions taken accordingly.
In contrast, we allow only a single set of decisions at the beginning of the
horizon;
and all subsequent costs due to production, lost-sales or backlog, and holding
inventory are consequences of this set of decisions.
The way we quantify our heuristic solution (from the upper-bound problem),
however, appears to have a similar flavor to the proven ``2-approximation''
status of the dual-balancing rule in \cite{levi}.
Specifically, the performance of our heuristic, via the ratio of the upper-
and lower-bound solutions, is ``2-plus'' --- where this slightly worse
performance guarantee than a 2-approximation algorithm can be attributed to the
fact that there is no recourse over the planning horizon in our model, which,
to make matters worse, can be infinitely long.
A complementary set of results for the lost-sales model has been recently
obtained by Goldberg {\it et al.}~\cite{goldberg}.
Another body of research related to our analysis concerns process flexibility;
see the recent paper of Chou {\it et al.}~\cite{chou} and the references
therein.
Specifically, process flexibility can be modeled by a random walk with
functional drift, called a generalized random walk in~\cite{chou}.
When proper functional forms are taken, it can be seen that the cost structure
under full flexibility has the same form as the cost structure of our backorder
model, while the cost structure under a so-called chaining option corresponds
to the structure of our lost-sales model.
Upon applying the methodology developed in this paper, the effectiveness of
chaining can be demonstrated and quantified for more general processes.
\subsection{Paper Organization}
The remainder of the paper is organized as follows.
In Section~\ref{sec:prelim}, we start with a formulation of our models ---
while focusing on the lost-sales model, we also cover the backorder model,
followed by a derivation of the main performance measures associated with
these models.
We then present preliminaries regarding the normal distribution function
and the loss function, and derive estimates for $\ex[L_N]$ in the case of
$\al=1$.
In Section~\ref{sec:bounds}, we continue with estimates of $\ex[L_N]$ for all
other cases, through constructing upper- and lower-bounds and also
deriving a corresponding Brownian approximation for the case of $\al=1/2$.
Optimization models that minimize total cost over production, lost sales
(or backorder) and inventory are studied in Section~\ref{sec:optimality},
followed by concluding remarks in Section~\ref{sec:conclusions}.
\section{Problem Formulation and Preliminaries}
\label{sec:prelim}
\subsection{Key Performance Measures}
Consider a planning horizon that consists of $N$ periods,
indexed by $n=1,\dots, N$.
For each period $n$, let $D_n$ denote the demand, a random variable;
and $x_n$ the production or supply quantity, a decision variable.
At the beginning of the planning horizon, we need to determine the
production quantities for all $N$ periods, as part of the capacity
planning problem.
Once this set of decisions is made, then in each period $n$ there will
be $x_n$ units available along with any units left over from previous
periods to supply demand $D_n$.
In the lost-sales model,
any demand surpassing the available supply is lost by the end of each period;
whereas in the backorder model, any unfilled demand will be
backlogged, to be supplied in a later period.
Correspondingly, the quantities of interest include the following.
\begin{itemize}
\item $L_n$: the cumulative demand
shortfall, in the lost-sales model, up to period $n$;
\item
$B_n$: the backlogged demand, in the backorder model, at the end of each
period $n$;
\item
$H_n$ and $H'_n$: the on-hand inventory left at the end of each period $n$,
respectively for the lost-sales and backorder models.
\end{itemize}
\begin{pro}
\label{pro:lost}
{\rm
Define $S_0:=0$ and $S_n:=D_1-x_1+\cdots +D_n-x_n$.
Further define $x^+:=\max\{x,0\}$ and $x^-:=-\min\{x,0\}$.
Then, for $n=1,\dots, N$, we have
\begin{eqnarray}
B_n= S_n^-, &\qquad& H'_n=S_n^+, \label{back}\\
L_n=\max_{0\le j\le n}\{S_j\}, &\qquad& H_n = L_n-S_n.
\label{lost}
\end{eqnarray}
}
\end{pro}
{\startb Proof.}
In the backorder model, any unmet demand (if $S_n>0$) or any leftover
inventory (if $S_n<0$) at the end of each period $n$ will be carried over;
hence, leading to the expressions in (\ref{back}).
The $H_n$ expression in (\ref{lost}) follows from
\begin{eqnarray*}
H_n=\sum_{i=1}^n x_i -\left(\sum_{i=1}^n D_i-L_n\right),
\end{eqnarray*}
i.e., total production minus total {\it supplied} demand equals inventory.
For the $L_n$ expression in (\ref{lost}), note that the result
holds trivially for $n=1$: $L_1=\max\{0,D_1-x_1\}$.
Suppose it holds for $n$. Then,
we have $L_n=S_k$ for some $k\le n$,
which means that there will be no more shortage for periods $k+1$ through $n$.
Therefore, the units available to supply $D_{n+1}$ consist of $x_{n+1}$ and
the leftover from periods $k+1$ through $n$, namely
$$-(D_{k+1}-x_{k+1} + \cdots +D_n-x_n).$$
(Note that the above expression is non-negative, as the quantity in parentheses
cannot exceed $0$ due to no shortage from periods $k+1$ through $n$.)
Hence, at the end of period $n+1$, shortage will occur if
$$D_{n+1}- [x_{n+1} - (D_{k+1}-x_{k+1} + \cdots +D_n-x_n)] >0 ,$$
in which case the cumulative shortage will be adding $S_k$ to the above expression,
with the sum being equal to $S_{n+1}$.
Otherwise, i.e., if the left-hand side above is $\le 0$, then
the cumulative shortage stays at $S_k$.
In summary, we have
\begin{eqnarray*}
L_{n+1}= \max\{S_k,S_{n+1}\}= \max\{L_n, S_{n+1}\}= \max_{0\le j\le n+1}\{S_j\},
\end{eqnarray*}
which is as desired.
\hfill$\Box$
\medskip
In what follows, we shall focus on $L_n$, since $H_n$ relates directly to $L_n$;
whereas $B_n$ and $H'_n$ are simply the positive and negative parts of the
random walk $S_n$, and thus both are easily accessible -- refer to (\ref{esn}) below.
Since the study of large $N$ asymptotics is our primary objective,
we focus on the normal distribution for the demand model. Specifically, let
$$D_n=\mu_n+\sig Z_n,$$
where $Z_n$ are independent and identically distributed (i.i.d.) standard
normal random variables, and $\mu_n$ and $\sig$ are the mean and standard
deviation of $D_n$, respectively.
The common $\sig$ can be interpreted as a forecast error; hence, it is
independent of the periods.
Alternatively, we can allow a period-dependent $\sig_n$, as long as the
sum $\sum_{i=1}^n\sig_i \sim \sig\sqrt{n}$, i.e., it grows at a rate of
$\sqrt n$.
In this case, the above expression will still serve as a good approximation
due to the central limit theorem.
On the production side, we write, for $ n=1,\dots, N$,
\begin{eqnarray}
\label{supply}
x_1+\cdots +x_n=\mu_1+\cdots +\mu_n +\kappa\sig n^\alpha,
\end{eqnarray}
where $\kappa$ and $\al \in [0,1]$ are two policy parameters
(or decision variables).
Note that the (cumulative) production quantity in \eqref{supply} consists of
two parts: the first part matches the mean of the (cumulative) demand; and
the second part can be interpreted as {\it safety stock} to offset
demand variability.
The parameter $\alpha$ is often referred to as the ``safety factor''.
This factor addresses the issue of how the safety stock should match up with
the demand variability, where the latter grows on the order $\sqrt n$.
Recall that $\kappa$ can take on positive or negative values:
when it is positive, $\kappa\sig n^\alpha$ is truly the safety stock;
when $\kappa$ is negative, this part is a calculated under-production,
suitable for settings where the production cost is much higher relative to
the lost-sales penalty.
Since (\ref{supply}) implies that the production quantity for each period $n$
is given by
\begin{eqnarray}
\label{supply1}
x_n=\mu_n +\kappa\sig [n^\alpha- (n-1)^\alpha], \qquad n=1,\dots, N ,
\end{eqnarray}
then the cumulative shortage process $L_n$ can be estimated in different ways
depending on the value of $\al$.
\subsection{Useful Facts about the Standard Normal Variable}
\label{sec:calculations}
We summarize below some useful expressions and estimates associated
with the standard normal variable. Let $Z$
denote the standard normal variate with density and distribution functions
denoted by $\phi (x)$ and $\Phi (x)$, respectively.
Define the ``shortfall function'' (or ``loss function'') as
\begin{eqnarray}
\label{g}
G(x):=\ex[(Z-x)^+]=\int_x^{+\infty}(z-x)\phi(z)dz =\phi(x)-x\bar\Phi(x) ,
\end{eqnarray}
where $\bar\Phi (x):=1-\Phi (x)$ and
the last equality follows directly from $\phi'(z)=-z\phi(z)$.
Now, we list properties of $G(x)$ that will be useful later for our purposes.
\begin{itemize}
\item[1.]
$G(x)$ is decreasing and convex in $x$ (since $(Z-x)^+$ is decreasing and convex in $x$).
\item[2.]
The part of $G(x)$ that has significant curvature is limited to the neighborhood of
the origin $x=0$. For large $x >0$,
we have $G(x){\buildrel\dot\over =} 0$ and $G(-x){\buildrel\dot\over =} x$.
\item[3.]
A direct derivation yields:
\begin{eqnarray}
\label{g_int}
2\int_a^b G(x)dx = \Phi(b) -\Phi(a)+bG(b)-aG(a).
\end{eqnarray}
\item[4.]
\begin{eqnarray}
\label{upperlower2}
0\le G(x)
\le \frac{\phi(x)}{x^2}, \qquad \forall x>0 .
\end{eqnarray}
\item[5.]
\begin{eqnarray}
\label{upperlower2asy}
\frac{x^2 G(x)}{\phi(x)} \to 1 \qquad {\rm as}\; x\to +\infty.
\end{eqnarray}
\end{itemize}
To verify \eqref{upperlower2} and \eqref{upperlower2asy}, observe that
\begin{eqnarray}
\label{upperlower}
\Big(1-\frac{1}{x^2}\Big)\frac{1}{x}\phi (x)
\le \bar\Phi(x)\le \frac{1}{x}\phi (x), \qquad x>0 ,
\end{eqnarray}
where the upper bound follows from
\begin{eqnarray*}
\int_x^\infty e^{-u^2/2} du
\le \int_x^\infty \Big(1+\frac{1}{u^2}\Big)e^{-u^2/2} du
= \frac{1}{x}e^{-x^2/2}
\end{eqnarray*}
and the lower bound follows from
\begin{eqnarray*}
\int_x^\infty e^{-u^2/2} du
\ge \int_x^\infty \Big(1-\frac{3}{u^4}\Big)e^{-u^2/2} du
= \Big(1-\frac{1}{x^2}\Big)\frac{1}{x}e^{-x^2/2}.
\end{eqnarray*}
The (second) inequality in \eqref{upperlower2} follows immediately
from the lower bound in (\ref{upperlower}).
Combining the upper and lower bounds in (\ref{upperlower}),
we have
$\bar\Phi(x){\buildrel\dot\over =} \frac{1}{x}\phi (x)$ for large (positive) $x$.
This, along with l'H\^opital's rule, helps to verify \eqref{upperlower2asy}.
\subsection{Estimating $\mathbf{\ex[L_N]}$ for $\mathbf{\alpha=1}$}
\label{sec:one}
When $\alpha=1$ we have, from (\ref{supply1}),
$x_n=\mu_n +\kappa\sig$; thus, $D_n-x_n=-\kappa\sig +\sig Z_n$.
In this case, the partial sums $\{S_n\}$ constitute a random walk,
and $\ex[L_n]$ follows from standard results concerning the maximum of a random walk.
Specifically, by Spitzer's Identity
(see \cite{spitzer}; also, Ross \cite{ross}, Proposition 7.1.5):
\begin{eqnarray}
\label{sp}
\ex[L_N] = \sum_{n=1}^N \frac{1}{n} \ex[S_n^+],
\end{eqnarray}
where $S_n \, {\buildrel\rm d\over =} \, -\kappa\sig n +\sig \sqrt{n} Z$.
Hence,
\begin{eqnarray}
\label{esn}
\ex[S_n^+]=\sig\sqrt{n}\ex[(Z-\kappa\sqrt{n})^+]
=\sig\sqrt{n} G (\kappa\sqrt{n}).
\end{eqnarray}
Approximating the summation by integration, we have
\begin{eqnarray}
\label{sumint}
\sum_{n=1}^N \frac{1}{n} \ex[S_n^+]{\buildrel\dot\over =}
\int_0^N \frac{1}{t} \ex[S_t^+] dt
=\s \int_0^N \frac{1}{\sqrt{t}}G (\kappa\sqrt{t}) dt
=\frac{2\s}{\kappa} \int_0^{\kappa\sqrt{N}} G (u) du .
\end{eqnarray}
Therefore, combining the above with (\ref{g_int}), we obtain
\begin{eqnarray}
\label{eln}
\ex[L_N] {\buildrel\dot\over =}
\frac{\s}{\kappa} \Big(\Phi(\kappa\sqrt{N})-\frac{1}{2}\Big)+
\s\sqrt{N}G(\kappa\sqrt{N}) .
\end{eqnarray}
When $\kappa >0$, the above is increasing in $N$.
As $N\to\infty$, the second term on the right-hand side vanishes, and the
first term converges to $\frac{\s}{2\kappa}$.
When $\kappa <0$, the second term on the right-hand side of (\ref{eln})
quickly becomes $\sig |\kappa| N$ as $N$ grows, and the first term approaches
a constant $\frac{\s}{2|\kappa|}$.
These results then can be summarized in the following proposition.
\begin{pro}
\label{pro:al1}
{\rm
For $\alpha =1$ and large $N$, we have
\begin{eqnarray}
\label{al1}
\ex[L_N] {\buildrel\dot\over =} \frac{\s}{2\kappa}, \quad{\rm when}\quad \kappa >0;
\qquad
\ex[L_N] {\buildrel\dot\over =} \frac{\s}{2|\kappa|} +\sigma |\kappa| N, \quad{\rm when}\quad \kappa <0 .
\end{eqnarray}
}
\end{pro}
\medskip
Hence, setting $\alpha=1$, the expected shortage is bounded by a constant,
or grows linearly in $N$, corresponding to a positive or negative $\kappa$,
respectively.
\section{Estimating $\mathbf{\ex[L_N]}$ for $\mathbf{\alpha\in [0,1)}$}
\label{sec:bounds}
When $\alpha <1$, it is generally a difficult problem to calculate the exact
value of $\ex[L_N]$, and thus we develop bounds to reveal its asymptotic order
as $N\rightarrow \infty$.
We shall initially focus on $\kappa >0$ in the next two subsections, and
then summarize the corresponding results for $\kappa \le 0$ in the following
subsection.
Lastly, we consider a Brownian approximation for the case when $\alpha = \frac12$.
\subsection{Case of $\mathbf{\alpha \in (\frac12, 1)}$ and $\mathbf{\kappa > 0}$}
From (\ref{lost}), taking into account $S_0=0$, we have
$$L_n=\max_{0\le j\le n}\{S_j\}= \max_{1\le j\le n}\{S_j^+\}, \qquad n=1,\dots, N.$$
Jensen's inequality implies
\begin{eqnarray}
\label{elnlb}
\ex[L_N] \ge \max_{1\le n\le N}\{\ex[S_n^+]\},
\end{eqnarray}
where, similar to (\ref{esn}), we obtain
\begin{eqnarray}
\label{esn1}
\ex[S_n^+]=\sig\sqrt{n}\ex[(Z-\kappa {n}^{\alpha -1/2})^+]
=\sig\sqrt{n} G (\kappa {n}^{\alpha -1/2}) .
\end{eqnarray}
From (\ref{upperlower2}), we know that
\begin{eqnarray*}
\sig\sqrt{n} G (\kappa {n}^{\alpha -1/2}) \le
\frac{\sig\sqrt{n}\phi(\kappa {n}^{\alpha -1/2})}{\kappa^2 {n}^{2\alpha -1}}
:=\frac{\sig\sqrt{n}\phi(u)}{u^2},
\qquad{\rm where}\qquad u:=\kappa {n}^{\alpha -1/2} .
\end{eqnarray*}
Hence, $\ex[S_n^+]$ will decrease to $0$
(since $\phi(u)$ decreases to $0$ exponentially fast as $u\to+\infty$).
The values of $n$ that achieve the maximum in the lower bound of (\ref{elnlb})
must satisfy the following
optimality equation (whose solution need not be unique):
$$\frac{1}{2} n^{-1/2} G (y)= n^{1/2} \bar\Phi(y)\kappa (\alpha -1/2)n^{\alpha -3/2}.$$
This simplifies to
\begin{eqnarray}
\label{opty0}
G (y)= (2\alpha -1)y\bar\Phi(y)
\qquad{\rm or}\qquad
\phi (y)= 2\alpha y\bar\Phi(y) ,
\end{eqnarray}
and from (\ref{upperlower}) we know that $\phi(y){\buildrel\dot\over =} y\bar\Phi(y)$ when
$y$ is large.
Since $\alpha >\frac{1}{2}$, the solution to the above equation exists.
Indeed, from (\ref{upperlower}), we have
$$y\bar\Phi(y)\ge \Big(1-\frac{1}{y^2}\Big)\phi(y), \qquad y>0 .$$
If we replace $y\bar\Phi(y)$ in (\ref{opty0}) by this lower bound, then the resulting solution
\begin{eqnarray}
\label{y0}
\bar y:= \Big(\frac{1}{2\alpha -1}\Big)^{\frac{1}{2}}
\end{eqnarray}
is an upper bound to the solution of the equations in (\ref{opty0}),
since $\phi (y)$ is decreasing in $y$ for $y >0$.
Therefore, in this case we obtain
\begin{eqnarray}
\label{elnlb2}
\ex[L_N]\ge \max_{0\le n\le N}\{\ex[S_n^+]\}{\buildrel\dot\over =}
\sig\Big(\frac{y}{\kappa}\Big)^{1/(2\alpha -1)} G(y),
\end{eqnarray}
where $y$ is the solution to (\ref{opty0}) and we know $y\in (0, \bar y)$.
For an upper bound, we have
$L_N =\max_{0\le n\le N} \{S_n\} \le \sum_{n=1}^N S_n^+$, which, along with
\eqref{esn1},
leads to
\begin{eqnarray}
\label{cal0}
\ex[L_N] \le \sigma \sum_{n=1}^N \sqrt{n} G(\kappa n^{\al - \frac12})
{\buildrel\dot\over =} \sigma \int_0^N \sqrt{t} G(\kappa t^{\al - \frac12}) dt.
\end{eqnarray}
Using a transformation of variable $u= t^{\al -\frac12}$, together with
$G(x) \le \phi(x) /x^2$ from (\ref{upperlower2}), we obtain
\begin{eqnarray}
\label{cal}
\int_0^N \sqrt{t} G(\kappa t^{\al - \frac12}) dt
&=&
\frac{2}{2\al-1} \int_0^{N^{\al -\frac12}} u^{\frac{4-2\al}{2\al -1}} G(\kappa u) du
\nonumber\\
&\le& \frac{2}{(2\al-1)\kappa^2} \int_0^{N^{\al -\frac12}} u^{\frac{6-6\al}{2\al -1}} \phi(\kappa u) du
\nonumber\\
&=& \frac{2}{2\al-1} \kappa^{-\frac{3}{2\al-1}} \int_0^{\kappa N^{\al -\frac12}} v^{\frac{6-6\al}{2\al -1}} \phi(v) dv
\nonumber\\
&\le & \frac{2}{2\al-1} \kappa^{-\frac{3}{2\al-1}} \int_0^\infty v^{\frac{6-6\al}{2\al -1}} \phi(v) dv
:=C_\al <\infty .
\end{eqnarray}
Since $\frac{6-6\al}{2\al -1} >0$, as $\al\in (\frac{1}{2}, 1)$, we can summarize our
results for this case as follows.
\begin{pro}
\label{pro:al>0.5}
{\rm
For $1/2< \al <1$ and $\kappa >0$, we have\\
\hspace*{0.15in} (a) the lower bound:
\begin{eqnarray}
\label{lb>0.5}
\ex[L_N]\ge
\sig\Big(\frac{y^*_\al}{\kappa}\Big)^{1/(2\alpha -1)} G(y^*_\al), \quad{\rm for}\quad N\ge \bar y_\al:=(2\al -1)^{-1/2},
\end{eqnarray}
\hspace*{0.38in} where $y^*_\al>0$ is the solution to the equation $\phi (y)=2\al y\bar\Phi (y)$
(and hence, independent of $N$); and\\
\hspace*{0.15in} (b) the upper bound: $\ex[L_N] \le \sig C_\al$, with $C_\al$
as specified in
(\ref{cal}).\\
Hence, in this case $\ex[L_N]$ is both upper- and lower-bounded by constants.
}
\end{pro}
\subsection{Case of $\mathbf{\alpha \in [0,\frac12]}$ and $\mathbf{\kappa > 0}$}
For $\alpha \in [0,\frac12]$, as $n\to\infty$, we have $G (\kappa {n}^{\alpha -1/2})\to G(0)$ or $G(\kappa)$
depending on whether $\al <\frac{1}{2}$ or $\al=\frac{1}{2}$.
Hence, from (\ref{elnlb}) and (\ref{esn1}), we obtain,
for sufficiently large $N$,
\begin{eqnarray}
\label{elnlb1}
\ex[L_N]\ge\sig G(\kappa)\sqrt{N}, \quad{\rm when}\; \al=\frac{1}{2}; \qquad
\ex[L_N]\ge\sig G(0)\sqrt{N}, \quad{\rm when} \; \al\in [0,\frac{1}{2}).
\end{eqnarray}
For an upper bound,
observe that the concave function $f(n)=n^{\alpha}$ is bounded from {\it below} by a linear function:
\begin{eqnarray}
\label{slowerrates}
n^\alpha \ge \frac{N^\alpha}{N} n:= \Theta n, \qquad {\rm for}\quad 0\le n\le N .
\end{eqnarray}
Defining $W_n:= -\kappa \sigma \Theta n + \sigma \sum_{k=1}^n Z_n$,
we know that $L_N \le \max_{n\le N} W_n$.
Meanwhile, $\max_{n\le N} W_n$ can be calculated through Spitzer's identity.
More specifically, we have
\begin{eqnarray*}
\ex[W_n^+] =\sig\sqrt{n} G (\kappa\Theta\sqrt{n}), \qquad 0\le n\le N,
\end{eqnarray*}
and
\begin{eqnarray*}
\sum_{n=1}^N \frac{1}{n} \ex[W_n^+]{\buildrel\dot\over =}
\int_0^N \frac{1}{t} \ex[W_t^+] dt
=\frac{2\s}{\kappa\Theta} \int_{0}^{\kappa\Theta \sqrt{N}} G(x)dx .
\end{eqnarray*}
Therefore, making use of \eqref{g_int}, we obtain
\begin{eqnarray}
\label{elnub0}
\ex[L_N] \le \sum_{n=1}^N \frac{1}{n} \ex[W_n^+]
{\buildrel\dot\over =}
\frac{\s}{\kappa\Theta} \Big(\Phi(\kappa\Theta \sqrt{N})-\frac{1}{2}\Big)
+{\s}\sqrt{N}G(\kappa\Theta \sqrt{N}).
\end{eqnarray}
As $N\to \infty$,
and according to the definition of $\Theta$ in (\ref{slowerrates}),
we have
$\Theta \sqrt{N} \to 0$ if $\alpha \in [0, \frac{1}{2})$ and
$\Theta \sqrt{N} \to 1$ if $\al=\frac{1}{2}$.
The second term on the right-hand side of (\ref{elnub0}) is of order $\sqrt{N}$,
whereas the first term can be written as
$\frac{\s\sqrt{N}}{\kappa\Theta\sqrt{N}} [\Phi(\kappa\Theta \sqrt{N})-\frac{1}{2}]$.
Since $\lim_{y\to 0} (\Phi(y)-\frac{1}{2})/y =\phi (0)$, we know this term is on
the order of $\sqrt{N}$ as well.
Hence,
\begin{eqnarray}
\label{elnub1}
\ex[L_N] \le
{\s}\sqrt{N}[ \phi (\kappa\Theta \sqrt{N})+G(\kappa\Theta \sqrt{N})] \le {\s}\sqrt{N}[ \phi (0)+G(0)]=2{\s}\phi (0)\sqrt{N} .
\end{eqnarray}
These results allow us to conclude as follows.
\begin{pro}
\label{pro:al<0.5}
{\rm
In the case of $\alpha \in [0,\frac12]$ and $\kappa > 0$, $\ex[L_N]$ grows on
the order of $\sqrt{N}$.
In particular, we have the lower- and upper-bounds for $\ex[L_N]$ in (\ref{elnlb1}) and (\ref{elnub1}).
}
\end{pro}
\subsection{Case of $\mathbf{\kappa \leq 0}$}
First consider the lower bounds.
For $\al\in (\frac{1}{2}, 1)$ and $\kappa <0$, we have
\begin{eqnarray*}
\ex[S_n^+] = \sigma \sqrt{n} G(\kappa n^{\al -1/2}){\buildrel\dot\over =} \sig |\kappa| n^\al,
\end{eqnarray*}
since $G(-x){\buildrel\dot\over =} x$ when $x>0$ is large.
Hence, for $N$ sufficiently large,
\begin{eqnarray}
\label{lbneg}
\ex[L_N] \ge \max_{0\le n \le N} \{ \ex[S_n^+]\} {\buildrel\dot\over =} \sigma |\kappa| N^\al,
\quad{\rm for} \; \al\in (\frac{1}{2}, 1) .
\end{eqnarray}
When $\al\in [0,\frac{1}{2} ]$ and $\kappa <0$, the lower bound for $\kappa>0$ in
(\ref{elnlb1}) remains valid; namely,
\begin{eqnarray}
\label{elnlbk<0}
\ex[L_N]\ge\sig G(\kappa)\sqrt{N}, \quad{\rm when}\; \al=\frac{1}{2}; \qquad
\ex[L_N]\ge\sig G(0)\sqrt{N}, \quad{\rm when} \; \al\in [0,\frac{1}{2}).
\end{eqnarray}
Next, we consider the upper bounds.
Define $Y_n := S_n +\kappa\sig n^\al$, for $n=0,1,\dots,N$.
Note that $Y_n$ is a random walk with zero drift and $Y_0=0$.
Then,
\begin{eqnarray}
\label{ubk<0}
L_N&=&\max_{0 \le n\le N}\{ S_n\}
= \max_{0 \le n\le N} \{Y_n - \kappa\sig n^\al \}
\nonumber\\
&\le& \max_{0 \le n\le N} \{Y_n\} + \max_{0 \le n\le N} \{- \kappa\sig n^\al \}
\nonumber\\
&=& \max_{0 \le n\le N} \{Y_n\} + (- \kappa\sig N^\al) ,
\end{eqnarray}
where the inequality follows from the subadditivity of the maximum operator,
and the last equality is due to $\kappa <0$.
Applying Spitzer's identity to the random walk $Y_n$,
similar to the derivation of $\ex[L_N]$ in (\ref{eln}) for the case of $\al=1$, we obtain
$$ \ex\Big[ \max_{0 \le n\le N} \{Y_n \}\Big] =\ex\Big[ \max_{1 \le n\le N} \{Y^+_n \} \Big]
= \sum_{n=1}^N \frac{\sig\sqrt{n}}{n} G(0){\buildrel\dot\over =} \sig G(0) \int_0^N \frac{dt}{\sqrt{t}} =2\sig G(0) \sqrt{N}.
$$
Hence, combining this with \eqref{ubk<0}, renders
\begin{align}
\label{ubk<0a}
\ex[L_N] \le 2\sig G(0) \sqrt{N} + |\kappa| \sig N^\al, \qquad \al \ge 0.
\end{align}
Our results for this case then can be summarized in the following proposition.
\begin{pro}
\label{pro:negk}
{\rm
For $\kappa<0$, $\ex[L_N]$ grows on the order of $\sig |\kappa| N^\al$ when
$\al\in(\frac{1}{2}, 1 )$;
it grows on the order of $\sig G(0) \sqrt{N} $ when $\al\in[0, \frac{1}{2})$;
and it grows on the order of $\sig G(k) \sqrt{N} $ when $\al= \frac{1}{2}$.
In particular, the lower bound of $\ex[L_N]$ follows (\ref{lbneg}) and
(\ref{elnlbk<0}), with its upper bound following (\ref{ubk<0a}).
}
\end{pro}
\medskip
Finally, when $\kappa =0$, we have from (\ref{ubk<0}) and (\ref{ubk<0a})
$$\ex[L_N]=\ex\Big[ \max_{0 \le n\le N} \{Y_n \}\Big] =2\sig G(0) \sqrt{N},$$
which is also consistent with setting $\kappa=0$ in the lower bound
(\ref{elnlbk<0}) and in the upper bound (\ref{ubk<0a}).
\subsection{Brownian approximation for case of $\mathbf{\alpha = \frac12}$}
\label{sec:hitting_time_approach}
The case of $\al= \frac12$ will turn out to produce the right trade-off between cost and service,
as we will see in the next section,
thus yielding the optimal order of production.
It is therefore appropriate to explore the possibility of a more accurate
estimation of $L_n$ in this particular case.
To this end, we study the continuous counterpart of $L_n$,
i.e., a Brownian motion with square-root drift.
More specifically, $\ex[L_N]$ can be approximated as
\begin{equation}
\ex[L_N] \;\; {\buildrel\dot\over =} \;\; \sqrt{\sigma}\ex\left[\sup_{1\le s\le t} \left\{B(s)-\frac{\kappa}{\sqrt{\sigma}}\sqrt{s}\right\}\right] ,
\label{eq:BrownianApproximation}
\end{equation}
where $B(t)$ is a standard Brownian motion.
Defining
\begin{align*}
\tau_x & := \inf \Big\{t > 1:B(t)=\frac{\kappa}{\sqrt{\sigma}}\sqrt{t}; B(1) = - x \Big\}
\end{align*}
for any $x\ge 0$, we then have
\begin{eqnarray}
\ex\left[\sup_{1\le s\le t} \left\{B(s)-\frac{\kappa}{\sqrt{\sigma}}\sqrt{s}\right\}\right]&=& \int_0^\infty
\mbox{\sf P}\left[\sup_{1\le s\le t} \left\{B(s)-\frac{\kappa}{\sqrt{\sigma}}\sqrt{s}\right\} \ge x \right] dx \nonumber \\
&=& \int_0^\infty \mbox{\sf P}\left[\tau_x < t \right] dx.
\label{eq:BA:HT}
\end{eqnarray}
Now, define a new process $Y(u) := B(e^{2u})/e^u$.
One can readily verify, as shown in~\cite{Breiman}, that $Y(u)$ is an
Ornstein-Uhlenbeck process with $Y(0)=x$.
Further define
\begin{equation*}
T_{a,b} \;\; := \;\; \inf \{u\ge 0: Y(u) \ge b, Y(0) = a\} .
\end{equation*}
We note that, for any $t \ge 0$,
\begin{equation}
\mbox{\sf P}[ \tau_x > e^{2t}] \;\; = \;\; \mbox{\sf P}[ T_{-x,\kappa/\sqrt{\sigma}} >t] ,
\label{eq:BA:FPT}
\end{equation}
a fact that has been known since~\cite{Breiman}.
Our task is thus reduced to the evaluation of $T_{a,b}$.
When $b=0$, an explicit expression for $T_{a,b}$ has been derived;
see, e.g.,~\cite{RicciardiSato,leblanc,GJYor}.
More generally when $b\neq 0$ (as is of interest here), however,
only the Laplace transform of the density of $T_{a,b}$ is known.
Using various methods, the moments of this distribution also have
been calculated;
refer to, e.g.,~\cite{Siegert51,Sato78,CRS81,NRS85,RicciardiSato}.
In theory, the distribution function of $T_{a,b}$ can be recovered
from all of its moments.
In practice, the distribution function of $T_{a,b}$ can be
effectively approximated by its finite moments up to a certain
order.
Moreover, as shown in \cite{BertsimasPopescu}, the error bound of
such an approximation can be computed through numerical methods,
such as semi-definite programming.
We therefore propose to use the first $K$ moments $M_{a,b}^1, \ldots, M_{a,b}^K$
of $T_{a,b}$ where $a=-x$ and $b=\kappa/\sqrt{\sigma}$, in combination
with \eqref{eq:BrownianApproximation}, \eqref{eq:BA:HT} and \eqref{eq:BA:FPT},
to obtain a more accurate estimation of $L_n$ in the case of $\al= \frac12$.
The methods and results in~\cite{Siegert51,NRS85,RicciardiSato} can be
used to calculate the first $K$ moments
$M_{a,b}^1, \ldots, M_{a,b}^K$.
When $K \leq 3$, we can additionally exploit the closed-form tight
bounds in~\cite{BertsimasPopescu} to further improve our approximation
of the distribution function of $T_{a,b}$.
(Finding the best possible bounds for $K \geq 4$ is NP-hard~\cite{BertsimasPopescu}.)
As a specific instance of our proposed approach, we consider in more
detail the case of utilizing the first three moments of $T_{a,b}$
(i.e., $K=3$).
Our starting point is the set of closed-form expressions for
$M_{a,b}^1, M_{a,b}^2, M_{a,b}^3$ provided in~\cite{NRS85} and
parameterized by $a=-x$ and $b=\kappa/\sqrt{\sigma}$.
Then, letting $X$ denote a generic random variable that follows the
probability distribution function of $T_{-x,\kappa/\sqrt{\sigma}}$,
we next obtain closed-form tight upper bounds on both $\mbox{\sf P}[X \le z]$
and $\mbox{\sf P}[X\ge z]$ in terms of the expressions for
$M_{-x,\kappa/\sqrt{\sigma}}^1, M_{-x,\kappa/\sqrt{\sigma}}^2, M_{-x,\kappa/\sqrt{\sigma}}^3$.
More specifically, we exploit the following propostion adapted from
Theorem 3.3 in~\cite{BertsimasPopescu}.
\begin{pro}\label{prop:BertsimasPopescu}
For a constant $\delta > 0$ and a nonegative real random variable
$X \: {\buildrel\rm d\over =} \: T_{-x,\kappa/\sqrt{\sigma}}$ with first three moments
$M_1 = M_{-x,\kappa/\sqrt{\sigma}}^1$, $M_2 = M_{-x,\kappa/\sqrt{\sigma}}^2$
and $M_3 = M_{-x,\kappa/\sqrt{\sigma}}^3$, the functionals
$\mbox{\sf P}[X > (1+\delta) M_1]$ and $\mbox{\sf P}[X < (1-\delta) M_1]$ can be approximated as
\begin{equation*}
\mbox{\sf P}[X > (1+\delta) M_1] {\buildrel\dot\over =} f_1(C_M^2,D_M^2, \delta) \qquad \mbox{and} \qquad \mbox{\sf P}[X < (1-\delta) M_1] {\buildrel\dot\over =} f_2(C_M^2,D_M^2, \delta) ,
\end{equation*}
respectively, where
\begin{align*}
f_1(C_M^2,D_M^2, \delta)& = \left\{
\begin{array}{cc} \min \left(\frac{C_M^2}{C_M^2+\delta^2}, \frac{1}{1+\delta} \cdot \frac{D_M^2}{D_M^2 + (C_M^2-\delta)^2}\right), & \qquad \delta > C_M^2 , \\
\frac{1}{1+\delta} \cdot \frac{D_M^2+(1+\delta)(C_M^2-\delta)}{D_M^2 + (1+C_M^2)(C_M^2-\delta)}, & \qquad \delta \leq C_M^2 , \end{array} \right. \\
f_2(C_M^2,D_M^2, \delta)& = 1- \frac{(C_M^2+\delta)^3}{(D_M^2 + (C_M^2+1)(C_M^2+\delta))(D_M^2 + (C_M^2+\delta)^2)} , \qquad \delta < 1 , \\
C_M^2 & = \frac{M_2-M_1^2}{M_1^2} , \\
D_M^2 & = \frac{M_1M_3-M_2^2}{M_1^4} .
\end{align*}
\end{pro}
Lastly, we combine Proposition~\ref{prop:BertsimasPopescu} and the expressions for
$M_{-x,\kappa/\sqrt{\sigma}}^1, M_{-x,\kappa/\sqrt{\sigma}}^2, M_{-x,\kappa/\sqrt{\sigma}}^3$
in \cite{NRS85} together with \eqref{eq:BrownianApproximation}, \eqref{eq:BA:HT},
\eqref{eq:BA:FPT} to obtain the more accurate approximation of $L_n$ in the case
of $\al= \frac12$, as desired.
\section{Cost Minimization and Asymptotically Optimal Solutions}
\label{sec:optimality}
In this section we consider instances of two cost minimization models for the lost-sale system of interest.
For the first model, the total cost consists of both the production cost and lost-sale penalty.
To be more precise,
we minimize $c\sum_{n=1}^N x_n + p \ex[L_N]$ where $c$ and $p$ are the per-unit production cost and lost-sales penalty, respectively.
From \eqref{supply},
ignoring the constant term $c(\mu_1+\cdots +\mu_N)$, we have
\begin{eqnarray}
\label{prodlost}
\min_{\alpha,\kappa} \quad c\sig\kappa N^\al + p \ex[L_N] .
\end{eqnarray}
For the second model, we incorporate the inventory (holding) cost in each period.
Specifically, we have
\begin{eqnarray}
\label{prodlostinv}
\min_{\alpha,\kappa} \quad c\sig\kappa N^\al + p \ex[L_N] + h \ex[H_N] ,
\end{eqnarray}
where $h$ is the per-unit inventory holding cost.
Recall from Proposition \ref{pro:lost} that the inventory (at the end) of
period $n$ is given by $H_n=L_n-S_n$, and therefore
$\ex[H_N]= \ex[L_N]+\kappa \sigma N^\alpha$.
Our first step is to find the optimal value of $\al$ in these optimization problems.
For this purpose,
let us start with a summary of the results we have derived so far concerning the asymptotics of
the expected loss sales in the limit as $N\to\infty$.
\begin{itemize}
\item[(i)]
When $\alpha=1$, $\ex[L_N]$ is bounded by a constant
$\frac{\sig}{2\kappa}$ if $\kappa >0$; whereas it grows linearly in $N$
if $\kappa <0$.
\item[(ii)]
When $\alpha\in (\frac{1}{2}, 1)$ and $\kappa >0$, $\ex[L_N]$ is bounded from above
and below by constants that are independent of $N$ (but dependent on $\al$
and $\kappa$); whereas it grows on the order of $N^\al$ if $\kappa <0$.
\item[(iii)]
When $\alpha\in [0,\frac{1}{2} ]$, $\ex[L_N]$ grows on the order of $\sqrt {N}$,
regardless of whether $\kappa >0$ or $\kappa \leq 0$.
\end{itemize}
From \eqref{prodlost},
if $\al \in (\frac{1}{2},1]$ and $\kappa >0$, then the second term in the objective will be bounded by a constant independent of $N$;
hence, for sufficiently large $N$, the objective is of order $N^\al$, i.e., the same order as the first term.
On the other hand,
if $\al \in (\frac{1}{2},1]$ and $\kappa <0$, then the first term will decrease while the second term will increase, both on the order of $|\kappa| N^\al$;
hence, if $c> p$, then $\kappa =-\infty$ minimizes the objective value, whereas if $c\le p$, then $\kappa \to 0$ minimizes the objective value.
Therefore, the only non-trivial solution of the optimization problem is to set
$\al\le \frac{1}{2}$, corresponding to the case summarized in (iii) above.
Here, the first term of the objective function is of order $N^\alpha$, which
is dominated by the second term of order $\sqrt {N}$ (except when $\alpha=\frac{1}{2}$).
Consequently, the objective value will also be of order $\sqrt{N}$, as is
further confirmed by replacing $\ex[L_N]$ with its lower bound via Jensen's
inequality (which applies in all cases):
$\ex[L_N] \ge \ex[\max_{1\le n\le N} S_n^+]\ge \ex[S_N^+]$.
This leads to the following minimization problem, which is a lower bound of
the original problem:
\begin{eqnarray}
\label{eqn:production_LB_1}
\min_z \quad cz + p \ex\left[\left( \sig\sum_{n=1}^NZ_n -z\right)^+\right] = cz+p\ex \big[(\sig \sqrt{N} Z- z\big)^+].
\end{eqnarray}
To allow for any $\al\in[0,1]$ and any $\kappa$, we write
$$z=\sig\kappa N^\al =\sig\sqrt{N}\kappa N^{\al-\frac{1}{2}}:= \sig\sqrt{N} y,$$
and then the optimal solution to minimizing $cy+p \ex[(Z-y)^+]$ is given by
\begin{eqnarray}
\label{ylb}
y^*=\bar\Phi^{-1}\Big(\frac{c}{p}\Big)=\Phi^{-1}\Big(\frac{p-c}{p}\Big).
\end{eqnarray}
With $y^*$ being a constant, independent of $N$, we must have that $\al=\frac{1}{2}$,
and hence $\kappa =y^*$.
The corresponding objective value then can be expressed as
\begin{eqnarray}
\label{yoptval}
\sigma \sqrt{N} [c y^* +pG (y^*)]=\sigma \sqrt{N} [c y^* +p\phi (y^*)-py^*\bar\Phi(y^*)]= \sigma \sqrt{N} p\phi(y^*) .
\end{eqnarray}
This confirms our previous statements; namely, the minimal overall cost cannot be
lower than order $\sqrt{N}$.
Moreover, we shall henceforth assume that $p\ge c$, unless noted otherwise,
because if $p<c$ then the lower-bound solution in (\ref{ylb}) already indicates
what will happen: make $y$ as small (negative) as possible; whereas the
objective value is lower bounded by the expression in (\ref{yoptval}).
We therefore only need to consider the case of $\al =\frac12$.
Furthermore, upon careful examination of the objective of the second optimization
model, we can reach the same conclusion.
In the remainder of this section, we start by considering asymptotically optimal solutions of
the above optimization models using two different approaches to obtain the desired solutions.
We then turn to incorporate inventory costs in our optimization models.
Finally, we consider the asymptotic equivalence of lost-sales and backorder models
under certain conditions.
\subsection{Asymptotically Optimal Solutions}
With our focus on $\alpha = \frac12$, we first obtain an asymptotically optimal solution through a Brownian approximation
of the objective function for large $N$.
We then obtain optimal solutions through an analysis of the asymptotic behavior of our upper and lower bounds of $\ex[L_N]$.
\subsubsection{Brownian Approximation}
\label{sec:Brownian_Approx}
Recall that the optimization problem for the production loss trade-off, when $\alpha$ is fixed to be $\frac12$, has the form:
\begin{eqnarray}
\label{prodlost_recall}
\min_{\kappa} \quad c\sig\kappa N^{1/2} + p \ex[L_N] .
\end{eqnarray}
Taking a closer look at the objective function, we conclude
\begin{align*}
c\sig\kappa N^{1/2} + p \ex[L_N] & = \left(c\sig\kappa + p \frac{\ex[L_N]}{ N^{1/2}}\right)N^{1/2} .
\end{align*}
Moreover,
\begin{align*}
\frac{\ex[L_N]}{ N^{1/2}} =\ex\left[\max_{n\le N}\frac{S_n}{N^{1/2}}\right]= \ex\left[\max_{n\le N}\frac{\sum_{i=1}^n\sig Z_i-\kappa \sig \sqrt{n} }{N^{1/2}}\right]
\end{align*}
where $Z_i$ are i.i.d.\ copies of a random variable following the standard normal distribution.
This leads to our next result.
\begin{thm}
Let $\rho(\kappa)= \sup_{t\le 1 } B_t -\kappa \sqrt{t}$, where $B_t $ is a standard Brown motion. Then,
\begin{align*}
\lim_{N\rightarrow \infty} \ex\left[\max_{n\le N}\frac{\sum_{i=1}^nZ_i-\kappa \sqrt{n} }{N^{1/2}}\right] \rightarrow \rho(\kappa) .
\end{align*}
\end{thm}
\proof{Proof.}
We first obtain
\begin{align*}
\max_{n\le N}\frac{\sum_{i=1}^nZ_i-\kappa \sqrt{n} }{N^{1/2}}& = \max_{t \in \{ \frac{n}{N}, n=1,2,\ldots, N\}} \frac{\sum_{i=1}^{Nt}Z_i-\kappa \sqrt{Nt} }{N^{1/2}}\\ & = \sup_{t \in (0,1]} \frac{\sum_{i=1}^{\lfloor Nt\rfloor}Z_i-\kappa \sqrt{\lfloor Nt \rfloor} }{N^{1/2}} .
\end{align*}
From the functional central limit theorem, for the summation of random variables (see e.g.,~\cite{chenyaobook}), we can conclude that
$$\frac{\sum_{i=1}^{\lfloor Nt\rfloor}Z_i-\kappa \sqrt{\lfloor Nt \rfloor} }{N^{1/2}} \Rightarrow B_t -\sqrt{t}.$$
The desired convergence then follows from the continuous mapping theorem.
\Halmos\endproof
The above theorem implies that, for any $\epsilon>0$ and when $N$ is large enough, we have
\begin{align*}
\Big| \frac{\ex[L_N]}{ N^{1/2}}-\rho(\kappa)\Big| < \epsilon
\end{align*}
uniformly on compact sets in $\kappa$.
Although there is no explicit formula for $\rho(\kappa)$, computational methods that make use of the related results in \cite{APP,Breiman,RicciardiSato,Sato78}
together with Monte-Carlo simulation can be employed;
in particular, these methods can be used to obtain accurate estimation of the value of $\rho(\kappa)$ and its derivative in order to solve the Brownian version of our optimization problem:
\begin{equation}
\min_{\kappa} \quad c\sig\kappa N^{1/2} + p \sig \rho(\kappa).
\label{BrownOpt}
\end{equation}
Let $\kappa^*$ denote an optimum of the problem \eqref{BrownOpt}; of course, $\kappa^*$ corresponds to a specific production plan.
We next show that this production plan is asymptotically optimal, defined as follows.
\begin{defn}
A production plan is {\it asymptotically optimal} if
\begin{align*}
\lim \sup_{N\rightarrow \infty} \frac{J_N(\kappa)}{J_N^*} \le 1.
\end{align*}
\end{defn}
\begin{thm}
The optimal solution of \eqref{BrownOpt}, $\kappa^*$, is asymptotically optimal for the problem \eqref{prodlost_recall}.
\end{thm}
\proof{Proof.}
Let $\kappa^*(N)$ denote an optimum of problem \eqref{prodlost_recall} for a given $N$.
Then, for any $N$, we have
\begin{align*}
\frac{J_N(\kappa)}{J_N^*} &= \frac{c\sig\kappa^* N^{1/2} + p \ex[L_N(\kappa^*)] } {c\sig\kappa^*(N) N^{1/2} + p \ex[L_N(\kappa^*(N))] }\\ & =
\frac{c\sig\kappa^* N^{1/2} + p \ex[L_N(\kappa^*)] - \left(c\sig\kappa^*(N) N^{1/2} + p \ex[L_N(\kappa^*(N))]\right)+ \left(c\sig\kappa^*(N) N^{1/2} + p \ex[L_N(\kappa^*(N))]\right)} {c\sig\kappa^*(N) N^{1/2} + p \ex[L_N(\kappa^*(N))] }
\\ & \le \frac{c\sig\kappa^* N^{1/2} + p \ex[L_N(\kappa^*)] - \left(c\sig\kappa^*(N) N^{1/2} + p \ex[L_N(\kappa^*(N))]\right)+ } {c\sig\kappa^*(N) N^{1/2} + p \ex[L_N(\kappa^*(N))] } + 1.
\end{align*}
We further know that the first term can be bounded by a constant, and thus goes to $0$ in the limit at $N \rightarrow \infty$.
The desired result of asymptotic optimality follows.
\Halmos\endproof
It is evident that the asymptotically optimal result is not restricted to the production optimization problem, as a similar conclusion can be also reached for the inventory optimization problem.
\subsubsection{Lower- and Upper-Bound Surrogate Problems}
\label{sec:prodlost}
We now turn to consider an alternative approach for obtaining optimal solutions through our upper and lower bounds of $\ex[L_N]$.
For finite $N>0$, we have already analyzed the lower bound surrogate problem in (\ref{yoptval}).
We next seek to minimize the upper bound with respect to the choice of $\kappa$.
From the upper bound for the case $\al=\frac{1}{2}$ and $\kappa >0$ in terms of
\eqref{slowerrates} and \eqref{elnub0},
upon replacing $c\sigma\sqrt{N}$ and $p\sigma\sqrt{N}$ with $c$ and $p$,
respectively, we want to solve the following optimization problem:
\begin{eqnarray}
\label{ubobj}
\min_{\kappa \ge 0} \; c\kappa + p \left( \frac{1}{\kappa} \Big[\frac{1}{2} - \bar\Phi(\kappa)\Big]+G(\kappa)\right).
\end{eqnarray}
The optimality condition is then given by
$$c+\frac{p}{\kappa^2} \Big(\kappa \phi(\kappa) +\bar\Phi(\kappa) -\frac{1}{2} \Big) = p \bar\Phi(\kappa), $$
which simplifies to
\begin{eqnarray}
\label{uboptsol}
c\kappa + p G(\kappa) = \frac{p}{\kappa} \Big(\frac{1}{2} - \bar\Phi(\kappa)\Big) .
\end{eqnarray}
Denoting the solution of \eqref{uboptsol} by $\kappa_u$ and observing
$\kappa_u\ge 0$ implies that $\bar\Phi(\kappa_u)\le \frac{1}{2}$.
Further denote the objective values of the lower- and upper-bound problems
in (\ref{yoptval}) and (\ref{ubobj}) by $V^\ell$ and $V^u$, respectively.
We write $\kappa_\ell:=y^*$ following (\ref{ylb}), which yields
\begin{eqnarray}
\label{kl}
\kappa_\ell :=\bar\Phi^{-1}\Big(\frac{c}{p}\Big)=\Phi^{-1}\Big(\frac{p-c}{p}\Big).
\end{eqnarray}
Then, from (\ref{yoptval}), (\ref{ubobj}) and \eqref{uboptsol}, we have
\begin{eqnarray}
\label{vuvell}
\frac{V^u}{V^\ell} = \frac{2[ c\kappa_u +pG(\kappa_u)] }{c\kappa_\ell +pG(\kappa_\ell)} \ge 2,
\end{eqnarray}
since $\kappa_\ell$ is the minimizer of $cy+pG(y)$.
To go in the opposite direction,
from (\ref{yoptval}), (\ref{ubobj}) and (\ref{uboptsol}), we can also write
the above ratio as
$$\frac{V^u}{V^\ell} = \frac{2 [\frac{1}{2} - \bar\Phi(\kappa_u)]}{\kappa_u \phi(\kappa_\ell)}.$$
Observe that $\frac{1}{x} [\frac{1}{2} - \bar\Phi(x)]$ is decreasing in $x\ge 0$
(and at $x=0$, the limit is $\phi (0)$, via l'H\^opital's rule), since its
derivative is $\frac{1}{x^2} [x\phi(x) + \bar\Phi(x)-\frac{1}{2}]$.
It is then straightforward to verify that $x\phi(x) + \bar\Phi(x)-\frac{1}{2}\le 0$
for $x\ge 0$.
We therefore have
$$\phi (0)\ge \frac{1}{\kappa_u} \Big(\frac{1}{2} - \bar\Phi(\kappa_u)\Big) ,$$
and hence
\begin{eqnarray}
\label{vuvell1}
\frac{V^u}{V^\ell} \le \frac{2\phi(0)}{\phi(\kappa_\ell)} .
\end{eqnarray}
In fact, we also know that $\kappa_\ell\le\kappa_u$, which follows from
$$\frac{p}{\kappa_\ell} \Big(\frac{1}{2} - \bar\Phi(\kappa_\ell)\Big)\ge \frac{p}{\kappa_u} \Big(\frac{1}{2} - \bar\Phi(\kappa_u)\Big)$$
(since $\kappa_u$ is the minimizer of the right-hand side), together with
the fact justified above that $\frac{1}{x} [\frac{1}{2} - \bar\Phi(x)]$ is decreasing in $x$.
When $p=2c$, we have $\kappa_\ell =0$; refer to (\ref{kl}).
Then, we will have $\kappa_u=0$ as well; refer to (\ref{uboptsol}).
These results render, for this case, $V^u=2V^\ell$.
In summary, we have established the following proposition.
\begin{pro}
\label{pro:lostbounds}
{\rm
Let $V^*$ denote the optimal value of the original problem in (\ref{prodlost}).
Let $V^\ell$ and $V^u$ denote the objective values of the lower- and
upper-bound problems in (\ref{yoptval}) and (\ref{ubobj}), respectively.
Then,
$$ \frac{V^u}{V^\ell}= \frac{2[ c\kappa_u +pG(\kappa_u)] }{c\kappa_\ell +pG(\kappa_\ell)}
=\frac{2 [\frac{1}{2} - \bar\Phi(\kappa_u)]}{\kappa_u \phi(\kappa_\ell)} $$
where $\kappa_\ell=y^*$ is the lower-bound solution following (\ref{ylb}),
$\kappa_u$ is the upper-bound solution to the optimality equation in
(\ref{uboptsol}), and $\kappa_\ell\le\kappa_u$.
Consequently, we have
$$2\le \frac{V^u}{V^\ell} \le \frac{2\phi(0)}{\phi(\kappa_\ell)}$$
and
$$\frac{V^\ell}{V^u}\le \frac{V^\ell}{V^*}\le \frac{V^u}{V^*} \le \frac{V^u}{V^\ell}.$$
}
\end{pro}
\medskip
Based on the foregoing analysis, we propose to use $\al=\frac{1}{2}$ and $\kappa=\kappa_u$,
the solution to the upper-bound problem $V^u$ (for $\kappa >0$), for our
approximate solution to the original optimization problem in (\ref{prodlost}).
When $\kappa <0$, we can still use the solution to $V^u$ as a heuristic, simply
because the upper- or lower-bounds in this case will not lead to meaningful solutions.
(For instance, the upper-bound problem is $\min_{\kappa\le 0} (c-p)\kappa$,
and thus the solution is: $\kappa^*=0$ if $c\le p$, and $\kappa^* =-\infty$
if $c>p$.)
All we need is to remove the constraint $\kappa \ge 0$ from the minimization
problem in (\ref{ubobj}), and then the optimality equation in (\ref{uboptsol})
still applies (with the right-hand side {\it increasing} in $\kappa$ for
$\kappa\le 0$).
However, it is no longer the case that $V^u$ provides an upper bound.
In fact, when $\kappa<0$, $V^u$ is a {\it lower} bound of $V^*$: the
constructed linear lower bound to the concave function $\sqrt n$, along with a
negative $\kappa$, implies subtracting {\it less} from the production costs,
which in turn must result in lower lost-sales penalties.
Thus, we have $V^u/V^* \le 1$, and the performance of $V^u$, as an
approximation to $V^*$, should be comparable to, if not better than, its
performance in the case of $\kappa >0$.
\subsection{Incorporating Inventory Cost}
\label{sec:inventory_cost}
Now, we turn to incorporate inventory costs in our analysis.
Referring back to \eqref{prodlostinv},
let $h/N$ be the holding cost per unit of inventory that is held for one period,
with $h>0$ a constant parameter.
The reason for including $N$ in the denominator can be explained as follows.
The application context we have in mind is a fast-moving production-inventory
system, where the planning horizon is of a fixed length, say one unit, and $h$
is the cost to hold one unit of inventory over this entire horizon.
By dividing the horizon into $N$ segments (periods), we are effectively using
$N$ as a scaling parameter on the demand.
Hence, as previously noted, instead of demand $D$ over the entire planning horizon,
we have been considering $D_n$ for each period $n=1,\dots, N$, with $D_n{\buildrel\rm d\over =} D$ and
independent of $n$.
Another way to motivate this approach is due to the fact that, if $h$ is not
scaled by $N$, then the inventory holding cost, because of its cumulative
nature over time, will dominate all other costs when $N$ grows.
Therefore, using $\al=\frac{1}{2}$, the inventory cost is given by
$$\frac{h}{N}\sum_{n=1}^N \sig\sqrt{n} [G(\kappa) +\kappa]{\buildrel\dot\over =}
\frac{2}{3} h \sig\sqrt{N} [G(\kappa) +\kappa]$$
for the lower bound;
and similarly, for the upper bound, we have
$$\frac{h}{N}\sum_{n=1}^N \sig\sqrt{n} \left(\frac{1}{\kappa} \Big[\frac{1}{2}-\bar\Phi(\kappa)\Big]
+G(\kappa) +\kappa\right){\buildrel\dot\over =}
\frac{2}{3} h \sig\sqrt{N} \left(\frac{1}{\kappa} \Big[\frac{1}{2}-\bar\Phi(\kappa)\Big]
+G(\kappa) +\kappa\right);$$
refer to \eqref{elnlb1} and \eqref{elnub0}.
Upon replacing $\frac{2}{3} h$ by $h$, the inventory costs derived above are
easily incorporated into both the Brownian approximation and the lower- and upper-bound
surrogate problems of the previous subsections:
simply replace $c$ by $c+h$, and $p$ by $p+h$.
\subsection{Asymptotic Equivalence of Lost-Sales and Backorder Models}
\label{sec:backorder}
Next, we consider another type of asymptotic results.
Suppose the penalty $p$ is large, say $p\to\infty$, whereas $N$ is fixed.
We first ask the question of whether it is possible to keep the expected
penalty cost $p\ex[L_N]$ finite (for a given $N$).
To this end,
let us consider the lower-bound solution with $\al=\frac{1}{2}$ and
$\kappa=\kappa_\ell=y^*$;
for simplicity, we shall write $y$ instead of $y^*$.
Recall from \eqref{ylb} that $y =\bar\Phi^{-1}(\frac{c+h}{p+h})$ when inventory
cost is included into the model, and thus $p\to\infty$ is equivalent to
$y\to\infty$.
Then, for any $n$, we have
$$p\ex[S_n^+] =\sig\sqrt{n} pG(y)=\sig\sqrt{n} \frac{cG(y)}{\bar\Phi(y)},$$
for which applying l'H\^opital's rule yields
$$\lim_{y\to\infty}\frac{G(y)}{\bar\Phi (y)}=\lim_{y\to\infty}\frac{\bar\Phi (y)}{\phi(y)}=\lim_{y\to\infty}\frac{1}{y}\to 0.$$
Namely, $p\ex[S_n^+]\to 0$ for every $n$.
In fact, with $S^+_n(y):=\sig\sqrt{n}(Z-y)^+$, this strategy also leads to
$\ex[S_n^+(y)]=\sig\sqrt{n} G(y)\to 0$ as $y\to\infty$ for every $n$.
Because $S^+_n(y)$ is decreasing in $y$, monotone convergence implies that
$S^+_n(y) \to 0$ as $y\to\infty$, for every $n=1,\dots, N$.
Hence, we must have $L_N(y)=\max_{1\le n\le N} S_n^+ (y)\to 0$ as $y\to\infty$,
and thus $\ex[L_N(y)]\to 0$ as $y\to\infty$.
Since the optimal strategy (to the original problem with inventory cost
included) can do no worse, we must have
\begin{eqnarray*}
&& \min_\kappa \Big(\sig\sqrt{N} (c+h)\kappa + (p+h)\ex[L_N (\kappa) ]\Big)
\nonumber\\
&\le& \sig\sqrt{N} (c+h)y + (p+h)\ex[L_N (y) ]
= \sig\sqrt{N} (c+h)y+o(y),
\end{eqnarray*}
where $y=\bar\Phi^{-1}(\frac{c+h}{p+h})$ as specified above.
On the other hand, based on the results established earlier, we know that the
right-hand side above is also a lower bound of the objective function, because
(with $\kappa^*$ denoting the minimizer of the first expression)
\begin{eqnarray*}
&& \min_\kappa \Big(\sig\sqrt{N} (c+h)\kappa + (p+h)\ex[L_N (\kappa) ]\Big) \\
& \ge& \sig\sqrt{N} (c+h)\kappa^* + (p+h)\ex[S^+_N (\kappa^*) ]\\
& \ge& \sig\sqrt{N} (c+h)y + (p+h)\ex[S^+_N (y) ]
= \sig\sqrt{N} (c+h)y+o(y),
\end{eqnarray*}
where the second inequality follows from $y$ being the minimizer of
the lower-bound problem.
We therefore have
\begin{eqnarray}
\label{lsub}
\min_\kappa \Big(\sig\sqrt{N} (c+h)\kappa + (p+h)\ex[L_N (\kappa) ]\Big)
= \sig\sqrt{N} (c+h)y+o(y).
\end{eqnarray}
Next, let us consider the backorder model.
Recall that $B_n=S_n^+$ is the number of backlogged units of demand and
that $H'_n=S_n^-$ is the inventory, both at the end of period $n$;
refer to Proposition \ref{pro:lost}.
Similar to the lost-sales model in the previous subsection, let $\frac{b}{N}$
and $\frac{h'}{N}$ be the per-unit backlog penalty and inventory holding costs
in each period, respectively.
Then, the (original) optimization problem can be expressed as
\begin{eqnarray}
\label{bk}
\min_\kappa \left( c\kappa + \frac{b}{N}\sum_{n=1}^N \ex[S^+_n (\kappa) ]
+\frac{h'}{N}\sum_{n=1}^N \ex[S_n^- (\kappa) ] \right)
= \sig\sqrt{N} \, \min_\kappa \left[ (c+h')\kappa + (b+h') G (\kappa) \right] ,
\end{eqnarray}
where $\frac{2}{3} b$ and $\frac{2}{3} h'$ are replaced by $b$ and $h'$
(analogous to the approach taken with $h$ in the previous subsection),
respectively, and taking into account
$S_n^-=S_n^+-S_n=S_n^+ +\sig\sqrt{n}\kappa$.
Let $y'$ denote the optimal solution to \eqref{bk}, clearly rendering
$y'=\bar\Phi^{-1}(\frac{c+h'}{b+h'})$.
Suppose we set $b=p$ and $h'=h$, and let $p\to\infty$.
Then, $y'=y$ and the left-hand side of (\ref{bk}) has the following
asymptotics for large $y$:
\begin{eqnarray}
\label{bklimit}
\sig\sqrt{N} \min_\kappa \Big( c\kappa + b \ex[S^+_n (\kappa) ] +h' [S_n^- (\kappa) ] \Big)
= (c+h)y' + o(y') =(c+h)y + o(y).
\end{eqnarray}
Upon comparing (\ref{lsub}) and (\ref{bklimit}), we have established the following result.
\begin{pro}
\label{pro:equiv}
{\rm
As $p \rightarrow \infty$, the objective values of the lost-sales and
backorder models become equivalent in that their ratio approaches $1$,
provided the lost-sales and backlog penalty costs are both equal to $p$,
and the purchasing and holding costs are also respectively equal in the
two models.
}
\end{pro}
\section{Conclusions}
\label{sec:conclusions}
A general class of high-volume, fast-moving production-inventory problems are studied.
Although we focus on the case in which the production is set to be the average demand
supplemented by a certain amount of safety stock, our assumptions on this safety stock
is quite general.
In the asymptotic regime, we identify the right order of the safety stock by studying
random walks with power drifts.
We also derive bounds and approximations for functionals of key performance measures
so that the production planning can be optimized under different settings.
Finally, our analysis provides another means and settings for establishing the asymptotic
equivalence of lost sale and backorder models, as observed by several authors using
different means and settings.
|
\section{Introduction}
\label{sec::introduction}
In this paper, we study longitudinal classification problems in which
the number of predictors can exceed the number of observations. The
setup: we observe $n$ individuals across discrete timepoints
$t=1,\ldots T$. At each timepoint we record $p$ predictor variables
per individual, and an outcome that places each individual into one of
$K$ classes. The goal is to construct a model that predicts the
outcome of an individual at time $t+\Delta$, given his or her
predictor measurements at time $t$.
Since we allow for the possibility that $p>n$, regularization must be
employed in order for such a predictive model (e.g., based on maximum
likelihood) to be well-defined. Borrowing from the extensive
literature on high-dimensional regression, we consider two well-known
regularizers, each of which also has a natural place in
high-dimensional longitudinal analysis for many scientific problems of
interest. The first is the {\it lasso} regularizer, which encourages
overall sparsity in the active (contributing) predictors at each
timepoint; the second is the {\it fused lasso} regularizer, which
encourages a notion of persistence or contiguity in the sets of active
predictors across timepoints.
Our work is particularly motivated by the analysis of a large data set
provided by the Cardiovascular Health Study Cognition Study (CHS-CS).
Over the past 24 years, the CHS-CS recorded multiple metabolic,
cardiovascular and neuroimaging risk factors for Alzheimer's disease
(AD), as well as detailed cognitive assessments for people of ages 65
to 110 years old \citep{lopez2003prevalence, saxton2004preclinical,
lopez2007incidence}.
As a matter of background, the prevalence of AD increases at an
exponential-like rate beyond the age of 65.
After 90 years of age, the incidence of AD increases dramatically,
from 12.7\% per year in the 90-94 age group, to 21.2\% per year in the
95-99 age group, and to 40.7\% per year for those older than 100 years
\citep{evans1989prevalence, fitzpatrick2004incidence,
corrada2010dementia}. Later, we examine data from 924 individuals
in the Pittsburgh section of the CHS-CS.
The objective is to use the data available from subjects at $t$ years
of age to predict the onset of AD at $t+10$ years of age
($\Delta=10$). For each age, the outcome variable assigns an
individual to one of 3 categories: normal, dementia, death. Refer to
Section \ref{sec::CHS} for our analysis of the CHS-CS data set.
\subsection{The multinomial fused lasso model}
Given the number of parameters involved in our general longitudinal
setup, it will be helpful to be clear about notation: see Table
\ref{tab:notation}. Note that the matrix $Y$ stores future outcome
values, i.e., the element $Y_{it}$ records the outcome of the
$i$th individual at time $t+\Delta$, where $\Delta \geq 0$ determines
the time lag of the prediction. In the following, we will generally
use the ``$\cdot$'' symbol to denote partial indexing; examples are
$X_{i\cdot t}$, the vector of $p$ predictors for individual $i$ at
time $t$, and $\beta_{\cdot tk}$, the vector of $p$ multinomial
coefficients at time $t$ and for class $k$. Also, Section
\ref{sec::GGD} will introduce an extension of the basic setup in which
the number of individuals can vary across timepoints, with $n_t$
denoting the number of individuals at each timepoint $t=1,\ldots T$.
\begin{table}[h!]
\centering
\begin{tabular}{| c | c |}
\hline
\textbf{Parameter} & \textbf{Meaning} \\
\hline \hline
$i=1, \ldots n$ & index for individuals \\ \hline
$j=1, \ldots p$ & index for predictors \\ \hline
$t=1,\ldots T$ & index for timepoints \\ \hline
$k=1, \ldots K$ & index for outcomes \\ \hline \hline
$Y$ & $n \times T$ matrix of (future) outcomes \\ \hline
$X$ & $n \times p \times T$ array of predictors\\ \hline
$\beta_0$ & $T \times (K-1)$ matrix of intercepts \\ \hline
$\beta$ & $p \times T \times (K-1)$ array of coefficients \\ \hline
\end{tabular}
\caption{Notation used throughout the paper.}
\label{tab:notation}
\end{table}
At each timepoint $t=1,\ldots T$, we use a separate multinomial logit
model for the outcome at time $t+\Delta$:
\begin{equation}
\begin{aligned}
\label{eq::multinomial}
\log \frac{\mathbb{P}(Y_{it}=1 | X_{i\cdot t} = x)}{\mathbb{P}(Y_{it}=K | X_{i\cdot t}
= x)}&= \beta_{0t1} + \beta_{\cdot t1}^T x \\
\log \frac{\mathbb{P}(Y_{it}=2 | X_{i\cdot t} = x)}{\mathbb{P}(Y_{it}=K | X_{i\cdot t}
= x)}&= \beta_{0t2} + \beta_{\cdot t2}^T x \\
& \vdots \\
\log \frac{\mathbb{P}(Y_{it}=K-1 | X_{i\cdot t} = x)}{\mathbb{P}(Y_{it}=K | X_{i\cdot
t} = x)}&= \beta_{0t(K-1)} + \beta_{\cdot t(K-1)}^T x.
\end{aligned}
\end{equation}
The coefficients are determined by maximizing a penalized log
likelihood criterion,
\begin{equation}
\label{eq::fusedmodel}
(\hat\beta_0, \hat\beta) \in \mathop{\mathrm{arg\,max}}_{\beta_0,\beta} \;
\ell(\beta_0,\beta) - \lambda_1 P_1(\beta) - \lambda_2 P_2(\beta),
\end{equation}
where $\ell(\beta_0,\beta)$ is the multinomial log likelihood,
\begin{equation*}
\ell(\beta_0,\beta) = \sum_{t=1}^T \sum_{i=1}^n \mathbb{P}(Y_{it}|X_{i\cdot
t}),
\end{equation*}
$P_1$ is the lasso penalty \citep{lasso},
\begin{equation*}
P_1(\beta) =\sum_{j=1}^p \sum_{t=1}^T \sum_{k=1}^{K-1}
|\beta_{jtk}|.
\end{equation*}
and $P_2$ is a version of the fused lasso penalty \citep{fuse} applied
across timepoints,
\begin{equation*}
P_2(\beta) = \sum_{j=1}^p \sum_{t=1}^{T-1} \sum_{k=1}^{K-1}
|\beta_{jtk}-\beta_{j(t+1)k}|.
\end{equation*}
(The element notation in \eqref{eq::fusedmodel} emphasizes the fact
that the maximizing coefficients \smash{$(\hat\beta_0,\hat\beta)$} need not
be unique, since the log likelihood $\ell(\beta_0,\beta)$ need not be
strictly concave---e.g., this is the case when $p>n$.)
In broad terms, the lasso and fused lasso penalties encourage sparsity
and persistence, respectively, in the estimated coefficients
\smash{$\hat\beta$}. A larger value of the tuning parameter $\lambda_1
\geq 0$ generally corresponds to fewer nonzero entries in
\smash{$\hat\beta$}; a larger value of the tuning parameter $\lambda_2
\geq 0$ generally corresponds to fewer change points in the piecewise
constant coefficient trajectories \smash{$\hat\beta_{j \cdot k}$}, across
$t=1,\ldots T$. We note that the form the log likelihood
$\ell(\beta_0,\beta)$ specified above assumes independence between the
outcomes across timepoints, which is a rather naive assumption given
the longitudinal nature of our problem setup. However, this naivety is
partly compensated by the role of the fused lasso penalty, which ties
together the multinomial models across timepoints.
It helps to see an example. We consider a simple longitudinal problem
with $n=50$ individuals, $T=15$ timepoints, and $K=2$ classes. At
each timepoint we sampled $p=30$ predictors independently from a
standard normal distribution. The true (unobserved) coefficient matrix
$\beta$ is now $30\times 15$; we set $\beta_{j\cdot}=0$ for $j=1\ldots
27$, and set the 3 remaining coefficients trajectories to be piecewise
constant across $t=1,\ldots 15$, as shown in the left panel of Figure
\ref{fig::example}. In other words, the assumption here is that only 3 of
the 30 variables are relevant for predicting the outcome, and these
variables have piecewise constant effects over time. We generated a
matrix of binary outcomes $Y$ according to the multinomial model
\eqref{eq::multinomial}, and computed the multinomial fused lasso
estimates \smash{$\hat\beta_0,\hat\beta$} in \eqref{eq::fusedmodel}. The
right panel of Figure \ref{fig::example} displays these estimates (all
but the intercept \smash{$\hat\beta_0$}) across $t=1,\ldots 15$, for a
favorable choice of tuning parameters $\lambda_1=2.5$,
$\lambda_2=12.5$; the middle plot shows the unregularized (maximum
likelihood) estimates corresponding to $\lambda_1=\lambda_2=0$.
\begin{figure}[!ht]
\centering
\includegraphics[width=\textwidth]{figs/example.pdf}
\caption{A simple example with $n=50$, $T=15$, $K=2$, and $p=30$. The
left panel displays the true coefficent trajectories across
timepoints $t=1,\ldots 15$ (only 3 of the 30 are nonzero); the
middle panel shows the (unregularized) maximum likelihood estimates;
the right panel shows the regularized estimates from
\eqref{eq::multinomial}, with $\lambda_1=2.5$ and $\lambda_2=12.5$.}
\label{fig::example}
\end{figure}
Each plot in Figure \ref{fig::example} has a $y$-axis that has been
scaled to suit its own dynamic range.
We can see that the multinomial fused lasso estimates, with an
appropriate amount of regularization, pick up the underlying trend in
the true coefficients, though the overall magnitude of coefficients is
shrunken toward zero (an expected consequence of the $\ell_1$
penalties). In comparison, the unregularized multinomial estimates
are wild and do not convey the proper structure. From the perpsective
of prediction error, the multinomial fused lasso estimates offer a
clear advantage, as well: over 30 repetitions from the same simulation
setup, we used both the regularized coefficient estimates (with
$\lambda_1=2.5$ and $\lambda_2=12.5$) and the unregularized estimates
to predict the outcomes on an i.i.d.\ test set. The average
prediction error using the regularized estimates was $0.114$ (with a
standard error of $0.014$), while the average prediction error from
the unregularized estimates was $0.243$ (with a standard error of
$0.022$).
\subsection{Related work and alternative approaches}
\label{sec::related}
The fused lasso was first introduced in the statistics literature by
\citet{fuse}, and similar ideas based on total variation, starting
with \citet{tv}, have been proposed and studied extensively in the
signal processing community. There have been many interesting
statistical applications of the fused lasso, in problems involving
the analysis of comparative genomic hybridization data
\citep{fusehot}, the modeling of genome association networks
\citep{graphguided}, and the prediction of colorectal cancer
\citep{lin2013}. The fused lasso has in fact been applied to the
study of Alzheimer's disease in \citet{xin2014}, though these authors
consider a very different prediction problem than ours, based on
static magnetic resonance images, and do not have the time-varying
setup that we do.
Our primary motivation, which is the focus of Section
\ref{sec::CHS}, is the problem of predicting the status of an
individual at age $t+10$ years from a number of variables measured
at age $t$.
For this we use the regularized multinomial model described in
\eqref{eq::multinomial}, \eqref{eq::fusedmodel}. We encode
$K=3$ multinomial categories as normal, dementia, and
death: these are the three possible outcomes for any individual at
age $t+10$. We are mainly interested in the prediction of
dementia; this task is complicated by the fact that risk factors
for dementia are also known to be risk factors for death
\citep{rosvall2009apoe}, and so to account for this, we include
the death category in the multinomial classification model.
An alternate approach would be
to use a Cox proportional hazards model \citep{cox1972}, where the
event of interest is the onset of dementia, and censorship corresponds
to death.
Traditionally, the Cox model is not fit with time-varying
predictors or time-varying coefficients, but it can be naturally
extended to the setting considered in this work, even using the
same regularization schemes. Instead of the multinomial model
\eqref{eq::multinomial}, we would model the hazard function as
\begin{equation}
\label{eq::hazard}
h(t+\Delta|X_{i \cdot t} = x) = h_0(t+\Delta) \cdot
\exp(x^T \beta_{\cdot t}),
\end{equation}
where $\beta \in \mathbb{R}^{p \times T}$ are a set of coefficients over time,
and $h_0$ is some baseline hazard function (that does not depend
on predictor measurements). Note that the hazard model
\eqref{eq::hazard} relates the instantaneous rate of failure (onset of
dementia) at time $t+\Delta$ to the predictor measurements at time
$t$. This is as in the multinomial model \eqref{eq::multinomial},
which relates the outcomes at time $t+\Delta$ (dementia or death) to
predictor measurements at time $t$. The coefficients in
\eqref{eq::hazard} would be determined by maximizing the
partial log likelihood with the analogous lasso and fused lasso
penalties on $\beta$, as in the above multinomial setting
\eqref{eq::fusedmodel}.
The partial likelihood approach can be viewed as a sequence of
conditional log odds models \citep{efron1977,kalb2002}, and therefore
one might expect the (penalized) Cox regression model described here
to perform similarly to the (penalized) multinomial
regression model pursued in this paper. In fact, the computational
routine described in Section \ref{sec::GGD} would apply to the Cox
model with only very minor modifications (that concern the gradient
computations). A rigorous comparison of the two approaches is beyond
the scope of the current manuscript, but is an interesting topic for
future development.
\subsection{Outline}
The rest of this paper is organized as follows. In Section
\ref{sec::GGD}, we describe a proximal gradient descent algorithm for
efficiently computing a solution \smash{$(\hat\beta_0,\hat\beta)$} in
\eqref{eq::multinomial}.
Next, we present an analysis of the CHS-CS data set in Section
\ref{sec::CHS}. Section \ref{sec::stability} discusses the stability
of estimated coefficients, and related concepts. In Section
\ref{sec::selection} we discuss numerous approaches for the selecting
the tuning parameters $\lambda_1,\lambda_2 \geq 0$ that govern the
strength of the lasso and fused lasso penalties in
\eqref{eq::multinomial}. In Section \ref{sec::discussion}, we
conclude with some final comments and lay out ideas for future work.
\section{A proximal gradient descent approach}
\label{sec::GGD}
In this section, we describe an efficient proximal gradient descent
algorithm for computing solutions of the fused lasso regularized
multinomial regression problem \eqref{eq::fusedmodel}. While a number
of other algorithmic approaches are possible, such as implementations
of the alternating direction method of multipliers \citep{admm}, we
settle on the proximal gradient method because of its simplicity, and
because of the extremely efficient, direct proximal mapping
associated with the fused lasso regularizer.
We begin by reviewing proximal gradient descent in generality, then we
describe its implementation for our problem, and a number of practical
considerations like the choice of step size, and stopping criterion.
\subsection{Proximal gradient descent}
\label{sec::proxintro}
Suppose that $g : \mathbb{R}^d \rightarrow \mathbb{R}$ is convex and differentiable,
$h : \mathbb{R}^d \rightarrow \mathbb{R}$ is convex, and we are interested in
computing a solution
\begin{equation*}
x^\star \in \mathop{\mathrm{arg\,min}}_{x \in \mathbb{R}^d} \; g(x) + h(x).
\end{equation*}
If $h$ were assumed differentiable, then the criterion
$f(x)=g(x)+h(x)$ is convex and differentiable, and repeating the
simple gradient descent steps
\begin{equation}
\label{eq::gradstep}
x^+ = x - \tau \nabla f(x)
\end{equation}
suffices to minimize $f$, for an appropriate choice of step size
$\tau$. (In the above, we write $x^+$ to denote the gradient descent
update from the current iterate $x$.) If $h$ is not
differentiable, then gradient descent obviously does not apply, but
as long as $h$ is ``simple'' (to be made precise
shortly), we can apply a variant of gradient descent that shares many
of its properties, called {\it proximal gradient descent}.
Proximal gradient descent is often also called composite or
generalized gradient descent, and in this routine we repeat the steps
\begin{equation}
\label{eq::proxstep}
x^+ = \mathrm{prox}_{h,\tau} \big( x - \tau \nabla g(x) \big)
\end{equation}
until convergence, where $\mathrm{prox}_{h,\tau} : \mathbb{R}^d \rightarrow \mathbb{R}^d$ is
the proximal mapping associated with $h$ (and $\tau$),
\begin{equation}
\label{eq::proxmap}
\mathrm{prox}_{h,\tau}(x) = \mathop{\mathrm{arg\,min}}_{z \in \mathbb{R}^d} \;
\frac{1}{2\tau} \|x-z\|_2^2 + h(z).
\end{equation}
(Strict convexity of the above criterion ensures that it has a unique
minimizer, so that the proximal mapping is well-defined.) Provided
that $h$ is simple, by which we mean that its proximal map
\eqref{eq::proxmap} is explicitly computable, the proximal gradient
descent steps \eqref{eq::proxstep} are straightforward and resemble the
classical gradient descent analogues \eqref{eq::gradstep}; we
simply take a gradient step in the direction governed by the smooth
part $g$, and then apply the proximal map of $h$. A slightly more
formal perspective argues that the updates \eqref{eq::proxmap} are the
result of minimizing $h$ plus a quadratic expansion of $g$, around the
current iterate $x$.
Proximal gradient descent has become a very popular tool for
optimization problems in statistics and machine learning, where
typically $g$ represents a smooth loss function, and $h$ a nonsmooth
regularizer. This trend is somewhat recent, even
though the study of proximal mappings has a long history of
in the optimization community (e.g., see \citet{prox} for a nice
review paper). In terms of convergence properties, proximal gradient
descent enjoys essentially the same convergence rates as gradient
descent under the analogous assumptions, and is amenable to
acceleration techniques just like gradient descent
(e.g., \citet{nestcomp}, \citet{fista}).
Of course, for proximal gradient
descent to be applicable in practice, one must be able to exactly (or
even approximately) compute the proximal map of $h$ in
\eqref{eq::proxmap}; fortunately, this is possible for many
optimization problems, i.e., many common regularizers $h$,
that are encountered in statistics. In our case, the proximal mapping
reduces to solving a problem of the form
\begin{equation}
\label{eq::1dfused}
\hat{\theta} = \mathop{\mathrm{arg\,min}}_{\theta} \frac{1}{2} \|x-\theta\|_2^2 +
\lambda_1 \sum_{i=1}^m |\theta_i| +
\lambda_2 \sum_{i=1}^{m-1} |\theta_i-\theta_{i+1}|.
\end{equation}
This is often called the fused lasso signal approximator (FLSA)
problem,
and extremely fast, linear-time algorithms exist to compute its
solution. In particular, we rely on an elegant dynamic
programming approach proposed by \citet{nickdp}.
\subsection{Application to the multinomial fused lasso problem}
\label{sec::proxapp}
The problem in \eqref{eq::fusedmodel} fits into the desired form for
proximal gradient descent, with $g$ the multinomial regression loss
(i.e., negative multinomial regression log likelihood) and $h$ the
lasso plus fused lasso penalties. Formally, we can rewrite
\eqref{eq::fusedmodel} as
\begin{equation}
\label{eq::fusedmodel2}
(\hat\beta_0, \hat\beta ) \in
\mathop{\mathrm{arg\,min}}_{\beta_0,\beta} \; g(\beta_0, \beta) + h(\beta_0,\beta),
\end{equation}
where $g$ is the convex, smooth function
\begin{equation*}
g(\beta_0, \beta) = \sum_{t=1}^T \sum_{i=1}^{n}
\left\{
\sum_{k=1}^{K-1} - \mathbb{I}(Y_{it}=k)
(\beta_{0tk}+X_{i\cdot t}\beta_{\cdot tk}) +
\log \left(1+\sum_{h=1}^{K-1} \exp(\beta_{0th} + X_{i\cdot
t}\beta_{\cdot th}) \right) \right\},
\end{equation*}
and $h$ is the convex, nonsmooth function
\begin{equation*}
h(\beta_0,\beta) = \lambda_1 \sum_{j=1}^p \sum_{t=1}^T
\sum_{k=1}^{K-1} |\beta_{jtk}| + \lambda_2 \sum_{j=1}^p
\sum_{t=1}^{T-1} \sum_{k=1}^{K-1} |\beta_{jtk}-\beta_{j(t+1)k}| .
\end{equation*}
Here we consider fixed values $\lambda_1,\lambda_2 \geq 0$. As
described previously, each of these tuning parameters will have a big
influence on the strength of their respective penalty terms, and hence
the properties of the computed estimate \smash{$(\hat\beta_0,\hat\beta)$};
we discuss the selection of $\lambda_1$ and $\lambda_2$ in
Section \ref{sec::selection}. We note that the intercept coefficients
$\beta_0$ are not penalized.
To compute the proximal gradient updates, as given in
\eqref{eq::proxstep}, we must consider two quantities: the
gradient of $g$, and the proximal map of $h$. First, we discuss the
gradient. As $\beta_0 \in \mathbb{R}^{T\times (K-1)}$,
$\beta \in \mathbb{R}^{p\times T \times (K-1)}$, we may consider the gradient
as having dimension
$\nabla g(\beta_0,\beta) \in \mathbb{R}^{(p+1) \times T \times (K-1)}$. We
will index this as $[\nabla g(\beta_0,\beta)]_{jtk}$ for $j=0,\ldots
p$, $t=1,\ldots T$, $k=1,\ldots K-1$; hence note that $[\nabla
g(\beta_0,\beta)]_{0tk}$ gives the partial derivative of $g$ with
respect to $\beta_{0tk}$, and $[\nabla g(\beta_0,\beta)]_{jtk}$ the
partial derivative with respect to $\beta_{jtk}$, for $j=1,\ldots p$.
For generic $t,k$, we have
\begin{equation}
\label{eq::gradient0}
[\nabla g(\beta_0,\beta)]_{0tk}
=
\sum_{i=1}^{n} \left( -\mathbb{I}(Y_{it}=k) +
\frac{ \exp(\beta_{0tk} + X_{i\cdot t}
\beta_{\cdot tk})}
{1+\sum_{h=1}^{K-1} \exp(\beta_{0th} +
X_{i\cdot t}\beta_{\cdot th})}\right),
\end{equation}
and for $j \geq 1$,
\label{eq::gradient}
\begin{equation}
[\nabla g(\beta_0,\beta)]_{jtk}
= \sum_{i=1}^{n}
\left( -\mathbb{I}(Y_{it}=k) X_{ijt} +
X_{ijt}\frac{ \exp(\beta_{0tk} +
X_{i\cdot t}\beta_{\cdot tk})}
{1+\sum_{h=1}^{K-1} \exp(\beta_{0th} +
X_{i\cdot t}\beta_{\cdot th})}\right).
\end{equation}
It is evident that computation of the gradient requires $O(npTK)$.
Now, we discuss the proximal operator. Since the intercept
coefficients $\beta_0 \in \mathbb{R}^{T\times (K-1)}$ are left unpenalized,
the proximal map over $\beta_0$ just reduces to the identity, and the
intercept terms undergo the updates
\begin{equation*}
\beta_{0tk}^+ = \beta_{0tk} - \tau [\nabla g(\beta_0,\beta)]_{0tk}
\;\;\;\text{for}\;\, t=1,\ldots T, \; k=1, \ldots K-1.
\end{equation*}
Hence we consider the proximal map over $\beta$ alone.
At an arbitrary input $x \in \mathbb{R}^{p \times T \times (K-1)}$, this is
\begin{equation*}
\mathop{\mathrm{arg\,min}}_{z \in \mathbb{R}^{p\times T \times (K-1)}} \;
\frac{1}{2\tau} \sum_{j=1}^p \sum_{t=1}^T
\sum_{k=1}^{K-1} (x_{jtk} - z_{jtk})^2 +
\lambda_1 \sum_{j=1}^p \sum_{t=1}^T
\sum_{k=1}^{K-1} |\beta_{jtk}| + \lambda_2 \sum_{j=1}^p
\sum_{t=1}^{T-1} \sum_{k=1}^{K-1} |\beta_{jtk}-\beta_{j(t+1)k}|,
\end{equation*}
which we can see decouples into $p(K-1)$ separate minimizations, one
for each predictor $j=1,\ldots p$ and class $k=1,\ldots K-1$. In
other words, the coefficients $\beta$ undergo the updates
\begin{multline}
\label{eq::bupdate}
\beta^+_{j\cdot k} =
\mathop{\mathrm{arg\,min}}_{\theta \in \mathbb{R}^T} \frac{1}{2}
\sum_{t=1}^T \Big(\big(\beta_{j\cdot k} - \tau
[\nabla g(\beta_0,\beta)]_{j\cdot k}\big) - \theta \Big)^2 +
\tau \lambda_1 \sum_{t=1}^T |\theta_t| +\tau \lambda_2
\sum_{t=1}^{T-1} |\theta_t-\theta_{t+1}|,\\
\text{for}\;\, j=1,\ldots p, \; k=1,\ldots K-1,
\end{multline}
each minimization being a fused lasso signal approximator problem
\citep{fuse}, i.e., of the form \eqref{eq::1dfused}. There are many
computational approaches that may be applied to such a problem
structure; we employ a specialized, highly efficient algorithm by
\citet{nickdp} that is based on dynamic programming. This algorithm
requires $O(T)$ operations for each of the problems in
\eqref{eq::bupdate}, making the total cost of the update $O(pTK)$
operations. Note that this is actually dwarfed by the cost of
computing the gradient $\nabla g(\beta_0,\beta)$ in the first place,
and therefore the total complexity of a single iteration of our proposed
proximal gradient descent algorithm is $O(npTK)$.
\subsection{Practical considerations}
\label{sec::practical}
We discuss several practical issues that arise in applying the
proximal gradient descent algorithm.
\subsubsection{Backtracking line search}
Returning to the generic perpsective for proximal gradient descent as
described in Section \ref{sec::proxintro}, we rewrite the proximal
gradient descent update in \eqref{eq::proxstep} as
\begin{equation}
\label{eq::genstep}
x^+ = x - \tau G_\tau (x),
\end{equation}
where $G_\tau(x)$ is called the {\it generalized gradient} and is
defined as
\begin{equation*}
G_\tau = \frac{x - \mathrm{prox}_{h,\tau}(x-\tau\nabla g(x))}{\tau}.
\end{equation*}
The update is rewritten in this way so that it more closely resembles
the usual gradient update in \eqref{eq::gradstep}. We can see that,
analogous to the gradient descent case, the choice of parameter
$\tau>0$ in
each iteration of proximal gradient descent determines the magnitude
of the update in the direction of the generalized gradient
$G_\tau(x)$. Classical analysis shows that if $\nabla g$ is
Lipschitz with constant $L>0$, then proximal gradient descent
converges with any fixed choice of step size $\tau \leq 1/L$
across all iterations. In most practical situations, however, the
Lipschitz constant $L$ of $\nabla g$ is not known or easily
computable, and we rely on an adaptive scheme for choosing an
appropriate step
size at each iteration; backtracking line search is one such scheme,
which is straightforward to implement in practice and guarantees
convergence of the algorithm under the same Lipschitz assumption on
$\nabla g$ (but importantly, without having to know its Lipschitz
constant $L$). Given a shrinkage factor $0 < \gamma < 1$, the
backtracking line search routine at a
given iteration of proximal gradient descent starts with $\tau=\tau_0$
(a large initial guess for the step size), and while
\begin{equation}
\label{eq::backtrack}
g\big(x-\tau G_\tau(x)\big) > g(x) - \tau \nabla g(x)^T G_\tau(x) +
\frac{\tau}{2}\|G_\tau(x)\|_2^2,
\end{equation}
it shrinks the step size by letting $\tau=\gamma \tau$. Once the exit
criterion is achieved (i.e., the above is no longer satisfied), the
proximal gradient descent algorithm then uses the current value of
$\tau$ to take an update step, as in \eqref{eq::genstep} (or
\eqref{eq::proxstep}).
In the case of the multinomial fused lasso problem, the generalized
gradient is of dimension $G_\tau(\beta_0,\beta) \in
\mathbb{R}^{(p+1)\times T\times (K-1)}$, where
\begin{equation*}
[G_\tau(\beta_0,\beta)]_{0\cdot\cdot} =
[\nabla g(\beta_0,\beta)]_{0\cdot\cdot},
\end{equation*}
and
\begin{equation*}
[G_\tau(\beta_0,\beta)]_{j\cdot k} =
\frac{\beta_{j\cdot k} - \mathrm{prox}_{\mathrm{FLSA},\tau}
(\beta_{j \cdot k} -\tau [\nabla g(\beta_0,\beta)]_{j\cdot k})}
{\tau} \;\;\; \text{for} \;\,
j=1,\ldots p, \;k=1,\ldots K-1.
\end{equation*}
Here $\mathrm{prox}_{\mathrm{FLSA},\tau}
(\beta_{j \cdot k} -\tau [\nabla g(\beta_0,\beta)]_{j\cdot k})$
is the proximal map defined by the fused lasso signal approximator
evaluated at $\beta_{j \cdot k} -\tau [\nabla
g(\beta_0,\beta)]_{j\cdot k}$, i.e., the right-hand side in
\eqref{eq::bupdate}. Backtracking line search now applies just as
described above.
\subsubsection{Stopping criteria}
The simplest implementation of proximal gradient descent would
run the algorithm for a fixed, large number of steps $S$. A more
refined approach would check a stopping criterion at the end of each
step, and terminate if such a criterion is met. Given a
tolerance level $\epsilon>0$, two common stopping criteria are then
based on the relative difference in function values, as in
\begin{equation*}
\text{stopping criterion 1: terminate if}\;\
C_1 = \frac{| f(\beta_0^+,\beta^+) - f(\beta_0,\beta) |}
{f(\beta_0,\beta)} \leq \epsilon,
\end{equation*}
and the relative difference in iterates, as in
\begin{equation*}
\text{stopping criterion 2: terminate if}\;\,
C_2 = \frac{\|( \beta_0^+,\beta^+) - (\beta_0,\beta)\|_2}
{\|(\beta_0,\beta)\|_2} \leq \epsilon.
\end{equation*}
The second stopping criterion is generally more stringent, and may be
hard to meet in large problems, given a small tolerance $\epsilon$.
For the sake of completeness, we outline the full proximal gradient
descent procedure in the notation of the multinomial fused lasso
problem, with backtracking line search and the first stopping
criterion, in Algorithms \ref{alg::proxgrad} and \ref{alg::backtracking} below.
\begin{algorithm}[!htb]
\begin{algorithmic}[1]
\INPUT Predictors $X$, outcomes $Y$, tuning parameter values
$\lambda_1,\lambda_2$, initial coefficient guesses
\smash{$(\beta_0^{(0)},\beta^{(0)})$}, maximum number of iterations $S$,
initial step size before backtracking $\tau_0$, backtracking shrinkage
parameter $\gamma$, tolerance $\epsilon$
\OUTPUT Approximate solution \smash{$(\hat\beta_0,\hat\beta)$}
\State $s = 1$, $C=\infty$
\While{($s \leq S$ and $C > \epsilon$)}
\State Find $\tau_s$ using backtracking,
Algorithm \ref{alg::backtracking}
(INPUT: \smash{$\beta_0^{(s-1)}, \beta^{(s-1)}, \tau_0, \gamma$})
\State Update the intercept:
\smash{$\beta_{0\cdot \cdot}^{(s)} =
\beta^{(s-1)}_{0\cdot \cdot} -
\tau_s [\nabla g(\beta_0^{(s-1)},\beta^{(s-1)})]_{0\cdot\cdot}$}
\For{$j=1,\ldots p$}
\For{$k=1,\ldots (K-1)$}
\State Update
\smash{$\beta^{(s)}_{j\cdot k} = \mathrm{prox}_{\mathrm{FLSA},\tau_s}
(\beta^{(s-1)}_{j \cdot k} -\tau_s [\nabla
g(\beta_0^{(s-1)},\beta^{(s-1)})]_{j\cdot k})$}
\EndFor
\EndFor
\State Increment $s = s+1$
\State Compute \smash{$C =
[f(\beta_0^{(s)},\beta^{(s)}) - f(\beta_0^{(s-1)},\beta^{(s-1)})]/
f(\beta_0^{(s-1)},\beta^{(s-1)})$}
\EndWhile
\State
\smash{$\hat\beta_0 = \beta_0^{(s)}$, $\hat\beta = \beta^{(s)}$} \\
\Return \smash{$(\beta_0, \hat\beta)$}
\end{algorithmic}
\caption{Proximal gradient descent for the multinomial fused lasso}
\label{alg::proxgrad}
\end{algorithm}
\begin{algorithm}[!htb]
\begin{algorithmic}[1]
\INPUT $\beta_0,\beta, \tau_0,\gamma$
\OUTPUT $\tau$
\State $\tau = \tau_0$
\While{(true)}
\State Compute
$[G_\tau(\beta_0,\beta)]_{0\cdot\cdot} =
[\nabla g(\beta_0,\beta)]_{0\cdot\cdot}$
\For{$j=1,\ldots p$}
\For{$k=1,\ldots (K-1)$}
\State Compute
$[G_\tau(\beta_0,\beta)]_{j\cdot k} =
[\beta_{j\cdot k} - \mathrm{prox}_{\mathrm{FLSA},\tau}
(\beta_{j \cdot k} -\tau [\nabla g(\beta_0,\beta)]_{j\cdot k})]/
\tau$
\EndFor
\EndFor
\If{$g((\beta_0,\beta) -\tau G_\tau(\beta_0,\beta)) >
g(\beta_0,\beta) - \tau [\nabla g(\beta_0,\beta)]^T
G_\tau(\beta_0,\beta) + \frac{\tau}{2}
\|G_\tau(\beta_0,\beta)\|_2^2$}
\State Break
\Else
\State Shrink $\tau = \gamma \tau$
\EndIf
\EndWhile \\
\Return $\tau$
\end{algorithmic}
\caption{Backtracking line search for the multinomial fused lasso}
\label{alg::backtracking}
\end{algorithm}
\subsubsection{Missing individuals}
\label{sec::missingIndiv}
Often in practice, some individuals are not present at some
timepoints in the longitudinal study, meaning that one or both of
their outcome values and predictor measurements are missing over
a subset of $t=1,\ldots T$. Let $I_t$ denote the set of
completely observed individuals (i.e., with both predictor
measurements and outcomes observed) at time $t$, and let $n_t=|I_t|$.
The simplest strategy to accomodate such missingness would be to
compute the loss function $g$ only observed individuals, so that
\begin{equation*}
g(\beta_0, \beta) = \sum_{t=1}^T \sum_{i \in I_t}
\left\{
\sum_{k=1}^{K-1} - \mathbb{I}(Y_{it}=k)
(\beta_{0tk}+X_{i\cdot t}\beta_{\cdot tk}) +
\log \left(1+\sum_{h=1}^{K-1} \exp(\beta_{0th} + X_{i\cdot
t}\beta_{\cdot th}) \right) \right\}.
\end{equation*}
An issue arises when the effective sample size $n_t$ is quite variable
across timepoints $t$: in this case, the penalty terms can have quite
different effects on the
coefficients $\beta_{\cdot\cdot t}$ at one time $t$ versus
another. That is, the coefficients $\beta_{\cdot\cdot t}$ at a time
$t$ in which $n_t$ is small experience a relatively small loss term
\begin{equation}
\label{eq::gt}
\sum_{i \in I_t}
\left\{
\sum_{k=1}^{K-1} - \mathbb{I}(Y_{it}=k)
(\beta_{0tk}+X_{i\cdot t}\beta_{\cdot tk}) +
\log \left(1+\sum_{h=1}^{K-1} \exp(\beta_{0th} + X_{i\cdot
t}\beta_{\cdot th}) \right) \right\},
\end{equation}
simply because there are fewer terms in the above sum compared to a
time with a larger effective sample size; however, the penalty term
\begin{equation*}
\lambda_1 \sum_{j=1}^p
\sum_{k=1}^{K-1} |\beta_{jtk}| + \lambda_2 \sum_{j=1}^p
\sum_{k=1}^{K-1} |\beta_{jtk}-\beta_{j(t+1)k}|
\end{equation*}
remains comparable across all timepoints, regardless of sample size.
A fix would be to scale the loss term in \eqref{eq::gt} by $n_t$ to
make it (roughly) independent of the effective sample size, so that
the total loss becomes
\begin{equation}
\label{eq:gnew}
g(\beta_0,\beta) =
\sum_{t=1}^T
\frac{1}{n_t}
\sum_{i \in I_t}
\left\{
\sum_{k=1}^{K-1} - \mathbb{I}(Y_{it}=k)
(\beta_{0tk}+X_{i\cdot t}\beta_{\cdot tk}) +
\log \left(1+\sum_{h=1}^{K-1} \exp(\beta_{0th} + X_{i\cdot
t}\beta_{\cdot th}) \right) \right\}.
\end{equation}
This modification indeed ends up being important for the
Alzheimer's analysis that we present in Section \ref{sec::CHS}, since
this study has a number of individuals in the tens at some timepoints,
and in the hundreds for others. The proximal gradient descent
algorithm described in this section extends to cover the loss in
\eqref{eq:gnew} with only trivial modifications.
\subsection{Implementation in C++ and R}
An efficient C++ implementation of the proximal gradient descent
algorithm described in this section, with an easy interface to R, is
available from the second author's website:
\url{http://www.stat.cmu.edu/~flecci}. In the future, this will be
available as part of the R package {\tt glmgen}, which broadly fits
generalized linear models under generalized lasso regularization.
\section{Alzheimer's Disease data analysis}
\label{sec::CHS}
In this section, we apply the proposed estimation method to the data
of the the Cardiovascular Health Study Cognition Study (CHS-CS), a
rich database of thousands of multiple cognitive, metabolic,
cardiovascular, cerebrovascular, and neuroimaging variables obtained
over the past 24 years for people of ages 65 to 110 years old
\citep{fried1991cardiovascular, lopez2007incidence}.
The complex relationships between age and other risk factors produce
highly variable natural histories from normal cognition to the
clinical expression of Alzheimer's disease, either as dementia or its
prodromal syndrome, mild cognitive impairment (MCI)
\citep{lopez2003prevalence, saxton2004preclinical, lopez2007incidence,
sweet2012effect, 2014mixed}.
Many studies involving the CHS-CS data have shown the importance of a
range of risk factors in predicting the time of onset of clinical
dementia. The risk of dementia is affected by the presence of the
APOE*4 allele, male sex, lower education, and having a family history
of dementia \citep{fitzpatrick2004incidence, tang1996relative,
launer1999rates}. Medical risks include the presence of systemic
hypertension, diabetes mellitus, and cardiovascular or cerebrovascular
disease \citep{kuller2003risk, irie2005type, skoog199615}. Lifestyle
factors affecting risk include physical and cognitive activity, and
diet \citep{verghese2003leisure, erickson2010physical,
scarmeas2006mediterranean}.
A wide range of statistical approaches has been considered in these
studies, including exploratory statistical summaries, hypothesis
tests, survival analyses, logistic regression models, and latent
trajectory models.
None of these methods can directly accommodate a large number of
predictors that can potentially exceed the number of observations.
A small number of variables was often chosen a priori to match the
requirements of a particular model, neglecting the full potential
of the CHS-CS data, which consists of thousands of variables.
The approach that we introduced in Section \ref{sec::introduction} can
accommodate an array of predictors of arbitrary dimension, using
regularization to maintain a well-defined predictive model and avoid
overfitting. Our goal is to identify important risk factors for
the prediction of the cognitive status at $t+10$ years of age
($\Delta=10$), given predictor measurements at $t$ years of age, for
$t=65,66, \ldots ,98$.
We use the penalized log likelihood criterion in
\eqref{eq::fusedmodel} to estimate the coefficients of the multinomial
logit model in \eqref{eq::multinomial}. The lasso penalty forces the
solution to be sparse, allowing us to identify a few important
predictors among the thousands of variables of the CHS-CS data.
The fused lasso penalty allows for a few change points in
the piecewise constant coefficient trajectories
\smash{$\hat\beta_{j \cdot k}$}, across $t=65,\ldots 98$. Justification
for this second penalty is based on the scientific intuition that
predictors that are clinically important should have similar effects
in successive ages.
\subsection{Data preprocessing}
We use data from the $n=924$ individuals in the Pittsburgh
section of the CHS-CS, recorded between 1990 and 2012.
Each individual underwent clinical and cognitive assessments at
multiple ages, all falling in the range $65, \ldots 108$.
The matrix of (future) outcomes $Y$ has dimension $n \times 34$:
for $i=1,\ldots 924$ and $t=65,\ldots 98$, the outcome $Y_{it}$
stores the cognitive status at age $t+10$ and can assume one of the
following values:
\begin{equation*}
Y_{it}=
\begin{cases}
1 & \text{if normal}\\
2 & \text{if MCI/dementia}\\
3 & \text{if dead}
\end{cases}.
\end{equation*}
MCI is included in the same class as dementia,
as they are both instances of cognitive impairment.
Hence the proposed multinomial model predicts the onset of
MCI/dementia, in the presence of a separate death category.
This is done to implicitly adjust for the confounding effect of death,
as some risk factors for dementia are also known to be risk factors
for death \citep{rosvall2009apoe}.
The array of predictors $X$ is composed of time-varying variables that
were recorded at least twice during the CHS-CS study, and
time-invariant variables, such as gender and race. A complication in
the data set is the ample amount of missingness in the array of
predictors. We impute missing values using a uniform rule for all
possible causes of missingness. A missing value at age $t$ is imputed
by taking the closest past measurement from the same individual, if
present. If all the past values are missing, the global median from
people of age $t$ is used. The only exception is the case of
time-invariant predictors, whose missing values are imputed by either
future or past values, as available.
Categorical variables with $m$ possible outcomes are converted to
$m-1$ binary variables and all the predictors are standardized to have
zero mean and unit standard deviation. This is a standard procedure in
regularization, as the lasso and fused lasso penalties puts
constraints on the size of the coefficients associated with each
variable \citep{tibshirani1997lasso}.
To be precise, imputation of missing values and standardization of the
predictors are performed within each of the folds used in the
cross-validation method for the choice of the tuning parameters
$\lambda_1$ and $\lambda_2$ (discussed below), and then again for the
full data set in the final estimation procedure that uses the selected
tuning parameters.
The final array of predictors $X$ has dimension $924 \times 1050
\times 34$, where $1050$ is the number of variables recorded over the
period of 34 years of age range.
\subsection{Model and algorithm specification}
In the Alzheimer's Disease application, the multinomial model
in \eqref{eq::multinomial} is determined by two equations, as there
are three possible outcomes (normal, MCI/ dementia, death); the
outcome ``normal'' is taken as the base class. We will refer to the
two equations (and the corresponding sets of coefficients) as the
``dementia vs normal'' and ``death vs normal'' equations,
respectively.
We use the proximal gradient descent algorithm described in Section
\ref{sec::GGD} to estimate the coefficients
that maximize the penalized log likelihood criterion in
\eqref{eq::fusedmodel}. The initializations \smash{$(\beta_0^{(0)},
\beta^{(0)})$} are set to be zero matrices, the maximum number of
iterations is $S=80$, the initial step size before backtracking is
$\tau_0=20$, the backtracking shrinkage parameter is $\gamma=0.6$ and
the tolerance of the first stopping criterion (relative difference in
function values) is $\epsilon=0.001$.
We select the tuning parameters by a 4-fold cross-validation procedure
that minimizes the misclassification error.
The selected parameters are $\lambda_1 = 0.019$ and $\lambda_2 =
0.072$, which yield an average prediction error of 0.316 (standard
error 0.009).
Section \ref{sec::selection} discusses more details on the model
selection problem.
The number $n_t$ of outcomes observed at age $t$ varies across time,
for two reasons: first, different subjects entered the study at
different ages, and second, once a subject dies at time $t_0$, we
exclude them consideration in the model formed at all ages $t>t_0$, to
predict the outcomes of individuals at age $t+10$.
The maximum number of outcomes is 604 at age 88, whereas the minimum
is 7 at age 108. We resort to the strategy described in Section
\ref{sec::missingIndiv} and use the scaled loss in \eqref{eq:gnew} to
compensate for the varying sample sizes.
\subsection{Results}
Out of the 1050 coefficients associated with the predictors described
above, 148 are estimated to be nonzero for at least one time point in
the 34 years age range. More precisely, for at least one age, 57
coefficients are nonzero in the ``dementia vs normal'' equation of the
predictive multinomial logit model, and 124 are nonzero in the ``death
vs normal'' equation.
\begin{figure}[!ht]
\centering
\includegraphics[width=\textwidth]{figs/results2.pdf}
\caption{CHS-CS data analysis. Left: relative importance plots for the
15 most important variables in the ``dementia vs normal'' and
``death vs normal'' equations of the multinomial logit model. Right:
corresponding estimated coefficients. The order of the legends
follow the order of the maximum/minimum values of the estimated
coefficient trajectories. Note that some coefficients are estimated to be
very close to 0 and the corresponding trajectories are hidden by other coefficients.}
\label{fig::results}
\end{figure}
\begin{table}[!ht]
\centering
\begin{tabular}{| c | l |}
\multicolumn{2}{c}{Dementia vs normal} \\ \hline
\textbf{Variable} & \textbf{Meaning (and coding for categorical
variables, before scaling)} \\ \hline
race01.2 & Race: "White" 1, else 0 \\ \hline
cdays59 & Taken vitamin C in the last 2 weeks? (number of days) \\ \hline
newthg68.1 & How is the person at learning new things wrt 10 yrs ago? "A bit worse" 1, else 0 \\ \hline
estrop39 & If you not currently taking estrogen, have you taken in the past? "Yes" 1, "No" 0 \\ \hline
fear05.1 & How often felt fearful during last week? "Most of the time" 1, else 0 \\ \hline
early39 & Do you usually wake up far too early? "Yes" 1, "No" 0 \\ \hline
gend01 & Gender: "Female" 1, "Male" 0 \\ \hline
hctz06 & Medication: thiazide diuretics w/o K-sparing. "Yes" 1, "No" 0 \\ \hline
race01.1 & Race: "Other (no white, no black)" 1, else 0 \\ \hline
orthos27 & Do you use a lower extremity orthosis? "Yes" 1, "No" 0\\ \hline
pulse21 & 60 second heart rate \\ \hline
grpsym09.1 & What causes difficulty in gripping? "Pain in arm/hand" 1, else 0 \\ \hline
sick03.2 & If sick, could easily find someone to help? "Probably False" 1, else 0 \\ \hline
digcor & Digit-symbol substitution task: number of symbols correctly coded \\ \hline
trust03.3 & There is at at least one person whose advice you really trust. "Probably true" 1, else 0 \\ \hline
\multicolumn{2}{c}{} \\
\multicolumn{2}{c}{Death vs normal} \\ \hline
\textbf{Variable} & \textbf{Meaning (and coding for categorical
variables, before scaling)} \\ \hline
digcor & Digit-symbol substitution task: number of symbols correctly coded \\ \hline
ctime27 & Repeated chair stands: number of seconds \\ \hline
gend01 & Gender: "Female" 1, "Male" 0 \\ \hline
cis42 & Cardiac injury score \\ \hline
hurry59.2 & Ever had pain in chest when walking uphill/hurry? "No" 1, else 0 \\ \hline
numcig59 & Number of cigarettes smoked per day \\ \hline
dig06 & Digitalis medicines prescripted? "Yes" 1, "No" 0 \\ \hline
smoke.3 & Current smoke status: "Never smoked" 1, else 0 \\ \hline
hlth159.1 & Would you say, in general, your health is.. ? "Fair" 1, else 0 \\ \hline
exer59 & If gained/lost weight, was exercise a major factor? "Yes" 1, "No" 0\\ \hline
nomeds06 & Number of medications taken\\ \hline
diabada.3 & ADA diabetic status? "New diabetes" 1, else 0\\ \hline
anyone & Does anyone living with you smoke cigarettes regularly? "Yes" 1, "No" 0\\ \hline
ltaai & Blood pressure variable: left ankle-arm index \\ \hline
whmile09.2 & Do you have difficulty walking one-half a mile? "Yes" 1, else 0 \\ \hline
\end{tabular}
\caption{The 15 most important variables in the two separate equations
of the multinomial logit model.}
\label{tab::variables}
\end{table}
Figure \ref{fig::results} shows the 15 most important variables in the
34 years age range, separately for the two equations. The measure of
importance is described in detail in Section \ref{sec::stability} and
is, in fact, a measure of stability of the estimated coefficients,
across 4 subsets of the data (the 4 training sets used
in cross-validation). The plots on the left show the relative
importance of the 15 variables with respect to the most important one,
whose importance was scaled to be 100. The plots on the right show,
separately for the two equations, the longitudinal estimated
coefficients for the 15 most important variables, using the data and
algorithm specification described above. The meaning of these
predictors and the coding used for the categorical variables are
reported in Table \ref{tab::variables}.
The nonzero coefficients that are not displayed in Figure
\ref{fig::results} are less important (according to our measure of
stability) and, for the vast majority, their absolute values are less
than 0.1.
We now proceed to interpret the results, keeping in mind that,
ultimately, we are estimating the coefficients of a multinomial logit
model and that the outcome variable is recorded 10 years in the future
with respect to the predictors.
For example, an increase in the value of a predictor with positive
estimated coefficient in the top right plot of Figure
\ref{fig::results} is associated with an increase of the (10 years
future) odds of dementia with respect to a normal cognitive status.
In what follows, to facilitate the exposition of results, our
statements are less formal.
Inspecting the ``dementia vs normal'' plot we see that,
in general, being Caucasian ({\tt race01.2}) is associated with a
decrease in the odds of dementia, while, after the age of 85, fear
({\tt fear05.1}), lack of available caretakers ({\tt sick03.2}), and
deterioration of learning skills ({\tt newthg68.1}) increase the odds
of dementia. Variables {\tt hctz06} (a particular diuretic) and {\tt
early39} (early wake-ups) have positive coefficients for the age
ranges $65, \ldots 78$ and $77, \ldots 91$, respectively, and hence,
if active, they account for an increase of the risk of dementia.
The ``death vs normal" plot reveals the importance of several
variables in the age range $65, \ldots 85$: longer time to rise from sitting in a chair
({\tt ctime27}), more cigarettes ({\tt numcig59}), higher
cardiac injury score ({\tt cis42}) are associated with an increase of
the odds of death. Other variables in the same age range, with analogous
interpretations, but lower importance, are {\tt diabada.3} ("new
diabetes"' diagnosis), {\tt hlth159.1} ("fair" health status), {\tt
dig06} (use of Digitalis), {\tt whmile09.2} (difficulty in
walking).
By contrast, in the same age range, good performance on the
digit-symbol substitution task ({\tt digcor}) accounts for a decrease
in the odds of death. Finally, regardless of the age, being a
non-smoker ({\tt smoke.3}) or being a woman ({\tt gend01}) decrease
the odds of death.
\begin{figure}[!ht]
\centering
\includegraphics[width=\textwidth]{figs/intercepts.pdf}
\caption{CHS-CS data analysis. Estimated intercept coefficients in the
two separate equations of the multinomial logit model.}
\label{fig::intercepts}
\end{figure}
Figure \ref{fig::intercepts} shows the intercept coefficients
\smash{$\hat \beta_{0 \cdot 1}$} and \smash{$\hat \beta_{0 \cdot 2}$},
which, we recall, are not penalized in the log likelihood criterion in
\eqref{eq::fusedmodel}. The intercepts account for time-varying
risk that is not explained by the predictors. In particular, the
coefficients $\hat \beta_{0 \cdot 2}$ increases over time, suggesting
that an increasing amount of risk of death can be attributed to a
subject's age alone, independent of the predictor measurements.
\subsection{Discussion}
The results of the proposed multinomial fused lasso methodology
applied to the CHS data are broadly consistent with what is known
about risk and protective factors for dementia in the elderly
\citep{lopez2013patterns}. Race, gender, vascular and heart disease,
lack of available caregivers, and deterioration of learning and memory
are all associated with an increased risk of dementia.
The results, however, provide critical new insights into the natural
progression of MCI/dementia. First, the relative importance of the
risk factors changes over time. As shown in Figure \ref{fig::results},
with the exception of race, risk factors for dementia become more
relevant after the age of 85. This is critical, as there is increasing
evidence \citep{kuller2011does} for a change in the risk profile for
the expression of clinical dementia among the oldest-old.
Second, the independent prediction of death, and the associated
risk/protection factors, highlight the close connection between risk
of death and risk of dementia. That is, performance on a simple, timed
test of psychomotor speed (digit symbol substitution task) is a very
powerful predictor of death within 10 years, as is a measure of
physical strength/frailty (time to arise from a chair). Other
variables, including gender, diabetes, walking and exercise, are all
predictors of death, but are known, from other analyses in the CHS and
other studies, to be linked to the risk of dementia. The importance of
these risk/protective factors for death is attenuated (with the
exception of gender) after age 85, likely reflecting survivor bias.
Taken together, these results add to the growing body of evidence
of the critical importance of accounting for
mortality in the analysis of risk for dementia, especially among the
oldest old \citep{kuller2011does}.
For our analysis we chose a 10 year time window for risk
prediction. Among individuals of age 65-75, who are cognitively
normal, this may be a scientifically and clinically reasonable time
window to use. However, had we similar data from individuals as young
as 45-50 years old, then we might wish to choose time windows of 20
years or longer. In the present case, it could be argued that a
shorter time window might be more scientifically and clinically
relevant among individuals over the age of 80 years, as survival times
of 10 years become increasingly less likely in the oldest-old.
\section{Measures of stability}
\label{sec::stability}
Examining the stability of variables in a fitted model, subject to
small perturbations of the data set, is one way to assess variable
importance. Applications of stability, in this spirit, have recently
gained popularity in the literature, across a variety of settings such
as clustering (e.g., \citet{stabcluster}), regression (e.g.,
\citet{stabselect}), and graphical models (e.g., \citet{stars}). Here
we propose a very simple stability-based measure of variable
importance, based on the definition of variable importance for trees
and additive tree expansions \citep{cart,esl}. We fit the multinomial
fused lasso estimate \eqref{eq::fusedmodel} on the data set
$X_{i\cdot\cdot}$, $Y_{i\cdot}$, for $i=i_1,\ldots i_m$, a subsample
of the total individuals $1,\ldots n$, and repeat this process $R$
times. Let \smash{$\hat\beta^{(r)}$} denote the coefficients from the
$r$th subsampled data set, for $r=1,\ldots R$. Then we define the
importance of variable $j$ for class $k$ as
\begin{equation}
\label{eq::ij}
I_{jk} = \frac{1}{RT} \sum_{r=1}^R \sum_{t=1}^T
|\hat\beta_{jtk}^{(r)}|,
\end{equation}
for each $j=1,\ldots p$ and $k=1, \ldots K-1$, which is the average
absolute magnitude of the coefficients for the $j$th variable and
$k$th class, across all timepoints, and subsampled data sets.
Therefore, a larger value of $I_{jk}$ indicates a higher variable
importance, as measured by stability (not only across subsampled data
sets $r$, but actually across timepoints $t$, as well). Relative
importances can be computed by scaling the highest variable importance
to be 100, and adjusting the other values accordingly; for simplicity
we typically consider relative variable importances in favor of
absolute ones, because the original scale has no real meaning.
There is some subtlety in the role of the tuning parameters
$\lambda_1,\lambda_2$ used to fit the coefficients
\smash{$\hat\beta^{(r)}$} on each subsampled data set $r=1,\ldots
R$. Note that the importance measure \eqref{eq::ij} reflects the
importance of a variable in the context of a fitting procedure that,
given data samples, produces estimated coefficients. The simplest
approach would be to consider the fitting procedure defined by the
multinomial fused lasso problem \eqref{eq::fusedmodel} at a fixed pair
of tuning parameter values $\lambda_1,\lambda_2$. But in practice, it
is seldom true that appropriate tuning parameter values are known
ahead of time, and one typically employs a method like
cross-validation to select parameter values (see Section
\ref{sec::selection} for a discussion of cross-validation and other
model selection methods). Hence in this case, to determine variable
importances in the final coefficient estimates, we would take care to
define our fitting procedure in \eqref{eq::ij}
to be the one that, given data samples, performs cross-validation on
these data samples to determine the best choice of
$\lambda_1,\lambda_2$, and then uses this choice to fit coefficient
estimates. In other words, for each subsampled data set $r=1,\ldots
R$ in \eqref{eq::ij}, we would perform cross-validation to determine
tuning parameter values and then compute \smash{$\hat\beta^{(r)}$} as the
multinomial fused lasso solution at these chosen parameter
values. This is more computationally demanding, but it is a more
accurate reflection of variable importance in the final model output
by the multinomial fused lasso under cross-validation for model
selection.
The relative variable importances for the CHS-CS data example from
Section \ref{sec::CHS} are displayed in Figure \ref{fig::results},
alongside the plots of estimated coefficients. Here we drew 4
subsampled data sets, each one containing 75\% of the total number of
individuals. The tuning parameter values have been selected by
cross-validation. The variable importances were defined to incorporate
this selection step into the fitting procedure, as explained
above. We observe that the variables with high positive or negative
coefficients for most ages in the plotted trajectories typically also
have among the highest relative importances. Another interesting
observation concerns categorical predictors, which (recall) have been
converted into binary predictors over multiple levels: often only some
levels of a categorical predictor are active in the plotted
trajectories.
\section{Model selection}
\label{sec::selection}
The selection of tuning parameters $\lambda_1,\lambda_2$ is clearly an
important issue that we have not yet covered. In this section, we
discuss various methods for automatic tuning parameter selection in
the multinomial fused lasso model \eqref{eq::fusedmodel}, and apply
them to a subset of the CHS-CS study data of Section \ref{sec::CHS},
with 140 predictors and 600 randomly selected individuals,
as an illustration. In particular, we consider the following methods
for model selection: cross-validation, cross-validation under the
one-standard-error rule, AIC, BIC, and finally AIC and BIC using
misclassification loss (in place of the usual negative log
likelihood). Note that cross-validation in our longitudinal setting
is performed by dividing the individuals $1,\ldots n$ into folds, and,
per its typical usage, selecting the tuning parameter pair
$\lambda_1,\lambda_2$ (over, say, a grid of possible values) that
minimizes the cross-validation misclassification loss. The
one-standard-error rule, on the other hand, picks the simplest
estimate that achieves a cross-validation misclassification loss
within one standard error of the minimum. Here ``simplest'' is
interpreted to mean the estimate with the fewest number of nonzero
component blocks. AIC and BIC scores are computed for a candidate
$\lambda_1,\lambda_2$ pair by
\begin{align*}
\mathrm{AIC}(\lambda_1,\lambda_2) &= 2 \cdot
\mathrm{loss}\big((\hat\beta_0,\hat\beta)_{\lambda_1,\lambda_2}\big) \,+\,
2 \cdot \mathrm{df}\big((\hat\beta_0,\hat\beta)_{\lambda_1,\lambda_2}\big), \\
\mathrm{BIC}(\lambda_1,\lambda_2) &= 2 \cdot
\mathrm{loss}\big((\hat\beta_0,\hat\beta)_{\lambda_1,\lambda_2}\big) \,+\,
\log{N_\mathrm{tot}} \cdot
\mathrm{df}\big((\hat\beta_0,\hat\beta)_{\lambda_1,\lambda_2}\big),
\end{align*}
and in each case, the tuning parameter pair is chosen (again, say,
over a grid of possible values) to minimize the score.
In the above, \smash{$(\hat\beta_0,\hat\beta)_{\lambda_1,\lambda_2}$}
denotes the multinomial fused lasso estimate \eqref{eq::fusedmodel} at
the tuning parameter pair $\lambda_1,\lambda_2$, and $N_\mathrm{tot}$
denotes the total number of observations in the longitudinal study,
$N_\mathrm{tot}=nT$ (or \smash{$N_\mathrm{tot}=\sum_{t=1}^T n_t$} in
the missing data setting). Also,
\smash{$\mathrm{df}((\hat\beta_0,\hat\beta)_{\lambda_1,\lambda_2})$} denotes
the degrees of freedom of the estimate
\smash{$(\hat\beta_0,\hat\beta)_{\lambda_1,\lambda_2}$}, and we employ the
approximation
\begin{equation*}
\mathrm{df}\big((\hat\beta_0,\hat\beta)_{\lambda_1,\lambda_2}\big) \approx
\text{\# of nonzero blocks in $(\hat\beta_0,\hat\beta)_{\lambda_1,\lambda_2}$},
\end{equation*}
borrowing from known results in the Gaussian likelihood case
\citep{genlasso,lassodf2}. Finally,
\smash{$\mathrm{loss}((\hat\beta_0,\hat\beta)_{\lambda_1,\lambda_2})$}
denotes a loss function considered for the estimate, which we take
either to be the negative multinomial log likelihood
\smash{$-\ell((\hat\beta_0,\hat\beta)_{\lambda_1,\lambda_2})$}, as is
typical in AIC and BIC \citep{esl}, or the misclassification loss, to
put it on closer footing to cross-validation. Note that both loss
functions are computed in-sample, i.e., over the training samples, and
hence AIC and BIC are computationally much cheaper than
cross-validation.
We compare these model selection methods on the subset of the CHS-CS data set. The
individuals are randomly split into 5 folds. We use 4/5 of the data
set to perform model selection and subsequent model fitting with the 6
techniques described above: cross-validation, cross-validation with
the one-standard-error rule, and AIC and BIC under negative log
likelihood and misclassification losses. To be perfectly clear, the
model selection techniques work entirely within this given 4/5 of the
data set, so that, e.g., cross-validation further divides this data
set into folds. In fact, we used 4-fold cross-validation to make this
division simplest. The remaining 1/5 of the data set is then used for
evaluation of the estimates coming from each of the 6 methods, and
this entire process is repeated, leaving out each fold in turn as the
evaluation set. We record several measures on each evaluation set:
the misclassification rate, true positive rate in identifying the
dementia class, true positive positive rate in identifying the
dementia and death classes combined, and degrees of freedom (number of
nonzero blocks in the estimate). Figure \ref{fig::selection} displays
the mean and standard errors of these 4 measures, for each of the 6
model selection methods.
\begin{figure}[!ht]
\centering
\includegraphics[width=\textwidth]{figs/ModelSelection}
\caption{
Comparison of different methods for selection of tuning parameters
$\lambda_1,\lambda_2$ on the CHS-CS data set. The x-axis in each plot
parametrizes the 6 different methods considered, which are, from left
to right: AIC and BIC under negative log likelihood loss, AIC and BIC
under misclassification loss, cross-validation, and cross-validation
with the one-standard-error rule. The upper left plot shows
(out-of-sample) misclassification rate associated with the estimates
selected by each method, averaged over 5 iterations. The segments
denote $\pm 1$ standard errors around the mean. The red dotted line
is the average misclassification rate associated with the naive
estimator that predicts all individuals as dead (the majority class).
The upper right and bottom left plots show different measures of
evaluation (again, computed out-of-sample): the true positive rate in
identifying the dementia class, respectively, the true positive rate
in identifying the dementia and death classes combined. Finally, the
bottom right plot shows degrees of freedom (number of nonzero blocks)
of the estimates selected by each method.}
\label{fig::selection}
\end{figure}
Cross-validation and cross-validation with the
one-standard-error rule both seem to represent a favorable balance
between the different evaluation measures. The cross-validation
methods provide a misclassification rate significantly better than
that of the null model, which predicts according to the majority class
(death), they yield two of the three highest true positive rates in
identifying the dementia class, and perform well in terms of
identifying the dementa and death classes combined (as do all methods:
note that all true positive rates here are about 0.75 or higher).
We ended up settling cross-validation under the usual rule, rather
than the one-standard-error rule, because the
former achieves the highest true positive rate in identifying the
dementia class, which was our primary concern in the CHS-CS data
analysis. By design, cross-validation with the one-standard-error
rule delivers a simpler estimate in terms of degrees of freedom (196
for the one-standard-error rule versus 388 for the usual rule) though
both cross-validation models are highly regularized in absolute
terms (e.g., the fully saturated model would have thousands of nonzero
blocks).
\section{Discussion and future work}
\label{sec::discussion}
In this work, we proposed a multinomial model for
high-dimensional longitudinal classification tasks. Our proposal
operates under the assumption that a sparse number of predictors
contribute more or less persistent effects across time. The
multinomial model is fit under lasso and fused lasso regularization,
which address the assumptions of sparsity and persistence,
respectively, and lead to piecewise constant estimated coefficient
profiles. We described a highly efficient computational algorithm for
this model based on proximal gradient descent, demonstrated the
applicability of this model on an Alzheimer's data set taken from the
CHS-CS, and discussed practically important issues such stability
measures for the estimates and tuning parameter selection.
A number of extensions of the basic model are well within reach.
For example, placing a group lasso penalty on the coefficients
associated with each level of a binary expansion for a categorical
variable may be useful for encouraging sparsity in a group sense
(i.e., over all levels of a categorical variable at once). As another
example, more complex trends than piecewise constant ones may be fit
by replacing the fused lasso penalty with a trend filtering penalty
\citep{l1tf,trendfilter}, which would lead to piecewise polynomial
trends of any chosen order $k$. The appropriateness of such a penalty
would depend on the scientific application; the use of a fused lasso
penalty assumes that the effect of a given variable is mostly constant
across time, with possible change points; the use of a quadratic trend
filtering penalty (polynomial order $k=2$) allows the effect to vary
more smoothly across time.
More difficult and open-ended extensions concern statistical inference
for the fitted longitudinal classification models. For example, the
construction of confidence intervals (or bands) for selected
coefficients (or coefficient profiles) would be an extremely useful
tool for the practitioner, and would offer more concrete and
rigorous interpretations than the stability measures described in
Section \ref{sec::stability}. Unfortunately, this is quite a
difficult problem, even for simpler regularization schemes (such as
a pure lasso penalty) and simpler observation models (such as linear
regression). But recent inferential developments for related
high-dimensional estimation tasks
\citep{zhangconf,montahypo2,vdgsignif,LTTT2013,exactlasso,exactlars}
shed a positive light on this future endeavor.
\bibliographystyle{plainnat}
|
\section{Introduction}
\label{sec:intro}
Orientifolds of Type II superstring theories admit generalized fluxes via a successive application of T-duality on the three form $H$-flux. The same results in a chain of geometric and non-geometric fluxes as
\bea
\label{eq:Tdual}
& & H_{ijk} \longrightarrow \omega_{jk}{}^i \longrightarrow Q_k{}^{ij} \longrightarrow R^{ijk},
\eea
and have led to impetus progress in constructing string solutions in connection with the gauged supergravities in recent years \cite{Kachru:2002sk,Hellerman:2002ax,Dabholkar:2002sy,Hull:2004in,Derendinger:2004jn, Derendinger:2005ph, Shelton:2005cf, Dall'Agata:2009gv, Aldazabal:2011yz, Aldazabal:2011nj,Geissbuhler:2011mx,Grana:2012rr,Dibitetto:2012rk, Andriot:2012wx, Andriot:2012an, Andriot:2011uh,Blumenhagen:2013hva,Andriot:2013xca,Andriot:2014qla,Blair:2014zba}. Generically, all of such fluxes appear as parameters in the four dimensional effective potential and hence can develop a suitable scalar potential for the purpose of moduli stabilization which has been among the central aspects towards constructing realistic string models. For this goal, it is always preferred to have compactification backgrounds of much more rich structure and as much ingredients as possible because the same can induce new possibilities to facilitate the demands of (semi-)realistic model building. On these lines, the application of non-geometric fluxes towards moduli stabilization and cosmological model building aspects have attracted great amount of interest \cite{deCarlos:2009qm, Danielsson:2012by, Blaback:2013ht, Damian:2013dq, Damian:2013dwa, Hassler:2014mla} in recent time.
String fluxes are closely related to the possible gaugings in the gauged supergravity \cite{ Derendinger:2004jn, Derendinger:2005ph, Shelton:2005cf, Aldazabal:2006up, Dall'Agata:2009gv, Aldazabal:2011yz, Aldazabal:2011nj,Geissbuhler:2011mx,Grana:2012rr,Dibitetto:2012rk, Villadoro:2005cu} and it is remarkable that the four dimensional effective potentials could be studied (without having a full understanding of their ten-dimensional origin) via merely knowing the forms of K\"ahler and super-potentials \cite{Danielsson:2012by, Blaback:2013ht, Damian:2013dq, Damian:2013dwa, Blumenhagen:2013hva, Villadoro:2005cu, Robbins:2007yv, Ihl:2007ah}. Being simpler and well understood in nature, the Type II toroidal orinetifolds provide a promising toolkit to begin with while looking at new aspects, and so is the case with investigating the effects of non-geometric fluxes. Further, unlike the case with Calabi Yau compactifications, the explicit and analytic form of metric being known for the toroidal compactification backgrounds make such backgrounds automatically the favorable ones for performing explicit computations and studying the deeper insights of non-geometric aspects; for example the knowledge of metric has helped in knowing the ten-dimensional origin of the geometric flux dependent \cite{Villadoro:2005cu} as well as the non-geometric flux dependent potentials \cite{Blumenhagen:2013hva}.
In our previous work \cite{Blumenhagen:2013hva}, we have performed a close investigation
of the effects of T-duality motivated fluxes via considering their presence in the induced four-dimensional superpotential as proposed in \cite{Shelton:2005cf, Aldazabal:2011nj}.
We determined the most general form of the $H$, $F$, $Q$, $R$-fluxes in terms of the generalised metric and derived the Bianchi identities among these fluxes.
On a simple toroidal orientifold of type IIA and its T-dual type IIB model with all T-dual invariant geometric and non-geometric NS-NS
and R-R fluxes turned-on, we have computed the induced scalar potential from the four-dimensional superpotential and subsequently we have oxidized the various pieces into an underlying ten-dimensional supergravity action \cite{Blumenhagen:2013hva}.
We found that, both in the NS-NS and in the R-R sector, the resulting
oxidized ten-dimensional action is compatible with the flux formulation of the Double Field Theory action \cite{Aldazabal:2011nj,Geissbuhler:2011mx}.
The connection between a string compactification and the gauged supergravity effective theory mentioned so far is not the full story \cite{Aldazabal:2008zza,Font:2008vd,Guarino:2008ik} for both the type II superstring theories. In a setup of type IIB superstring theory compactified on ${\mathbb T}^6/{\left({\mathbb Z}_2 \times {\mathbb Z}_2\right)}$, it was argued that the additional fluxes are needed to ensure S-duality invariance of underlying low energy type IIB supergravity. The resulting modular completed fluxes can be arranged into spinor representations of $SL(2,{\mathbb Z})^7$, and can be described globally via a non-geometric compactification of F-theory when there is a geometric local description in terms of ten-dimensional supergravity \cite{Aldazabal:2008zza}. The Jacobi identities of the flux algebra then lead to the general form of the Bianchi identites in F-theory compactifications. The compactification manifold with $T$- and $S$-duality appears to be an $U$-fold \cite{Hull:2004in, Aldazabal:2008zza, Kumar:1996zx, Hull:2003kr} where local patches are glued
by performing $T$- and $S$-duality transformations.
As a result, a generalization of our previous work \cite{Blumenhagen:2013hva} to include the S-dual version of the non-geometric Q-flux, called P-flux, is necessary.
It is expected to provide a direct connection between the four-dimensional superpotential and the stringy aspects of the original T- and S-duality
invariant ten-dimensional supergravity.
Recently, axionic-inflation has received a lot of interest due to the
possible detection of primordial gravitational waves claimed by the BICEP2 collaboration \cite{Ade:2014xna}. The recent result of PLANCK \cite{Ade:2015lrj} implies that the value of tensor-to-scalar ratio $(r)$ around 0.2 (as claimed by BICEP2) can be explained by the foreground dust. Nevertheless, because of having an upper bound as $r< 0.11$, constructing models to realize non-trivially large values of $r$ are compatible as well as desired from the point of view of the possible future detection of gravitational waves. In the context of axion driven inflationary models developed in Type IIB/F-theory compactifciation, many proposals have emerged \cite{Hassler:2014mla,Grimm:2007hs,Blumenhagen:2014gta,Grimm:2014vva,Marchesano:2014mla,Arends:2014qca,Long:2014dta,Gao:2014uha,Ben-Dayan:2014lca,Garcia-Etxebarria:2014wla} in the recent times. In the original axion-monodromy inflation \cite{McAllister:2008hb, Flauger:2009ab}, the involutively odd $C_2$ axions have been proposed as being the inflaton candidate. The specific Calabi-Yau orientifolds which could support such odd axions along with their (F-term) moduli stabilization aspects have been studied in \cite{Blumenhagen:2008zz, Gao:2013pra,Grimm:2011dj,Gao:2013rra}. Regarding one of the recent axionic inflation models, a No-Go theorem has been proposed \cite{Blumenhagen:2014nba} as a challenge of creating a mass-hierarchy between universal axion and dilaton in type IIB orientifold compactifciation. The same has been of interest for constructing axion-monodromy inflationary model involving the universal axion \cite{Blumenhagen:2014gta,Gao:2014uha}. Equipped with the modular completed fluxes, we examine the original No-Go theorem. We find that despite of relatively much richer structure for universal-axion/dilaton dependences of the full potential, quite surprisingly, the No-Go statement still holds in a two field analysis, and thus showing its robustness.
This paper is organized as follows: in section \ref{sec:orientifold}, we present the basic set-up of the type IIB on $ {\mathbb T}^6 / \left(\mathbb Z_2\times \mathbb Z_2\right)$ orientifold and the general fluxes allowed to write out a generic form of superpotential involving the two NS-NS fluxes ($H, Q$) and their respective S-dual ($F, P$) fluxes. In section \ref{sec:Rearrangement}, a detailed study of the four-dimensional scalar potential induced by the flux superpotential is performed which takes us to propose a dimensional oxidation into the underlying ten-dimensional action in section {\ref{sec:S-dual}. The form of Chern-Simons terms reproducing the respective 3-brane and 7-brane tadpoles are also consistently invoked while considering the $SL(2,\mathbb{Z})$ invariance. Next, as an application to the potential we derived, in section \ref{sec:no-go}, we examine the role of non-geometric fluxes, specially S-dual P-fluxes, which they could play in the context of the No-go theorem mentioned in \cite{Blumenhagen:2014nba}. In the end, we summarize the results followed by two short appendices (\ref{appendix_0}) and (\ref{sec:D-term}) detailing some intermediate steps and the strategy followed for invoking the flux combinations needed for oxidation purpose.
\section{Type IIB on $ {\mathbb T}^6 / \left(\mathbb Z_2\times \mathbb Z_2\right)$ orientifold and fluxes}
\label{sec:orientifold}
Following the notations of \cite{Blumenhagen:2013hva}, let us briefly revisit the relevant features of a setup within type IIB superstring theory compactified on $ {\mathbb T}^6 / \left(\mathbb Z_2\times \mathbb Z_2\right)$ with the
two $\mathbb Z_2$ actions being defined as
\bea
\label{thetaactions}
& & \theta:(z^1,z^2,z^3)\to (-z^1,-z^2,z^3)\\
& & \overline\theta:(z^1,z^2,z^3)\to (z^1,-z^2,-z^3)\, . \nonumber
\eea
Further, the orientifold action is: ${\cal O} \equiv \Omega\, I_6 \, (-1)^{F_L}$ where $\Omega$ is the worldsheet parity, $F_L$ is left-fermion number while the holomorphic involution $I_6$ being defined as
\bea
\label{eq:orientifold}
& & I_6 : (z^1,z^2,z^3)\rightarrow (-z^1,-z^2,-z^3)\,,
\eea
resulting in a setup with the presence of $O3/O7$-plane. The complex coordinates $z_i$'s on $T^6=T^2\times T^2\times T^2$ are defined as
\bea
z^1=x^1+ i U_1 x^2, ~ z^2=x^3+ i U_2 x^4,~ z^3=x^5+ i U_3 x^6 ,
\eea
where the three complex structure moduli $U_i$'s can be written as
$U_i= u_i + i\, v_i,\,\,i=1,2,3$.
Now choosing the following basis of closed three-forms
\bea
\label{formbasis}
\alpha_0&=dx^1\wedge dx^3\wedge dx^5\,, ~
\beta^0=dx^2\wedge dx^4\wedge dx^6\, ,\nonumber\\
\alpha_1&=dx^1\wedge dx^4\wedge dx^6\, , ~
\beta^1=dx^2\wedge dx^3\wedge dx^5\, ,\\
\alpha_2&=dx^2\wedge dx^3\wedge dx^6\, , ~
\beta^2=dx^1\wedge dx^4\wedge dx^5\, ,\nonumber\\
\alpha_3&=dx^2\wedge dx^4\wedge dx^5\, ,~
\beta^3=dx^1\wedge dx^3\wedge dx^6\, \nonumber
\eea
satisfying $\int \alpha_I\wedge \beta^J=-\delta_I{}^J$, the
holomorphic three-form $\Omega_3=dz^1\wedge dz^2\wedge dz^3$ can be expanded as
\bea
& & \Omega_3= \alpha_0 + i \, (U_1 \beta^1 + U_2 \beta^2 + U_3 \beta^3)-i U_1
U_2 U_3 \beta^0 \nonumber\\
& & \hskip1.5cm -U_2 U_3 \alpha_1- U_1 U_3 \alpha_2 - U_1 U_2 \alpha_3\, .
\eea
The additional chiral variable are axion-dilaton
\bea
& & S= \, e^{-\phi}-i \, C_{0}\, .
\eea
and the K\"ahler moduli, generically being encoded in the complexified four-cycle volumes given as
\bea
J^c= \frac{1}{2} e^{-\phi} J \wedge J +i \, C^{(4)}\; .
\eea
In our case, these moduli are
\bea
T_1 = \tau_1+ i \, C^{(4)}_{3456}, ~ ~T_2 =\tau_2+ i \, C^{(4)}_{1256} , ~ ~T_3 = \tau_3+ i \, C^{(4)}_{1234},
\eea
where the real parts can be expressed in terms of the two-cycle volumes $t_i$
as,
$
\tau_1 = e^{-\phi} \, t_2 \, t_3, \,\,\tau_2 = e^{-\phi} \, t_3 \, t_1, \, \,
\,\,\tau_3 = e^{-\phi} \, t_1 \, t_2.
$
We also need to express the two-cycle volumes $t_i$ in terms of the
four-cycles volumes $\tau_i$ as,
\bea
\label{twoinfour}
& & t_1=\sqrt{\tau_2\, \tau_3\over \tau_1\, s}\,, ~~ t_2=\sqrt{\tau_1\, \tau_3\over \tau_2\, s}\,, ~~ t_3=\sqrt{\tau_1\, \tau_2\over \tau_3\, s}\,
\eea
with $s={\rm Re}(S)$.
Now, the non-vanishing components of the metric
in string frame are
\bea
\label{eq:gMN}
g_{MN}={\rm blockdiag}\Big({e^{\phi\over 2}\over {\sqrt{\tau_1\, \tau_2
\, \tau_3}}} \, \, \tilde g_{\mu\nu}, \, \, g_{ij}\Big) \, .
\eea
Further, the string frame internal metric $g_{ij}$
is also block-diagonal and
has the following non-vanishing components,
\bea
& & \hskip-0.8cm g_{11}=\frac{t_1}{u_1}\,, ~ g_{12}=-\frac{t_1 v_1}{u_1} =g_{21}\, , ~
g_{22}=\frac{t_1(u_1^2+v_1^2)}{u_1}\, ,\nonumber\\
& & \hskip-0.8cm g_{33}=\frac{t_2}{u_2}\, ,~ g_{34}=-\frac{t_2 v_2}{u_2}=g_{43}\, , ~
g_{44}=\frac{t_2(u_2^2+v_2^2)}{u_2}\, ,\\
& & \hskip-0.8cm g_{55}=\frac{t_3}{u_3}\,, ~ g_{56}=-\frac{t_3 v_3}{u_3}=g_{65}\,, ~
g_{66}=\frac{t_3(u_3^2+v_3^2)}{u_3}\, .\nonumber
\eea
These internal metric components can be written out in more a suitable form, to be utilized later, by using the four cycle volumes $\tau_i$'s and the same is given as under,
\bea
& & \hskip-0.8cm g_{11}=\frac{\sqrt{\tau_2} \, \sqrt{\tau_3}}{\sqrt{s} \, u_1 \, \sqrt{\tau_1}}\,, ~ g_{12}=-\frac{v_1 \, \sqrt{\tau_2} \, \sqrt{\tau_3}}{\sqrt{s} \, u_1 \, \sqrt{\tau_1}} =g_{21}\, , ~
g_{22}= \frac{\left(u_1^2 + v_1^2\right)\sqrt{\tau_2} \, \sqrt{\tau_3}}{\sqrt{s} \, u_1 \, \sqrt{\tau_1}} ,\nonumber\\
& & \hskip-0.8cm g_{33}=\frac{\sqrt{\tau_1} \, \sqrt{\tau_3}}{\sqrt{s} \, u_2 \, \sqrt{\tau_2}}\,, ~ g_{34}=-\frac{v_2 \, \sqrt{\tau_1} \, \sqrt{\tau_3}}{\sqrt{s} \, u_2 \, \sqrt{\tau_2}} =g_{43}\, , ~
g_{44}= \frac{\left(u_2^2 + v_2^2\right)\sqrt{\tau_1} \, \sqrt{\tau_3}}{\sqrt{s} \, u_2 \, \sqrt{\tau_2}} ,\\
& & \hskip-0.8cm g_{55}=\frac{\sqrt{\tau_1} \, \sqrt{\tau_2}}{\sqrt{s} \, u_3 \, \sqrt{\tau_3}}\,, ~ g_{56}=-\frac{v_3 \, \sqrt{\tau_1} \, \sqrt{\tau_2}}{\sqrt{s} \, u_3 \, \sqrt{\tau_3}} =g_{65}\, , ~
g_{66}= \frac{\left(u_3^2 + v_3^2\right)\sqrt{\tau_1} \, \sqrt{\tau_2}}{\sqrt{s} \, u_3 \, \sqrt{\tau_3}} ,\nonumber
\eea
Since the background fluxes $\omega^i{}_{jk}$ and $R^{ijk}$ are odd under
the orientifold projection, the only invariant fluxes
are the following components of the three-forms $H_3$ and $F_{3}$ as under,
\bea
& & \overline{H}:\quad {\overline H}_{135}\,,\ {\overline H}_{146}\, ,\ {\overline H}_{236}\, ,\ {\overline H}_{245}, {\overline H}_{246}\,, \ {\overline H}_{235}\, ,\ {\overline H}_{145}\, ,\ {\overline H}_{136}\, ,\\
& & \overline{F}:\quad {\overline F}_{135}\,,\ {\overline F}_{146}\, ,\ {\overline F}_{236}\, ,\ {\overline F}_{245}\, , {\overline F}_{246}\,, \ {\overline F}_{235}\, ,\ {\overline F}_{145}\, ,\ {\overline F}_{136} \nonumber
\eea
and the components of non-geometric $Q$ and P-fluxes, which can be collectively given as $\overline A\equiv Q$ or $P$:
\bea
& & \hskip-0.5cm \overline A: \, \, \, \overline A_1{}^{35}\,,\ \overline A_2{}^{45}\,,\ \overline
A_1{}^{46}\,,\ \overline A_2{}^{36}\, , \, \overline A_5{}^{13}\,,\ \overline A_6{}^{23}\,,\ \\
& & \hskip-0.5cm\hskip0.9cm \overline A_5{}^{24}\,,\ \overline A_6{}^{14}, \, \overline A_3{}^{51}\,,\ \overline A_4{}^{61}\,,\ \overline
A_3{}^{62}\,,\ \overline A_4{}^{52}\, ,\nonumber\\
& & \hskip-0.5cm \hskip0.9cm\overline A_2{}^{35}\,,\ \overline A_5{}^{23}\,,\ \overline
A_3{}^{52}\,,\ \overline A_2{}^{46} , \, \overline A_4{}^{51}\,,\ \overline A_1{}^{45}\,,\nonumber\\
& & \hskip-0.5cm \hskip0.9cm \overline A_5{}^{14}\,,\ \overline A_4{}^{62}\, , \, \overline A_6{}^{13}\,,\ \overline A_3{}^{61}\,,\ \overline
A_1{}^{36}\,,\ \overline A_6{}^{24}~. \nonumber
\eea
Now, the complete form of flux induced superpotential is given as \cite{Aldazabal:2006up},
\bea
\label{typeIIBW}
& & \hskip-0.5cm W = \frac{1}{4} \int_X \left(\overline {F} - i \, S \, \overline{H}\right) \wedge \Omega_3 - \frac{i}{4}\int_{X} \left[\left(\overline{Q} \,- i \, S \, \overline{P}\right)\bullet
J^c\right]\wedge \Omega_3 ,
\eea
where the three-form of type $\overline{A}\bullet J^c={1\over 6} (\overline{A}\bullet J^c)_{ijk}\,
dx^i\wedge dx^j\wedge dx^k$ is defined as
\bea
(\overline{A}\bullet J^c)_{ijk}={3\over 2}\, \overline{A}_{[\underline{i}}{}^{mn\,}
J^c_{mn\underline{jk}]}\, \, \, \, \, \, \, \, \, {\rm for} \, \, \, A \in\{Q, P\}. \nonumber
\eea
Together with the K\"ahler potential
\bea
\label{typeIIBK}
& & \hskip-0.5cm K = -\ln\left( \frac{S +\overline{S}}{2}\right) -\sum_{i=1,2,3}\ln\left( \frac{U_i +\overline{U_i}}{2}\right)- \, \sum_{i=1,2,3}\ln\left( \frac{T_i +\overline{T_i}}{2}\right)
\eea
it allows now to compute the F-term contribution of the effective four-dimensional scalar potential by utilizing the following standard relation
\bea
\label{VF}
V_{F}=e^{K}\Big(K^{i\bar\jmath}D_i W\, D_{\bar\jmath} \overline W-3
|W|^2\Big)\,.
\eea
As it is well reflected from the superpotential, the inclusion of dual P-flux provides a modular completion under the $SL(2, \mathbb{Z})$ transformation \cite{Aldazabal:2006up}:
\bea
\label{eq:SL2Z}
& & S\to \frac{kS-i\ell}{imS+n}\ , \quad kn-\ell m = 1\ , \quad k,\
\ell,\ m,\ n\in \mathbb{Z}\ ; \nonumber\\
& & \begin{pmatrix}F_3\\ H_3\end{pmatrix}\to\begin{pmatrix}k&\ell\\
m&n\end{pmatrix}\begin{pmatrix}F_3\\ H_3\end{pmatrix}, \, \, \, \, \, \begin{pmatrix}Q\\ P\end{pmatrix}\to\begin{pmatrix}k&\ell\\
m&n\end{pmatrix}\begin{pmatrix}Q\\ P\end{pmatrix}\ .
\eea
Let us mention that for our example there are no two-forms
anti-invariant under the orientifold projection
so that no $B_2$ and $C_2$ moduli are present.
The $SL(2, \mathbb{Z})$ self-dual action for IIB will result in further transformation on the rest of the massless bosonic spectrum. To explicitly check the modular invariance of the four-dimensional scalar potential, we consider a simplified version of S-duality transformation given as $S \to {1}/{S}$, under which chiral variables, together with R-R field and various fluxes transform in the following manner,
\bea
\label{eq:S-duality}
& & S \to \frac{1}{S}, \, \, T_{\alpha} \to {T_{\alpha}}, \, \, U_m \to {U_m}, \\
& & \hskip-0.0cm {H}_{ijk} \to {F}_{ijk}, \, \, {F}_{ijk} \to -{H}_{ijk}, {Q}^{ij}_{k} \to - {P}^{ij}_{k}, \, \, {P}^{ij}_{k}\to {Q}^{ij}_{k}. \nonumber
\eea
Here it should be noted that the Einstein-frame chiral coordinate $T_{\alpha}$ is invariant only in an orientifold with no odd axions, i.e. $h^{11}_-(X_6/{\cal O}) = 0$ \cite{Grimm:2007xm}. Under this S-duality, the superpotential and the K\"ahler potential (at the tree level) transform as:
\bea
\label{eq:S-duality1}
e^K \longrightarrow |S|^2 \, e^K\, , \, \, \, W \longrightarrow -\frac{i}{S} \, W.
\eea
which finally results in a S-duality invariant F-term potential $V_F$.
\section{Rearrangement of F-term scalar potential}
\label{sec:Rearrangement}
In this section we will present the full F-term scalar potential in the form of various ``suitable" pieces to be later utilized for the oxidation purpose in the next section.
Using the expressions of K\"ahler potential and superpotential given in eqns. (\ref{typeIIBW}-\ref{typeIIBK}), the full F-term scalar potential results in 9661 terms appearing in the form of quadratic-terms in four $H, F, Q$ and $P$-fluxes. Now, let us consider the following new flux-combinations which we have invoked after a very tedious terms-by-term investigation of the scalar potential,
\begin{subequations}
\label{eq:fluxOrbits}
\bea
\label{eq:fluxOrbits1}
&& {\cal H}_{ijk} = {h}_{ijk}~,\quad\quad {\cal Q}^{ij}_{k} = {Q}^{ij}_{k} - C_0 \, ~{P}^{ij}_{k} ~, \\
&& {\cal F}_{ijk}= {f}_{ijk} - C_0 \, ~{h}_{ijk}~, \quad\quad {\cal P}^{ij}_{k} = {P}^{ij}_{k}~ ,\nonumber
\eea
where
\bea
\label{eq:fluxOrbits2}
& & {h}_{ijk} = \left({H}_{ijk} +\frac{3}{2}\, {P}_{[\underline{i}}{}^{lm} C^{(4)}_{lm\underline{jk}]}\right)~, \, \, \, \, \, {f}_{ijk}=\left({F}_{ijk} +\frac{3}{2}\, {Q}_{[\underline{i}}{}^{lm} C^{(4)}_{lm\underline{jk}]}\,\right).
\eea
\end{subequations}
The importance/relevance of these flux combinations will be clearer as we proceed across the various sections of this article. By using these new flux orbits which generalizes the results of \cite{Blumenhagen:2013hva, Font:2008vd}, one can rewrite the old-flux squared terms (like $H^2, F^2$ etc.) into a set of new-flux squared terms (like ${\cal H}^2, {\cal F}^2$ etc.) in a useful manner. A close inspection of the full F-term scalar potential make it possible to rearrange the various terms into the following interesting pieces,
\begin{subequations}
\bea
\label{eq:potentialSdualEin}
& & \hskip-1.5cm {\bf V_{F}} = {\bf V_{{\cal H}{\cal H}}}+ {\bf V_{{\cal F}{\cal F}}} + {\bf V_{{\cal Q}{\cal Q}}} + {\bf V_{{\cal P}{\cal P}}} + {\bf V_{{\cal H}{\cal Q}}} + {\bf V_{{\cal F}{\cal P}}}+ {\bf V_{{\cal Q}{\cal P}}} \\
& & \hskip+1.5cm + {\bf V_{{\cal H}{\cal F}}}+ {\bf V_{{\cal F}{\cal Q}}}+ {\bf V_{{\cal H}{\cal P}}} + ....... \nonumber
\eea
where dots denote a collection of terms which could not be rearranged in new flux combinations, however such terms are precisely canceled by using the Bianchi identities which we will elaborate on later. The explicit expressions of various pieces in eqn. (\ref{eq:potentialSdualEin}) are given as under,
\bea
\label{eq:detailedV}
& & \hskip-2.5cm {\bf V_{{\cal H}{\cal H}}}= \frac{s}{4\, {\cal V}_E} \, \biggl[\frac{1}{3!}\overline{\cal H}_{ijk}\, \overline{\cal H}_{i'j'k'}\, g_E^{ii'}\, g_E^{jj'} g_E^{kk'}\biggr]\, \nonumber\\
& & \hskip-2.5cm {\bf V_{{\cal F}{\cal F}}}= \frac{1}{4 \, s \, {\cal V}_E}\biggl[\frac{1}{3!}\overline{\cal F}_{ijk}\, \overline{\cal F}_{i'j'k'}\, g_E^{ii'} \, g_E^{jj'} g_E^{kk'}\biggr] \nonumber\\
& & \hskip-2.5cm {\bf V_{{\cal Q}{\cal Q}}}= \frac{1}{4\, s \, {\cal V}_E} \, \biggl[3 \times \left(\frac{1}{3!}\, \overline{\cal Q}_k{}^{ij}\, \overline{\cal Q}_{k'}{}^{i'j'}\,g^E_{ii'} g^E_{jj'} g_E^{kk'} \right) + \, 2 \times \left(\frac{1}{2!}\overline{\cal Q}_m{}^{ni}\, \overline{\cal Q}_{n}{}^{mi'}\, g^E_{ii'}\right) \biggr]\nonumber\\
& & \hskip-2.5cm {\bf V_{{\cal P}{\cal P}}}= \frac{s}{4\, {\cal V}_E} \, \biggl[3 \times \left(\frac{1}{3!}\, \overline{\cal P}_k{}^{ij}\, \overline{\cal P}_{k'}{}^{i'j'}\,g^E_{ii'} g^E_{jj'} g_E^{kk'} \right) + \, 2 \times \left(\frac{1}{2!}\overline{\cal P}_m{}^{ni}\, \overline{\cal P}_{n}{}^{mi'}\, g^E_{ii'}\right) \biggr]\nonumber\\
& & \hskip-2.5cm {\bf V_{{\cal H}{\cal Q}}}= \frac{1}{4\, {\cal V}_E} \, \biggl[{\bf (-2)} \times \left(\frac{1}{2!} \overline{\cal H}_{mni} \, \overline{\cal Q}_{i'}{}^{mn}\, g_E^{ii'}\right)\biggr] \\
& & \hskip-2.5cm {\bf V_{{\cal F}{\cal P}}}= \frac{1}{4 \, {\cal V}_E} \, \biggl[{\bf (+2)} \times \left(\frac{1}{2!} \overline{\cal F}_{mni} \, \overline{\cal P}_{i'}{}^{mn}\, g_E^{ii'}\right)\biggr]\nonumber\\
& & \hskip-2.5cm {\bf V_{{\cal Q}{\cal P}}}= \frac{1}{4 \, {\cal V}_E}\, \biggl[{\bf (+2)} \times \left(\frac{1}{3!} \, \, (3\, \overline{\cal P}_{n}{}^{l'm'} \, g^E_{l' l} \, g^E_{m' m})\right)\,\, {\cal E}_E^{ijklmn} \, \, \left(\frac{1}{3!} \, (3 \, \overline{\cal Q}_{k}{}^{i'j'}\, g^E_{i'i} g^E_{j' j}) \, \right) \biggr] \nonumber\\
& & \hskip-1.5cm \equiv \frac{1}{4\, {\cal V}_E}\, \biggl[{\bf (+2)} \times \left(\frac{1}{2!} \, \, (\overline{\cal P}_{k'}{}^{i j} \, g_E^{k' k})\right)\, \, {\cal E}^E_{ijklmn} \, \, \left(\frac{1}{2!} (\overline{\cal Q}_{n'}{}^{l,m}\, g_E^{n'n}) \, \right) \biggr] \nonumber\\
& & \hskip-2.5cm {\bf V_{{\cal H}{\cal F}}}= \frac{1}{4 \, {\cal V}_E} \biggl[{\bf (+2)} \times \left(\frac{1}{3!} \, \times\, \frac{1}{3!} \, \, \overline{\cal H}_{ijk} \,\, {\cal E}_E^{ijklmn} \, \, \overline{\cal F}_{lmn}\right) \biggr] \nonumber\\
& & \hskip-2.5cm {\bf V_{{\cal F}{\cal Q}}}= \frac{1}{4\, s \, {\cal V}_E} \, \, \Big[{\bf (+2)} \times \biggl( \frac{1}{2!} \, \times\, \frac{1}{2!} \, \overline {\cal Q}_{i}{}^{j'k'} \, \overline {\cal F}_{j'k'j} \, \, \, \, \tau^E_{klmn} \,\,\, {\cal E}_E^{ijklmn} \, \, \biggr)\Big] \nonumber\\
& & \hskip-2.5cm {\bf V_{{\cal H}{\cal P}}}= \frac{s}{4 \, {\cal V}_E} \, \, \Big[{\bf (+2)} \times \biggl( \frac{1}{2!} \, \times\, \frac{1}{2!} \, \overline {\cal P}_{i}{}^{j'k'} \, \overline {\cal H}_{j'k'j} \, \, \, \, \tau^E_{klmn} \,\,\, {\cal E}_E^{ijklmn} \, \, \biggr)\Big]. \nonumber
\eea
\end{subequations}
In order to understand and appreciate the nice structures within the aforementioned expressions, we need to supplement the followings,
\begin{itemize}
\item{ In the rearrangement process, we have utilized some Einstein- and string-frame conversion relations given as ${\cal V}_E = s^{3/2}\, {\cal V}_s, \, g^E_{ij} = g_{ij} \, \sqrt{s}$ and $g_E^{ij} = g^{ij}/\sqrt{s}$ which helps us in seeing the S-duality invariance manifest.}
\item{The Levi-civita tensors are defined in terms of antisymmetric Levi-civita symbols $\epsilon_{ijklmn}$ and the same are given as: ${\cal E}^E_{ijklmn}=\sqrt{|g_{ij}|} \, \,\epsilon_{ijklmn} = {\cal V}_E \, \, \epsilon_{ijklmn}$ while ${\cal E}_E^{ijklmn}=\epsilon^{ijklmn}/\sqrt{|g_{ij}|} =\, \epsilon^{ijklmn}/{\cal V}_E $. Further, the Einstein- and string-frame Levi-civita tensors are related as: ${\cal E}_{ijklmn} = s^{-3/2} \, {\cal E}^E_{ijklmn}$ and ${\cal E}^{ijklmn} = s^{3/2} \, \, {\cal E}_E^{ijklmn}$.}
\item{$\tau^E_{klmn}$ denotes the four-form components corresponding to the saxionic counterpart of $C^{(4)}$ RR axions with the only non-zero components being the four-cycle volume moduli, which are given as $\tau^E_{3456} = \tau_1, \tau^E_{1256} = \tau_2, \tau^E_{1234} = \tau_3$ in the notations developed in the earlier section (\ref{sec:orientifold}).}
\end{itemize}
Another motivation for the collection of terms in eqn. (\ref{eq:detailedV}) being written out only in terms of Einstein frame quantities is the fact that in our later analysis of investigating a No-Go about the universal-axion/dilaton mass splitting, we want all (inverse-)metric appearance to be independent of the dilaton. The more on this aspect will be clear in section \ref{sec:no-go}. Further, to reflect the involvement and difficulties while invoking the right combinations of flux-orbits as well as the scalar potential rearrangement, it is important to mention the following counting of terms in various pieces of the rearrangement given in eqs.(\ref{eq:potentialSdualEin})-(\ref{eq:detailedV}),
\bea
& & \hskip-1.5cm \# ({\bf V_{{\cal H}{\cal H}}}) = 1054, \, \, \, \# ({\bf V_{{\cal F}{\cal F}}}) = 4108, \, \, \, \# ({\bf V_{{\cal Q}{\cal Q}}}) = 1071, \, \, \, \# ({\bf V_{{\cal P}{\cal P}}}) = 288,\nonumber\\
& & \# ({\bf V_{{\cal H}{\cal Q}}}) = 450,\, \, \, \# ({\bf V_{{\cal F}{\cal P}}}) = 450,\, \, \, \# ({\bf V_{{\cal Q}{\cal P}}}) = 324, \\
& & \# ({\bf V_{{\cal H}{\cal F}}}) = 128,\, \, \, \# ({\bf V_{{\cal F}{\cal Q}}}) = 288,\, \, \, \# ({\bf V_{{\cal H}{\cal P}}}) = 72 \nonumber
\eea
In addition, there are 1968 terms which are removed by using Bianchi identities and are denoted as dots in eqn. (\ref{eq:potentialSdualEin}). All these numbers sum up to a total of 9661 which is the number of terms in the F-term scalar potential. For a complete detail of term-by-term analysis by turning-on a subset of fluxes at a time, see appendix (\ref{appendix_0}).
Now, the following important observations can be made out of the eqns.(\ref{eq:potentialSdualEin})-(\ref{eq:detailedV}) along with the new-orbit arrangements as mentioned in eqs. (\ref{eq:fluxOrbits1})-(\ref{eq:fluxOrbits2}),
\begin{itemize}
\item{Not only the full potential (\ref{eq:potentialSdualEin})} is manifestly S-duality invariant, but also the internal pieces $({\bf V_{{\cal H}{\cal H}}}+{\bf V_{{\cal F}{\cal F}}})$, $({\bf V_{{\cal P}{\cal P}}}+{\bf V_{{\cal Q}{\cal Q}}})$, $({\bf V_{{\cal H}{\cal Q}}}+{\bf V_{{\cal F}{\cal P}}})$ and $({\bf V_{{\cal F}{\cal Q}}}+{\bf V_{{\cal H}{\cal P}}})$ form S-duality invariant combinations while the remaining two pieces ${\bf V_{{\cal Q}{\cal P}}}$ and ${\bf V_{{\cal H}{\cal F}}}$ are self-dual.
\item{In the absence of S-dual $P-$fluxes, one completely reproduces the results of \cite{Blumenhagen:2013hva}. Moreover, from eqns. (\ref{eq:fluxOrbits1}) and (\ref{eq:fluxOrbits2}), one can see that similar to the fact that inclusion of $Q$-fluxes corrects $F_3$-orbit by $(C_4 \bullet Q)$-type terms, the further inclusion of their dual P-fluxes modifies $H_3$-orbit with $(C_4 \bullet P$)-type terms.}
\item{As well expected, the S-dual completion results in a more symmetrical NS-NS and RR-sector flux orbits as one can see that similar to a RR-sector flux $F_3$ having a correction of type ${\cal F}_{ijk}= F_{ijk} - C_0 \, H_{ijk}$ in Taylor-Vafa construction (and as ${\cal F}_{ijk} = {f}_{ijk} - C_0 \,{h}_{ijk}$ in the current generalized version), now we have a NS-NS flux receiving a similar type of correction from a RR-flux in the form as ${\cal Q}_k{}^{ij} = Q_k{}^{ij} - C_0 \, P_k{}^{ij}$.}
\item{A relative minus sign in ${\bf V_{{\cal F}{\cal P}}}$ terms is observed as compared to those of ${\bf V_{{\cal H}{\cal Q}}}$ terms and the same is because of the definition of S-duality is as given in eq. (\ref{eq:S-duality}).}
\item{Invoking the peculiar form of ${\bf V_{{\cal Q}{\cal P}}}$ is necessary as well as crucial for the oxidation process as it contains many terms of PQ-type in which all the six-indices are different and being so, such terms can neither be washed away by using PQ-type Bianchi identities nor by the anti-commutation constraints because all such respective constraints as given in eqn. (\ref{eq:BIs}) involve summation over one index. Further, the term ${\bf V_{{\cal Q}{\cal P}}}$ could be written in two equivalent ways due to the following identity,
\bea
& & \hskip-1cm \epsilon^{ijklmn} g_{i i'} g_{j j'} g_{k k'} g_{l l'} g_{m m'} g_{n n'} = |det(g_{ij})| \, \epsilon_{i'j'k'l'm'n'}
\eea}
\item{The last three terms, namely ${\bf V_{{\cal H}{\cal F}}}$, ${\bf V_{{\cal F}{\cal Q}}}$ and ${\bf V_{{\cal H}{\cal P}}}$, are topological in nature, and so can be anticipated to be related to the contributions coming from various local sources such as brane/orientifold-tadpoles as we will elaborate now.}
\end{itemize}
\subsubsection*{Details of contributions from brane- and orientifold-sources}
In order to have the total scalar potential, the F-term contributions have to be supplemented with the D-term contributions subject to certain constraints coming from Bianchi identities (\ref{eq:BIs}) as described in detail in the appendix \ref{appendix_0}. Now as seen from the form of K\"ahler- and super-potentials, the eqn. (\ref{eq:S-duality1}) ensures that the F-term contribution is invariant under S-duality. Therefore in order to have an overall S-duality invariance, the D-terms should also be invariant and the same demands using generalized flux orbits instead of the normal ones as we will see in a moment. Further, as the pieces ${\bf V_{{\cal H}{\cal F}}}$, ${\bf V_{{\cal F}{\cal Q}}}$ and ${\bf V_{{\cal H}{\cal P}}}$ do not involve the metric unlike the rest of the terms in (of eqn. (\ref{eq:detailedV})) and happens to be topological in nature. Moreover, the combination $({\bf V_{{\cal H}{\cal F}}}+{\bf V_{{\cal F}{\cal Q}}}+{\bf V_{{\cal H}{\cal P}}})$ is indeed S-duality invariant. Therefore, the same should be (a part of) the contributions to be compensated by imposing RR Bianchi identities or via adding the respective contribution from the various local sources. The well anticipated contributions needed from brane- and orientifold- sources to cancel the topological pieces of eqn. (\ref{eq:detailedV}) can be considered as,
\bea
\label{eq:D-term}
{\bf V_{D}} = - {\bf V_{{\cal H}{\cal F}}}- {\bf V_{{\cal F}{\cal Q}}} - {\bf V_{{\cal H}{\cal P}}},
\eea
Here, it should be noted that $\bf{V_D}$ which is defined in terms of generalized flux combinations has a structure which is more than mere flux contributions and also contain the standard brane/orientiofld contributions coming from various local sources. Generically speaking, this ${\bf V_{D}}$ contains pieces from $D3$-brane, $O3$-plane as well as from all the 7-branes $(D7, NS7_i, I7_i)$ as we will see in the next section where the motivation for this collection written in terms of generalized flux-combinations would be clearer for oxidation purpose. Further, it should be noted that ${\bf V_{D}}$ contributions have some pieces which can be nullified by using certain Bianchi identities given in eqn. (\ref{eq:BIs}). For example, the piece $(-{\bf V_{{\cal H}{\cal P}}})$ will have certain terms of $HP$- and $PP$-types, and the later ones can be entirely washed away by using some of the $PP$-type Bianchi identities. To be precise, out of 72 terms of $(-{\bf V_{{\cal H}{\cal P}}})$, 48 are washed away while 24 terms survive
Now let us verify that the contributions, given in eqn. (\ref{eq:D-term}) which we also needed to compensate the topological pieces of eqn. (\ref{eq:detailedV}), indeed contain the generalized versions of D3/D7 tadpole-terms given in \cite{Blumenhagen:2013hva} with the inclusion of P-fluxes. For example, subject to applying the non-trivial Bianchi identities (\ref{eq:BIs}), switching off the P-flux recovers the following D3-tadpole terms \cite{Taylor:1999ii,Blumenhagen:2003vr},
\bea
\label{eq:VD3a}
V_{D3}^{HF} = - 2 \times \frac{1}{4 \, {\cal V}_E^2} \, \Big[ 20\, \overline{H}_{[\underline{123}}
\overline F_{\underline{456}]} \Big] \in (- {\bf V_{{\cal H}{\cal F}}}),
\eea
in addition to the following D7-tadpole terms (of \cite{Blumenhagen:2013hva}) given as under,
\bea
\label{eq:VD7a}
& & \hskip-1.3cm V_{D7}^{QF} = -2 \times \frac{1}{4\, s \, {\cal V}_E^2} \, \,
\Big[\overline Q_{[\underline{1}}{}^{jk} \, \overline F_{jk\underline{2}]} \, \tau_1+ \overline
Q_{[\underline{3}}{}^{jk} \, \overline F_{jk\underline{4}]}\, \tau_2+ \overline
Q_{[\underline{5}}{}^{jk} \, \overline F_{jk\underline{6}]} \, \tau_3 \Big] \in (- {\bf V_{{\cal F}{\cal Q}}}).
\eea
Now let us apply the reverse logic to motivate that in order to have S-duality invariance in the D-term contributions ($V_{D3}^{HF} + V_{D7}^{QF}$) of \cite{Blumenhagen:2013hva}, the use of generalized flux orbits is quite natural and necessary. For this purpose, consider the D7-tadpole terms $V_{D7}^{QF}$ as given in eqn. (\ref{eq:VD7a}) and invoke the terms needed for modular completion under transformations in eqn. (\ref{eq:S-duality}). Now as $1/s\longrightarrow (C_0^2 + s^2)/s$ with a S-dual of $V_{D7}^{QF}$ being of $(H P)$-type, and having in mind that $V_{D3}^{HF}$ is self-dual, one would need (at least) the following piece for a modular completion of D-term contributions,
\bea
\label{eq:VD7b}
& & \hskip-1.3cm V_{D7}^{PH} = - 2 \times \frac{\left(C_0^2+s^2\right)}{4\, s \, {\cal V}_E^2} \, \,
\Big[\overline P_{[\underline{1}}{}^{jk} \, \overline H_{jk\underline{2}]} \, \tau_1+ \overline
P_{[\underline{3}}{}^{jk} \, \overline H_{jk\underline{4}]}\, \tau_2+ \overline
P_{[\underline{5}}{}^{jk} \, \overline H_{jk\underline{6}]} \, \tau_3 \Big]
\eea
One should note that the additional piece with $C_0^2/s$ coefficient gets naturally absorbed into $(-{\bf V_{{\cal F}{\cal Q}}})$ when generalized version of fluxes ${\cal F}, {\cal H}, {\cal Q}$ and ${\cal P}$ fluxes are considered.
Thus using generalized flux combinations rearranges the terms appropriately taking care of modular completion.
Another reason which indicates the need of our generalized flux orbits (\ref{eq:fluxOrbits1}-\ref{eq:fluxOrbits2}) essential is the fact that, the 128 terms of cross-piece $(-{\bf V_{{\cal H}{\cal F}}})$ is reduced to 32 terms and 96 terms are removed via (HQ-FP) and PP-type Binachi identities. In addition to $V_{D3}^{HF}$ which consists of 8 terms of HF-type as mentioned in eqn. (\ref{eq:VD3a}), it also results in 24 more terms of $(P^{ij}_{k} Q^{lm}_{n} \epsilon_{ijklmn})$-type which (being topological) are different from those sitting inside ${\bf V_{{\cal Q}{\cal P}}}$. Noting that neither of the QP-type Bianchi identities nor the additional anti-commutative relation in eqn. (\ref{eq:BIs}) correspond to such PQ-terms because such constraints have at least one index of QP-term being summer over,
one should find a way to accommodate such PQ-type terms in the full picture. Interestingly, considering the generalized flux-combinations automatically does it via $(-{\bf V_{{\cal H}{\cal F}}})$, and thus resulting in no need for supplementing such strange topological terms of ${\cal Q}{\cal P}$-type.
Although, there are some more interesting aspects based on S-duality transformation of eight-form RR potential $C^{(8)}$ appearing as a triplet of eight-forms being related to produce a D-term of $({H}{ Q} + { F}{ P})$-type, however we postpone this issue to the next section, where we will discuss all the (oxidized) ten dimensional aspects
Thus, finally following all these taxonomy of terms and taking care of contributions from the various local sources, we reach a nicely structured form of the full scalar potential given as,
\bea
\label{eq:Fullpotential}
& & \hskip-1.5cm {\bf V_{Full}} = {\bf V_{F}} + {\bf V_{D}} = {\bf V_{{\cal H}{\cal H}}}+ {\bf V_{{\cal F}{\cal F}}} + {\bf V_{{\cal Q}{\cal Q}}} + {\bf V_{{\cal P}{\cal P}}} + {\bf V_{{\cal H}{\cal Q}}} + {\bf V_{{\cal F}{\cal P}}}+ {\bf V_{{\cal Q}{\cal P}}}
\eea
Now with this much ingredient in hand we are in a position to conjecture a modular completed version of the dimensional oxidation proposed in \cite{Blumenhagen:2013hva}.
\section{S-dual non-geometric type-IIB action: Dimensional oxidation to 10D}
\label{sec:S-dual}
With the analysis done in the previous section, a close inspection of the resulting full scalar potential, ${\bf V_{Full}} = {\bf V_{F}} + {\bf V_{D}}$ obtained as a sum of F-terms and local source contributions, reveals that all those terms can be recovered (up to satisfying a set of Bianchi identities) via a dimensional reduction from a set of generalized kinetic terms in a ten-dimensional action which, in string frame, is given as,
\begin{subequations}
\bea
\label{eq:oxiaction}
& & \hskip-1.5cm S={1\over 2}\int d^{10} x\, \sqrt{-g} \, \, \Big( {\bf {\cal L}_{\cal H \cal H}}+{\bf {\cal L}_{\cal F \cal F}} +{\bf {\cal L}_{\cal Q \cal Q}}+{\bf {\cal L}_{\cal P \cal P}} + {\bf {\cal L}_{{\cal H}{\cal Q}}} + {\bf {\cal L}_{{\cal F}{\cal P}}} + + {\bf {\cal L}_{{\cal Q}{\cal P}}} \Big)
\eea
where
\bea
\label{eq:detailedV0}
& & {\bf {\cal L}_{\cal H \cal H}}= -{e^{-2\phi}\over 2} \, \biggl[\frac{1}{3!}\overline{\cal H}_{ijk}\, \overline{\cal H}_{i'j'k'}\, g^{ii'}\, g^{jj'} g^{kk'}\biggr]\, \nonumber\\
& & {\bf {\cal L}_{\cal F \cal F}}= -{{1}\over 2} \, \biggl[\frac{1}{3!}\overline{\cal F}_{ijk}\, \overline{\cal F}_{i'j'k'}\, g^{ii'} \, g^{jj'} g^{kk'}\biggr] \\
& & {\bf {\cal L}_{\cal Q \cal Q}}= -{e^{-2\phi}\over 2} \, \biggl[3 \times \left(\frac{1}{3!}\, \overline{\cal Q}_k{}^{ij}\, \overline{\cal Q}_{k'}{}^{i'j'}\,g_{ii'} g_{jj'} g^{kk'} \right)+ \, 2 \times \left(\frac{1}{2!}\overline{\cal Q}_m{}^{ni}\, \overline{\cal Q}_{n}{}^{mi'}\, g_{ii'}\right) \biggr]\nonumber\\
& & {\bf {\cal L}_{\cal P \cal P}}= -{e^{-4\phi}\over 2} \, \biggl[3 \times \left(\frac{1}{3!}\, \overline{\cal P}_k{}^{ij}\, \overline{\cal P}_{k'}{}^{i'j'}\,g_{ii'} g_{jj'} g^{kk'} \right)+ \, 2 \times \left(\frac{1}{2!}\overline{\cal P}_m{}^{ni}\, \overline{\cal P}_{n}{}^{mi'}\, g_{ii'}\right) \biggr]\nonumber\\
& & {\bf {\cal L}_{{\cal H}{\cal Q}}}= -{e^{-2\phi}\over 2} \, \biggl[{\bf (-2)} \times \left(\frac{1}{2!} \overline{\cal H}_{mni} \, \overline{\cal Q}_{i'}{}^{mn}\, g^{ii'}\right)\biggr]\nonumber\\
& & {\bf {\cal L}_{{\cal F}{\cal P}}}= -{e^{-2\phi}\over 2} \, \biggl[{\bf (+2)} \times \left(\frac{1}{2!} \overline{\cal F}_{mni} \, \overline{\cal P}_{i'}{}^{mn}\, g^{ii'}\right)\biggr]. \nonumber\\
& & {\bf {\cal L}_{{\cal Q}{\cal P}}}= -{e^{-3\phi}\over 2} \, \biggl[{\bf (+2)} \times \left(\frac{1}{2!} \, \, (\overline{\cal P}_{k'}{}^{i j} \, g^{k' k})\right)\, {\cal E}_{lmnijk} \, \left(\frac{1}{2!} \, \, (\overline{\cal Q}_{n'}{}^{l,m}\, g^{n'n}) \, \right)\biggr]. \nonumber
\eea
\end{subequations}
This modular completed oxidation generalizes the results of \cite{Blumenhagen:2013hva}. Here, the new flux-orbits are the same as defined earlier in eqns. (\ref{eq:fluxOrbits1}-\ref{eq:fluxOrbits2}) while the (inverse-)metric components are written in their respective string frame expressions using ${\cal V}_E = s^{3/2}\, {\cal V}_s, \, g^E_{ij} = g_{ij} \, \sqrt{s}$ and $g_E^{ij} = g^{ij}/\sqrt{s} \,$. Recall that string frame Levi-civita tensor is related to its Einstein frame expression as ${\cal E}_{lmnijk} = s^{-3/2} \, {\cal E}^E_{lmnijk}$. The presence of Levi-civita tensor in ${\bf {\cal L}_{{\cal Q}{\cal P}}}$ is quite anticipated for the invariance of the same as under S-duality one has $\{Q \rightarrow -P, P\rightarrow Q\}$.
Further, for capturing the correct coefficients of the respective flux-squared quantities such as $|{\cal H}|^2, |{\cal F}|^2$ etc. to those of previous section via dimensional reduction of the 10D action proposed, one has to use the ten-dimensional metric given in eqn. (\ref{eq:gMN}) as the following,
\bea
& & \hskip-1.0cm \int d^{10} x\, \sqrt{-g} \, (....) \simeq \int d^{4} x\, \sqrt{-g_{\mu\nu}} \left(\frac{1}{s^{4} \, {\cal V}_s^2}\right) \times \left(\int d^{6} x\, \sqrt{-g_{mn}}\right) \, \times (..........) ~~ \\
& & \hskip2.2cm \simeq \int d^{4} x\, \sqrt{-g_{\mu\nu}} \times \left(\frac{1}{s^{4} \, {\cal V}_s}\right) \times (..........). \nonumber
\eea
as $\int d^{6} x\, \sqrt{-g_{mn}} \equiv {\cal V}_s$ gives the string-frame 6D volume. As we can see now, the S-duality invariance in the oxidized ten-dimensional action written in string frame is not explicitly manifest as opposed to the analysis of previous section in which we kept the expressions in terms of Einstein frame quantities. In order to see the full S-duality invariance of the 10D action (\ref{eq:oxiaction}), one has to take care of transformation of the integral measure as well.
\subsubsection*{Comments on local-source contributions relating to CS-action in 10D}
After proposing the oxidized ten-dimensional kinetic terms, now let us also focus on the contributions ${\bf V_D}$, given in eqn. (\ref{eq:D-term}), which could be thought of being related to the ten-dimensional Chern-Simons terms of the following $SL(2,\mathbb{Z})$-invariant types \cite{Aldazabal:2006up, Aldazabal:2008zza, Font:2008vd, Guarino:2008ik}\footnote{Here, we have a sign difference in the first and last terms involving $C^{(4)}$ and $\tilde C'^{(8)}$, as compared to those in \cite{Aldazabal:2006up, Aldazabal:2008zza}. This is because of the presence of a relative minus sign in $C_4$ (and $C_0$ also) while defining the chiral variables $T$ (and $S$) as compared to their respective definitions in \cite{Aldazabal:2006up, Aldazabal:2008zza}.},
\bea
\label{eq:CS1}
& & \hskip-1.2cm S_{CS} \sim - \int C^{(4)} \wedge {F} \wedge {H} \\
& & - \int C^{(8)} \wedge {Q}\bullet {F}\, + \int \tilde{C}^{(8)} \wedge {P}\bullet {H} \, - \int C'^{(8)} \wedge ({Q} \bullet {H} + {P} \bullet {F}) \nonumber
\eea
The first line is related to $D3/O3$-tadpoles while the second line corresponds to various 7-brane tadpoles. The first term is manifestly S-duality invariant as the RR four-form $C^{(4)}$ is $SL(2,\mathbb{Z})$ invariant, while for checking the S-duality invariance in the second line terms, one needs to consider the fact that the eight-form RR potential appears as an $SL(2, \mathbb{Z})$ triplet ($C^{(8)}$ , $\tilde C^{(8)}$ , $C'^{(8)}$) of eight-forms. These eight-form triplet components follow the S-dual transformation as
\bea
\label{eq:C8transform}
& & C^{(8)} \to - \tilde C^{(8)},\,\, \tilde C^{(8)} \to - C^{(8)},\,\, C'^{(8)}\to - C'^{(8)}
\eea
The first two of these transformation relations ensure the S-duality invariance between the first two terms of the second line of eqn. (\ref{eq:CS1}) relating $D7$-brane and S-dual $NS7_i$-brane tadpoles \cite{Aldazabal:2006up}. Further, the sign change of $C'^{(8)}$ under S-duality ensure the survival of S-duality odd combination of fluxes $({Q} \bullet {H} + {P} \bullet {F})$ which results in the so-called $I_{7_i}$-brane tadpoles. Further as the eight-form potentials ($C^{(8)}$ , $\tilde C^{(8)}$ , $C'^{(8)}$) correspond to the dual of axion-dilaton $S$, therefore there should be some way to reduce the same into two propagating degrees of freedoms, and we will see it happening precisely while using our new flux-orbits.
Now, let us explicitly investigate the origin of our D-brane tadpoles given in eqn. (\ref{eq:D-term}) through the respective ten-dimensional Chern-Simons' action, and see how those could get related to eqn. (\ref{eq:CS1}). For this purpose, let us reconsider the expressions D-brane tadpoles being the following pieces written from eqn. (\ref{eq:D-term}) as under,
\bea
& & \hskip-1.0cm {\bf V_D} = V_1 + V_2; \\
& & V_{1} = -\frac{1}{2 \, {\cal V}_E} \biggl[ \left(\frac{1}{3!} \, \times\, \frac{1}{3!} \, \, \overline{\cal H}_{ijk} \,\, {\cal E}_E^{ijklmn} \, \, \overline{\cal F}_{lmn}\right) \biggr] \nonumber\\
& & V_{2} = - \frac{1}{2\, s \, {\cal V}_E} \, \, \Big[ \biggl( \frac{1}{2!} \, \times\, \frac{1}{2!} \, \overline {\cal Q}_{i}{}^{j'k'} \, \overline {\cal F}_{j'k'j} \, \, \, \, \tau^E_{klmn} \,\,\, {\cal E}_E^{ijklmn} \, \, \biggr)\Big] \nonumber\\
& & \hskip 1cm - \frac{s}{2 \, {\cal V}_E} \, \, \Big[\biggl( \frac{1}{2!} \, \times\, \frac{1}{2!} \, \overline {\cal P}_{i}{}^{j'k'} \, \overline {\cal H}_{j'k'j} \, \, \, \, \tau^E_{klmn} \,\,\, {\cal E}_E^{ijklmn} \, \, \biggr)\Big] \nonumber
\eea
At a first glance, it appears that terms of ${\bf V_D}$ can be trivially related to all the piece of CS action in eqn (\ref{eq:CS1}) except the last terms with a piece $({Q} \bullet {H} + {P} \bullet {F})$. However, one should recall that ${\bf V_D}$ is written out in terms of generalized flux combinations while CS-terms in eqn. (\ref{eq:CS1}) are written using normal fluxes. Let us make some more taxonomy of the respective terms. Writing back these expressions in terms of older fluxes by using our new-flux orbit definitions in eqn. (\ref{eq:fluxOrbits1}-\ref{eq:fluxOrbits2}), we find an interesting rearrangement of terms \footnote{For explicit details related to which of the terms are nullified by Bianchi identities, see full expressions of $V_1$ and $V_2$ in terms of non-generalized fluxes given in the appendix (\ref{sec:D-term}).},
\bea
& & \hskip-0.5cm V_{1} \equiv V_{1}^a + V_{1}^b \, ; \, \nonumber\\
& & V_1^a \, =\, - \, \frac{1}{2 \, {\cal V}_E} \biggl[ \left(\frac{1}{3!} \, \times\, \frac{1}{3!} \, \, \overline{H}_{ijk} \,\, {\cal E}_E^{ijklmn} \, \, \overline{F}_{lmn}\right) \biggr] \nonumber\\
& & \hskip0.0cm V_1^b = - \, \, \frac{1}{2 \, {\cal V}_E} \biggl[ \left(\frac{1}{3!} \, \times\, \frac{1}{3!} \, \, \left(\frac{3}{2}\, {P}_{[\underline{i}}{}^{l'm'} C^{(4)}_{l'm'\underline{jk}]}\right) \,\, {\cal E}_E^{ijklmn} \, \, \left(\frac{3}{2}\, {Q}_{[\underline{l}}{}^{l'm'} C^{(4)}_{l'm'\underline{mn}]}\right) \right) \biggr] \nonumber\\
& & \hskip-0.5cm V_{2} \equiv V_{2}^a + V_{2}^b +V_2^c \, ; \, \\
& & V_2^a = - \, \, \frac{1}{2\, s \, {\cal V}_E} \, \, \Big[ \biggl( \frac{1}{2!} \, \times\, \frac{1}{2!} \, \overline {Q}_{i}{}^{j'k'} \, \overline {F}_{j'k'j} \, \, \, \, \tau^E_{klmn} \,\,\, {\cal E}_E^{ijklmn} \, \, \biggr)\Big] \nonumber\\
& & \hskip0.0cm V_2^b = - \, \, \frac{s^2+C_0^2}{2 \, s\, {\cal V}_E} \, \, \Big[\biggl( \frac{1}{2!} \, \times\, \frac{1}{2!} \, \overline {P}_{i}{}^{j'k'} \, \overline {H}_{j'k'j} \, \, \, \, \tau^E_{klmn} \,\,\, {\cal E}_E^{ijklmn} \, \, \biggr)\Big] \nonumber\\
& & V_2^c = -\, \, \frac{1}{2 \,\, {\cal V}_E} \, \times \left(\frac{-C_0}{s}\right)\, \Big[\biggl( \frac{1}{2!} \, \times\, \frac{1}{2!} \, \left(\overline {Q}_{i}{}^{j'k'} \, \overline {H}_{j'k'j}+\overline {P}_{i}{}^{j'k'} \, \overline {F}_{j'k'j}\right) \, \, \, \, \tau^E_{klmn} \,\,\, {\cal E}_E^{ijklmn} \, \, \biggr)\Big] \nonumber
\eea
Now, we have a couple of peculiar and very interesting observations to make,
\begin{itemize}
\item{The term $V_1^a$ simply corresponds to the well-known D3-tadpoles in a setup without non-geometric fluxes, and
\bea
& & V_1^a \in - \int C^{(4)} \wedge {F} \wedge {H}
\eea. }
\item{Using $s\to s/(s^2+C_0^2)$ along with flux and eight-form transformations, it is clear that $V_2^a+V_2^b$ is S-duality invariant and
\bea
& & V_2^a+V_2^b \in - \int C^{(8)} \wedge {Q}\bullet {F}\, + \int \tilde{C}^{(8)} \wedge {P}\bullet {H}.
\eea}
\item{The term $V_2^c$ with anti- S-dual combination $({Q} \bullet {H} + {P} \bullet {F})$ survives because the coefficient is also anti- S-dual as $C_0/s \to -\, C_0/s$ under S-duality. Thus, we are able to recover the last term in CS-action as
\bea
& & V_2^c \in -\int C'^{(8)} \wedge ({Q} \bullet {H} + {P} \bullet {F}).
\eea}
\item{In addition to the four type of terms we discussed, if we use non-generalized flux orbits, there is an additional term in form of $V_1^b$. Note that, this piece contains some terms of PQ-types in which all six flux-indices are different, and so such terms can neither be nullified by using any PQ- Bianchi identities nor using any of the anti-commutation constraints of QP-type. Therefore one needs to either introduce a new CS-term of type
\bea
& & V_1^b \in - \int C^{(4)} \wedge {\tilde P} \wedge {\tilde Q}
\eea
where $\tilde{P}_{lmn} = \left(\frac{3}{2}\, {P}_{[\underline{l}}{}^{l'm'} C^{(4)}_{l'm'\underline{mn}]}\right)$ and $\tilde{Q}_{lmn} = \left(\frac{3}{2}\, {Q}_{[\underline{l}}{}^{l'm'} C^{(4)}_{l'm'\underline{mn}]}\right)$,
or else one should seek for another way of absorbing those terms into the standard picture by rearranging the field strengths. In our case, it is the later one which happens to be true via using new flux-combinations.}
\end{itemize}
As we have mentioned earlier, these new observations also support the need of using our generalized flux combinations. In the new flux orbits, we not only embed all terms of $ ({Q} \bullet {H} + {P} \bullet {F})$ coupled with $C'^{(8)}$ eight-form into terms of type ${{\cal F}{\cal Q}}$ and ${{\cal H}{\cal P}}$, but also this helps in absorbing the additional strange looking PQ-type terms into ${\cal H} \wedge {\cal F}$. One should note that using generic form of (QH-PF) Bianchi identity, which is given as $Q_{[\overline k}{}^{ij} H_{\overline l \overline m]j} - P_{[\overline k}{}^{ij} F_{\overline l \overline m]j} = 0$, will generically not allow the nullification of the respective terms of $({Q} \bullet {H} + {P} \bullet {F})$ though it can reduce the number of such terms \footnote{However, as pointed out in \cite{Aldazabal:2006up}, this combination $({\cal Q} \bullet {\cal H} + {\cal P} \bullet {\cal F})$ does not have RR character and, in particular cases,
this term can be nullified. For example, by using the following simplified version of Binachi identities does so,
\bea
{\cal Q}_{[\overline k}{}^{ij} H_{\overline l \overline m]j} = 0 = P_{[\overline k}{}^{ij} F_{\overline l \overline m]j}\,\,\,\,\,
\Longrightarrow \,\,\,\, Q \bullet H + P \bullet F =0 .
\eea
Unlike this simplified case, there are examples of flux choices giving non-zero $I_{7_i}$-brane tadpoles in \cite{Guarino:2008ik}.}. Subsequently, we propose the following generalized form of the ten-dimensional Chern-Simons' action written in terms of new flux-orbits as under,
\bea
\label{eq:CS}
& & \hskip-2cm S_{CS} \sim - \int C^{(4)} \wedge {\cal F} \wedge {\cal H} - \int C^{(8)} \wedge {\cal Q}\bullet {\cal F}\, + \int \tilde{C}^{(8)} \wedge {\cal P}\bullet {\cal H} \, \, . \, \,
\eea
\section{Robustness of the No-Go for universal-axion and dilaton mass splitting }
\label{sec:no-go}
The general four dimensional scalar potential with the inclusion of all four types of (non-)geometric fluxes ($H, F, Q$ and $P$), depend on all the 14 real moduli/axions, and a schematic form would be as under,
\bea
{\bf V} \equiv {\bf V}\left(s, c_0, \tau_i, \rho_i, u_i, v_i \right) \, \, \, \, \, \, \, \forall i \in \{1, 2, 3\}.
\eea
Here, the scalar potential ${\bf V}$, as mentioned in eqs. (\ref{eq:Fullpotential}), denotes the sum of F-and D-term contributions which, can also be obtained from the dimensional reduction of 10D action proposed in eqns. (\ref{eq:oxiaction})-(\ref{eq:detailedV0}) along with the generalized flux orbits as in eqns. (\ref{eq:fluxOrbits1})-(\ref{eq:fluxOrbits2}). After collecting all the terms for dependencies of universal axion $c_0$ and the dilaton $s$, the very general scalar potential takes the following form,
\bea
\label{eq:Vpotsimple}
V = \left(\frac{a_1}{s} + a_2 + a_3 \, s \right) + \frac{a_4}{s} \, c_0 + \frac{a_5}{s} \, c_0^2
\eea
where $a_i$'s are generically some functions of various fluxes and all moduli/axions except universal axion $c_0$ and the dilaton $s$. This form of rearrangement of terms has been made to facilitate the study of a two-field dynamics. The extremization conditions for $c_0$ and $s$ are simply given as
\bea
& & \frac{\partial V}{\partial c_0} = \frac{a_4+ 2 \, a_5 \, c_0}{s}\, , \, \, \, \frac{\partial V}{\partial s} = -\frac{a_1 + a_4 \, c_0 + a_5 \, c_0^2 - a_3 \, s^2}{s^2}\, . \,
\eea
This shows that if one wants $\frac{\partial V}{\partial c_0}=0$ without fixing $c_0$, then one needs to satisfy flux constraints $a_4 = 0 = a_5$, and subsequently $\frac{\partial V}{\partial s} = -\frac{a_1 - a_3 \, s^2}{s^2}$. Now, the most crucial thing which happens to be true, is the fact that
\bea
a_3 = a_5
\eea
and the same implies that {\it ``the dilaton $s$ can not be fixed via $\frac{\partial V}{\partial s} = 0$ unless the universal axion $c_0$ is fixed via $\frac{\partial V}{\partial c_0} = 0$"}. Note that all the $a_i$-parameters generically depend on all the other moduli/axions except the universal axion and dilaton, nevertheless the above quoted argument holds independent of the fact whether those additional moduli or axions are stabilized or not. This is because of the fact that this argument is independent of the details of $a_i$s and follows from the extremization conditions of $c_0$ and dilaton. Moreover, it is worth to note that the condition: $a_3 = a_5$, holds irrespective of imposing the Bianchi identities or adding counter tadpole -terms. Now to support our arguments, we compare the scaler potential given in eqs. (\ref{eq:potentialSdualEin})-(\ref{eq:detailedV}) with our eqn. (\ref{eq:Vpotsimple}), and we get the following explicit expressions of $a_i$'s,
\bea
& & \hskip-0.5cm a_1 = \frac{1}{4\, {\cal V}_E} \times \frac{1}{3!}\, \left(\overline{F}_{ijk}+\frac{3}{2}\,\overline{Q}_{[\underline{i}}{}^{lm} C^{(4)}_{lm\underline{jk}]}\right) \left(\overline{F}_{i'j'k'} +\frac{3}{2}\,\overline{Q}_{[\underline{i'}}{}^{l'm'} C^{(4)}_{l'm'\underline{j'k'}]}\,\right) g_E^{ii'} \, g_E^{jj'} g_E^{kk'} \nonumber\\
& & \hskip-0.5cm \hskip0.7cm + \frac{1}{4\, {\cal V}_E} \, \biggl[3 \times \left(\frac{1}{3!}\, \overline{ Q}_k{}^{ij}\, \overline{ Q}_{k'}{}^{i'j'}\,g^E_{ii'} g^E_{jj'} g_E^{kk'} \right)+ \, 2 \times \left(\frac{1}{2!}\overline{ Q}_m{}^{ni}\, \overline{ Q}_{n}{}^{mi'}\, g^E_{ii'}\right) \biggr]\, ,\nonumber\\
& & \hskip-0.5cm a_2 = \frac{1}{4 \, {\cal V}_E} \, \biggl[\, 2 \times \left(\frac{1}{2!} \overline{F}_{mni} \, \overline{P}_{i'}{}^{mn}\, g_E^{ii'}\right)-2 \times \left(\frac{1}{2!} \overline{H}_{mni} \, \overline{Q}_{i'}{}^{mn}\, g_E^{ii'}\right) \nonumber\\
& & \hskip0.5cm + 2 \times \left(\frac{1}{2!} \, \, (\overline{P}_{k'}{}^{i j} \, g_E^{k' k})\right)\, {\cal E}^E_{lmnijk} \, \left(\frac{1}{2!} \, \, (\overline{Q}_{n'}{}^{l,m}\, g_E^{n'n}) \, \right)\biggr] \, , \nonumber\\
& & \hskip-0.5cm a_3 = \frac{1}{4\, {\cal V}_E} \times \frac{1}{3!}\, \left(\overline{H}_{ijk}+\frac{3}{2}\,\overline{P}_{[\underline{i}}{}^{lm} C^{(4)}_{lm\underline{jk}]}\right) \left(\overline{H}_{i'j'k'} +\frac{3}{2}\,\overline{P}_{[\underline{i'}}{}^{l'm'} C^{(4)}_{l'm'\underline{j'k'}]}\,\right) g_E^{ii'} \, g_E^{jj'} g_E^{kk'} \nonumber\\
& & \hskip-0.5cm \hskip0.7cm + \frac{1}{4\, {\cal V}_E} \, \biggl[3 \times \left(\frac{1}{3!}\, \overline{ P}_k{}^{ij}\, \overline{ P}_{k'}{}^{i'j'}\,g^E_{ii'} g^E_{jj'} g_E^{kk'} \right)+ \, 2 \times \left(\frac{1}{2!}\overline{ P}_m{}^{ni}\, \overline{ P}_{n}{}^{mi'}\, g^E_{ii'}\right) \biggr] \nonumber\\
& & \nonumber\\
& & \equiv a_5 \, ,\\
&& \nonumber\\
& & \hskip-0.5cm a_4 = \frac{(- 2)}{4\, {\cal V}_E} \times \frac{1}{3!}\, \left(\overline{F}_{ijk}+\frac{3}{2}\,\overline{Q}_{[\underline{i}}{}^{lm} C^{(4)}_{lm\underline{jk}]}\right) \left(\overline{H}_{i'j'k'} +\frac{3}{2}\,\overline{P}_{[\underline{i'}}{}^{l'm'} C^{(4)}_{l'm'\underline{j'k'}]}\,\right) g_E^{ii'} \, g_E^{jj'} g_E^{kk'}\nonumber\\
& & \hskip-0.5cm \hskip0.7cm+ \frac{(- 2)}{4\, {\cal V}_E} \, \biggl[3 \times \left(\frac{1}{3!}\, \overline{P}_k{}^{ij}\, \overline{Q}_{k'}{}^{i'j'}\,g^E_{ii'} g^E_{jj'} g_E^{kk'} \right) + \, 2 \times \left(\frac{1}{2!}\overline{P}_m{}^{ni}\, \overline{Q}_{n}{}^{mi'}\, g^E_{ii'}\right) \biggr] ~. \nonumber
\eea
Let us point out that by looking at the S-duality transformation, we observe that:
\bea
& & a_1 \leftrightarrow a_3 \equiv a_5, a_2 \rightarrow a_2, a_4 \rightarrow -a_4.
\eea
which using $s\longrightarrow s/(c_0^2 + s^2)$ and $c_0/s \longrightarrow - c_0/s$ ensure the S-duality invariance of the total potential in $\{c_0, s\}$ variables. The full potential can be written in S-dual pieces after using $a_3 = a_5$ as below,
\bea
\label{eq:Vpotsimple1}
V = a_2 + \left(\frac{a_1}{s} + {a_3} \, \frac{(c_0^2 + s^2)}{s} \right) + {a_4} \, \frac{c_0}{s}
\eea
The next question is whether it is possible to create a hierarchy via additional fluxes when we stabilize $c_0$ and $s$ simultaneously. To address this question will need a complete minimization analysis of the full scalar potential with 14 scalars along with an overall 64 flux components ! Although it will be a bit strong assumption to make, let us consider the parameters $a_i$'s as constants and simply investigate the dynamics of two fields, namely the universal axion and the dilaton, appearing in the same chiral multiplet $S$. Subsequently, the Hessian at one set of critical point: $\overline c_0 = -\frac{a_4}{2 \, a_5}, \overline s = \frac{\sqrt{4 a_1 \, a_5 - a_4^2}}{2 \sqrt{a_3}\, \sqrt{a_5}}$ is given as under
\bea
& & V_{c_0 c_0} = \frac{4\, \sqrt{a_3} \, a_5^{3/2}}{\sqrt{4 a_1 \, a_5 - a_4^2}}, \, V_{c_0 s} = 0 = V_{s c_0}, \, V_{s s} = \frac{4\, \sqrt{a_5} \, a_3^{3/2}}{\sqrt{4 a_1 \, a_5 - a_4^2}},
\eea
which implies that
\bea
\frac{m_{c_0}^2}{m_s^2} = \frac{a_5}{a_3} = 1.
\eea
So, with this two-field analysis, we can anticipate that it is not possible to have mass splitting of the chiral multiplet $S = e^{-\phi} - i\, C_0$ even with the inclusion of non-geometric fluxes. However as mentioned earlier, for the complete analysis, one has to investigate the full Hessian matrix of size $14\times14$, and carefully look at the non-trivial off-diagonal entries while diagonalizing the mass-matrix.
Thus, our investigation recovers the claim of \cite{Blumenhagen:2014nba} about the impossibility keeping the universal axion massless while stabilizing the dilaton in the simplest Taylor-Vafa construction \cite{Taylor:1999ii,Blumenhagen:2003vr} in the absence of non-geometric ($Q$, $P$) fluxes. In addition, our analysis supports for the validity of the first part of the No-Go theorem \cite{Blumenhagen:2014nba} that {while considering a two-field dynamics, one can not have a mass splitting in universal axion and dilaton masses even with the help of S-dual pairs of non-geometric fluxes}. However, models with additional contributions to the scalar potential
may also avoid this no-go theorem.
Such corrections can involve D-brane instanton
effects to the non-perturbative superpotential, or perturbative
corrections to the K\"ahler potential for the K\"ahler moduli, which can break the no-scale structure and such effects should be studied in great detail.
\section{Conclusion}
In this article, we propose a S-duality invariant ten-dimensional supergravity action via dimensional oxidation of a four-dimensional scalar potential, obtained by utilizing the K\"ahler- and super-potential expressions for a toroidal orientifold of type IIB superstring theory in the presence of non-geometric fluxes. In this context, we have generalized the flux orbits of \cite{Blumenhagen:2013hva} with the inclusion of RR P-flux being S-dual of the non-geometric Q-flux, and these generalized flux combinations appearing in ten-dimensional kinetic terms are as follows,
\begin{subequations}
\bea
&& {\cal H}_{ijk} = {h}_{ijk}~, \, \, \, \, \, {\cal Q}^{ij}_{k} = {Q}^{ij}_{k} - C_0 \, ~{P}^{ij}_{k} ~, \nonumber\\
&& {\cal F}_{ijk}= {f}_{ijk} - C_0 \, ~{h}_{ijk}~, \, \, \, \, \, {\cal P}^{ij}_{k} = {P}^{ij}_{k}~ .\nonumber
\eea
where
\bea
& & \hskip-1cm {h}_{ijk} = \left({H}_{ijk} +\frac{3}{2}\, {P}_{[\underline{i}}{}^{lm} C^{(4)}_{lm\underline{jk}]}\right)~, ~~~{f}_{ijk}=\left({F}_{ijk} +\frac{3}{2}\, {Q}_{[\underline{i}}{}^{lm} C^{(4)}_{lm\underline{jk}]}\,\right). \nonumber
\eea
\end{subequations}
We have motivated and exemplified the need for the use of these generalized flux-combinations in many stages; not only in nicely arranging the ten-dimensional kinetic terms out of F-term contribution of the scalar potential but also in consistently reproducing the S-dual version of the ten dimensional Chern-Simons'-terms via the D-brane tadpoles. In addition, we find that using our new flux orbits, only two propagating dofs out of the three eight-form triplet potentials ($C^{(8)}$ , $\tilde C^{(8)}$ , $C'^{(8)}$) survive which is consistent as well as desirably compatible because RR eight-form is dual to the axion-dilaton $S$.
As an application of the explicit expressions obtained, we examined the recently proposed No-Go theorem \cite{Blumenhagen:2014nba} about the impossibility of mass-splitting of axion-dilaton chiral multiplet, and investigating a two-field dynamics with fields $c_0$ and $s$ assuming that all the other moduli/axion are fixed at their minimum, we find that the No-Go result still holds with the inclusion of non-geometric Q- and its S-daul P-flux as well. However, for a final conclusion, one needs to minimize the full potential by considering the dynamics of all the 14 scalars with the presence of 64 consistent flux parameters. Further, it would be also interesting to check for the possibility alleviating the No-Go by non-perturbative effects in the presence of Non-geometric fluxes. Although, with the present poor understanding, it is hard to make any conclusion about the influence of non-geometric fluxes through non-perturbative effects, nevertheless, something robust happening at tree level would be expected to remain intact by sub-leading corrections. It would be also crucially important to perform a very detailed moduli stabilization, and to hunt for other combination of axionic directions which could be sufficiently lighter for satisfying the inflationary requirements.
\section*{Acknowledgments}
We gratefully acknowledge very significant learning on the subject from Ralph Blumenhagen and Daniela Herschmann during a previous collaboration. We are also very thankful to the referee for her/his very useful and enlightening suggestions and queries. XG would like to thank Lara Anderson, James Gray and Seung-Joo Lee for useful discussions. The work of XG was supported in part by NSF grant PHY-1417337. PS was supported by the Compagnia di San Paolo contract ``Modern Application of String Theory'' (MAST) TO-Call3-2012-0088.
\clearpage
|
\section{Introduction}
Some remarkable area inequalities for stable marginally outer trapped surfaces (MOTS) have been proven recently \cite{2011PhRvL.107e1101D},
\cite{Dain:2011kb}, \cite{2011PhRvD..84l1503J}, \cite{Simon:2011zf}, \cite{Clement:2012vb}, \cite{Fajman:2013ffa}.
In particular, for axially symmetric configurations with area $A$ and angular momentum $J$, there is the bound \cite{2011PhRvL.107e1101D},
\cite{2011PhRvD..84l1503J}
\be\label{AJ}
|J|\leq \frac{A}{8\pi},
\ee
which is saturated for extreme Kerr black holes. Although a cosmological constant $\Lambda$ does not
explicitly enter into
\eqref{AJ}, this inequality holds in the presence of a non-negative $\Lambda$.
On the other hand, when $\Lambda > 0 $,
stable MOTS obey the lower bound
\be\label{ALambda}
A \le 4\pi \Lambda^{-1},
\ee
saturated for the extreme Schwarzschild-deSitter horizon \cite{Hayward:1993tt}. This readily implies the universal
upper bound
\be\label{JLambda}
|J| \le (2\Lambda)^{-1}
\ee
which, however, can never be saturated even in theory (leaving practical considerations aside in view of the fact that $\Lambda^{-1}$ is of order $10^{122}$.
The situation bears some analogy to stable MOTS in (not necessarily axially symmetric) spacetimes with
electromagnetic fields and electric and magnetic charges $Q_E$ and $Q_M$.
In this case the inequalities $A \ge 4\pi Q^2$ \cite{Dain:2011kb} with $Q^2 = Q_E^2 + Q_M^2$
(saturated for extreme Reissner-Nordstr\"om horizons) and $A \le 4\pi
\Lambda^{-1}$ imply the (unsaturated) bound $Q^2 \le \Lambda^{-1}$.
There is however the stronger bound \cite{Simon:2011zf}
\be\label{AQLambda}
\Lambda A^2 - 4\pi A + 16 \pi^2 Q^2 \le 0
\ee
which is saturated for extreme Reissner-Nordstr\"om-deSitter configurations and,
moreover, improves the universal charge bound to $Q^2 \le (4\Lambda)^{-1}$.
Returning to the present axially symmetric case, the main objective of this article is to incorporate
explicitly the cosmological constant into inequality \eqref{AJ} and determine how it controls the allowed
values of the angular momentum. We prove the following theorem.
\begin{Theorem}
\label{main}
Let ${\cal S}$ be an axially symmetric, stable MOTS together with an
axially symmetric 4-neighborhood of ${\cal S}$ called $({\cal N}, g_{ij})$.
On $({\cal N}, g_{ij})$ we require Einstein's equations to hold,
with $\Lambda > 0$ and with matter satisfying the dominant energy condition.
Then the angular momentum $J$ and the area $A$ of ${\cal S}$ satisfy
\begin{eqnarray}
|J| & \le & \frac{A}{8\pi} \sqrt{\left(1 - \frac{\Lambda A}{4\pi} \right) \left(1 - \frac{\Lambda A}{12\pi} \right)}
\label{A1}\ ,
\\
|J| &\le & J_{max} = \frac{3 \sqrt{2}}{8 \Lambda \sqrt[4]{3}} \left(1 - \frac{1}{\sqrt{3}} \right) \approx
\frac{0.17}{\Lambda}.
\label{A2}
\end{eqnarray}
Here (\ref{A1}) is saturated precisely for the 1-parameter family of extreme Kerr-deSitter (KdS) horizons
while the universal bound (\ref{A2}) is saturated for one particular such configuration.
\end{Theorem}
The proof of this theorem will be sketched in Sect.
\ref{sec:struc}, while details are postponed to Sect. 5.
We discuss now its scope and the main differences, similarities and difficulties
compared to the ones cited above.
As $\Lambda > 0$, the main inequality \eqref{A1} is stronger than both
\eqref{AJ} and \eqref{ALambda}; in particular it forbids the black hole to
rotate as fast as its non-cosmological counterpart.
Concerning the saturation of \eqref{A1}, we observe the same pattern as in the previous inequalities: the extreme
solutions set a bound to the maximum values of charges and/or angular momentum.
The non-vanishing cosmological constant does not change this property of extreme black holes.
Inequality \eqref{A2} is obtained in a straightforward manner from \eqref{A1} and makes use of an interesting feature of the extreme KdS family.
Given $\Lambda>0$ there exists a maximum value for the angular momentum which is attained at a certain value of the area $A$. This property is
not shared by extreme Kerr horizons ($\Lambda=0$), where the value of $A$
determines the angular momentum as $8\pi |J| = A$. Note also that, as opposed to \eqref{JLambda}, \eqref{A2} is sharp and improves the numerical
factor from 0.5 to 0.17 approximately.
As stated in Theorem \ref{main}, the inequality \eqref{A1} holds between the area and angular momentum of stable MOTS's. Nevertheless, due to the analogy between stable MOTS
and stable minimal surfaces in maximal slices, one can prove an analogous
result for this type of surfaces as well (see \cite{Clement:2012vb} for a discussion of the similarities of these surfaces within the context of geometric inequalities).
Note that matter satisfying the dominant energy condition (DEC) is allowed. The energy condition is required in order to dispose of the matter terms and to arrive at the
'clean' inequality \eqref{A1} where matter does not appear explicitly. However, for electromagnetic fields we expect to obtain an inequality between area,
angular momentum, electromagnetic charges $Q_E$, $Q_M$ and cosmological constant which should reduce to \eqref{A1} for $Q=0$ and to \eqref{AQLambda} when $J=0$.
We discuss a corresponding conjecture in Sect. \ref{sec:disc}.
We now comment on the proof Theorem \ref{main} which is not a straightforward generalisation of previous results. To
explain this we recall briefly the basic strategy
of \cite{2011PhRvL.107e1101D}, \cite{2011PhRvD..84l1503J} that leads to \eqref{AJ}.
Starting with the stability condition one obtains a lower bound for the area of the MOTS in terms of a ``mass functional'' $\fm$. This $\fm$ is the key quantity in the proof,
and depends only on the twist potential and the norm of the axial Killing vector. The
non-negative cosmological constant and the matter terms
(satisfying the DEC) neither appear in $\fm$ nor later in the discussion in this
case. Therefore, the problem reduces to vacuum and with $\Lambda=0$. Then, a variational principle is used to obtain a lower bound for $\fm$.
The key point in this step is the relation between $\fm$ and the ``harmonic energy'' of maps between
the two-sphere and the hyperbolic plane. This allows to use and adapt a powerful theorem by Hildebrandt et al.
\cite{Hildebrandt}
on harmonic maps, which gives existence and uniqueness of the minimiser for $\fm$. This minimiser, in turn, gives the right hand side of \eqref{AJ}.
In the present work where we strengthen (\ref{AJ}) to
(\ref{A1}), two important obstacles appear.
Firstly, the area $A$ now appears not only as upper bound on the corresponding functional $\fm$ but also explicitly in $\fm$ itself. This makes the
variational principle hard to formulate. We overcome this problem in essence by ``freezing'' $A$ as well as $J$ to certain values
corresponding to an extreme KdS configuration, and by adapting the dynamical variables in $\fm$ suitably. Secondly,
the relation of $\fm$ to harmonic maps mentioned above no longer persists, whence the proof of existence and uniqueness of a minimiser for $\fm$ has to be done here from scratch.
We proceed by proving first that every critical point of $\fm$ is a local minimum. Finally we use the mountain pass theorem in order to get the corresponding global statement.
Our paper is organised as follows.
In Sect. 2 we recall and adapt some preliminary material, in particular the definition of angular momentum
for general 2-surfaces, as well as the definition of a stable MOTS.
In Sect. 3 we discuss relevant aspects of the KdS metric, focusing on the extreme case.
In Sect. 4 we sketch the proof of Theorem \ref{main}, postponing the core of the
argument to three key propositions which are proven in Sect. 5.
In Sect. 6 we conjecture a generalisation of our inequality to the case with
electromagnetic field along the lines mentioned above already, and we also discuss briefly the case $\Lambda < 0$.
\section{Preliminaries}
\subsection{The geometric setup}
We consider a manifold ${\cal N}$ which is topologically a 4-neighborhood of an embedded 2-surface ${\cal S}$
of spherical topology. ${\cal N}$ carries a metric $g_{ij}$ and a Levi-Civita connection $\nabla_i$.
(Latin indices from $i$ onwards run from 0 to 3, and the metric has signature $(-,+,+,+)$).
The field equations are
\begin{equation}
\label{ein}
G_{ij} = - \Lambda g_{ij} + 8\pi T_{ij}
\end{equation}
where $\Lambda$ is the cosmological constant, and the energy momentum tensor $T_{ij}$ satisfies
the dominant energy condition.
In Sections 2 and 3 we allow $\Lambda$ to have either sign; this enables us to
compare with and to carry over useful formulas from work which focuses on
Kerr-anti-deSitter, in particular \cite{Caldarelli:1999xj} and
\cite{Cho:2008vr}.
We next introduce null vectors $\ell^i$ and $k^i$ spanning
the normal plane to ${\cal S}$ and normalized as $\ell^i k_i = -1$.
We denote by $q_{ij} = g_{ij} + 2 l_{(i} k_{j)}$ the induced metric on ${\cal S}$,
the corresponding Levi-Civita connection by $D_i$ and the Ricci scalar by ${}^2\!R$.
$\epsilon_{ij}$ and $dS$ are respectively the volume element and the
area measure on ${\cal S}$. The normalisation $l_i k^i = -1$ leaves a (boost) rescaling freedom $\ell'^i =f \ell^i$, $k'^i = f^{-1}
k^i$. While this rescaling affects some quantities introduced below in an obvious
way, our key definitions such as the angular momentum (\ref{ang}) and the definition of stability
(\ref{stab}) are invariant, and the same applies to all our results.
The expansion $\theta^{(\ell)}$, the shear $\sigma^{(\ell)}_{ij}$
and the normal fundamental form $\Omega_i^{(\ell)}$
associated with the null normal $\ell^i$ are given by
\begin{equation}
\label{expsh}
\theta^{(\ell)}=q^{ij}\nabla_i\ell_j \ \ , \ \
\sigma^{(\ell)}_{ij}= {q^k}_i {q^l}_j \nabla_k \ell_l -
\frac{1}{2}\theta^{(\ell)}q_{ij} \ \, \ \
\Omega^{(\ell)}_i = -k^j {q^k}_i \nabla_k \ell_j \ .
\end{equation}
\subsection{Twist and angular momentum}
We now assume that ${\cal S}$ as well as $\Omega^{(\ell)}_i $ are axially symmetric,
i.e. there is a Killing vector $\eta^i$ on ${\cal S}$ such that
\begin{equation}
\label{Lie}
{\cal L}_\eta q_{ij} = 0 \qquad {\cal L}_\eta \Omega^{(\ell)}_{i} = 0.
\end{equation}
The field $\eta^i$ is normalized so that its integral curves have length $2\pi$.
We define the angular momentum of ${\cal S}$ as
\begin{equation}
\label{ang}
J =\frac{1}{8\pi}\int_{\cal S} \Omega_i^{(\ell)} \eta^i dS \ ,
\end{equation}
which will be related to the Komar angular momentum shortly.
By Hodge's theorem, there exist scalar fields $\omega$ and $\lambda$ on
${\cal S}$, defined up to constants,
such that $\Omega^{(\ell)}_{i}$ has the following decomposition
\begin{equation}
\label{omlam}
\Omega^{(\ell)}_{i}= \frac{1}{2\eta} \epsilon_{ij}D^j \omega +D_j \lambda.
\end{equation}
From axial symmetry it follows that
\begin{equation}
\label{eq:16}
\eta^i \Omega^{(\ell)}_{i}=
\frac{1}{2\eta} \epsilon_{ij}\eta^i D^j \omega
= \frac{1}{2}\eta^{-1/2}\xi^i D_i \omega
\end{equation}
where $\eta = \eta^i \eta_i$ and $\xi^i$ is a unit vector tangent to ${\cal S}$ and
orthogonal to $\eta^i$.
We now recall from \cite{2011PhRvL.107e1101D} that on any axially symmetric 2-surface one can
introduce a coordinate system such that
\begin{equation}
\label{can1}
q_{ij}dx^idx^j = e^{2c}e^{-\sigma} d\theta^2 + e^{\sigma} \sin^2 \theta d\varphi^2
\end{equation}
for some function $\sigma$ and a constant $c$ which is related to the area
$A$ of ${\cal S}$ via $A = 4\pi e^c$.
In such a coordinate system we can write $J$ as
\begin{equation}
\label{Jom}
J =- \frac{1}{8}\int_{0}^\pi \omega' ~ d\theta
= -\frac{1}{8} \left[\omega (\pi)- \omega(0) \right],
\end{equation}
where here and henceforth a prime denotes the derivative w.r.t. $\theta$.
From now onwards we assume that the Killing vector $\eta^i$
on ${\cal S}$ extends to ${\cal N}$ as a Killing vector of $g_{ij}$.
Of course this implies (\ref{Lie}). Moreover, it follows that ${\cal L}_{\eta} l = {\cal L}_{\eta} k = 0$.
Using the first equation we obtain
\begin{equation}\label{etaOmega}
\eta^i \Omega^{(\ell)}_{i} = - k^j \ell^i \nabla_i \eta_j.
\end{equation}
Inserting \eqref{etaOmega} in (\ref{ang}) we see that it indeed coincides with the Komar angular
momentum
\begin{equation}
\label{AK}
J = \frac{1}{8\pi} \int_{\cal S} \nabla^i \eta^j dS_{ij}.
\end{equation}
We finally introduce the twist vector
\begin{equation}
\omega_i = \epsilon_{ijkl}\eta^{j}\nabla^{k}\eta^l.
\end{equation}
If the energy momentum tensor vanishes on ${\cal N}$, we have $\nabla_{[i}\omega_{j]} = 0$.
Hence there exists a twist potential $\omega$, defined up to a constant,
such that $\omega_i = \nabla_i \omega$.
The restriction of this scalar field to ${\cal S}$ is easily seen to coincide with the
$\omega$ introduced in (\ref{omlam}), which justifies the notation.
In what follows we will refer to the pair $(\sigma,\omega)$ on ${\cal S}$ as
the \textit{data}.
\subsection{Stable marginally outer trapped surfaces}
We now take ${\cal S}$ to be a marginally trapped surface defined by $\theta^{(\ell)}=0$.
We will refer to $\ell^i$ as the {\em outgoing} null vector, which leads to
the name marginally outer trapped surface (MOTS).
Moreover, following \cite{Andersson:2007fh} (Sect. 5) we
now consider a family of two-surfaces in a neighborhood of ${\cal S}$
together with respective null normals $l_i$ and $k_i$ and we impose the following additional requirements on ${\cal S}$
and its neighborhood.
\begin{Definition}
\label{MOTS}
A marginally trapped surface ${\cal S}$
is stable if there exists an outgoing ($-k^i$-oriented) vector
$X^i= \gamma \ell^i - \psi k^i$, with
$\gamma \geq 0$ and $\psi > 0$, such that the variation $\delta_X$
of $\theta^{(\ell)}$ with respect to $X^i$ fulfills the condition
\begin{equation}
\label{stab}
\delta_X \theta^{(\ell)} \geq 0.
\end{equation}
\end{Definition}
Two remarks are in order here.
Firstly, it is easy to see (cf. Sect. 5 of \cite{Andersson:2007fh}) that stability of ${\cal S}$
w.r.t. some direction $X^i$ implies stability w.r.t all directions
``tilted away from'' $\ell^i$.
In particular, since
$\delta_{-\psi k} \theta^{(\ell)} \ge \delta_X \theta^{(\ell)}$
stability w.r.t. any $X^i$ implies stability in the past
outgoing null direction $-k^i$.
This latter condition suffices as requirement for all our results.
The other remark concerns the relation between stability and axial symmetry.
We recall that in \cite{2011PhRvL.107e1101D}, \cite{2011PhRvD..84l1503J}, inequality (\ref{AJ}) was proven under
the symmetry requirements (\ref{Lie}) and under a stability condition
similar to Definition \ref{MOTS} which, however, required $\psi$ to be
axially symmetric as well. (Axial symmetry of $\gamma$ was also assumed but
not used in the proof). In contrast, in the present theorem (\ref{main}) we
impose the stronger symmetry requirement that ${\cal S}$
as well as its neighborhood ${\cal N}$ are axially symmetric.
In this case it suffices to impose the stability condition (\ref{MOTS})
as above, namely without explicitly requiring axial symmetry of $\psi$,
since the existence of an axially symmetric function $\widetilde \psi$
then follows automatically, cf. Thm. 8.2. of \cite{Andersson:2007fh}.
Moreover, for {\it strictly} stable MOTS (which satisfy $\delta_X \theta^{(\ell)} \not\equiv 0$
in addition to (\ref{stab})) there follows even axial symmetry of the
surface itself if its neighborhood is axially symmetric (cf. Thm. 8.1. of
\cite{Andersson:2007fh}).
\section{Kerr-deSitter}
In this section we review some relevant properties of the event horizons of the Kerr-deSitter (KdS)
solutions, making use of \cite{Caldarelli:1999xj}, \cite{Cho:2008vr}, and references therein.
Other aspects of the rich and complex structure of these spacetimes can be found in
\cite{MR0424186}.
\vs
\subsection{The metric, the horizon and the angular momentum}
In ``Boyer-Lindquist'' coordinates, the KdS metric is
\begin{equation}
\label{Kerr}
ds^2 = - \frac{\zeta}{\rho^2} \left(dt - \frac{a \sin^2 \theta}{\kappa} d\phi \right)^2
+ \frac{\rho^2}{\zeta} dr^2 + \frac{\rho^2}{\chi}
d\theta^2 + \frac{\chi \sin^2 \theta}{\rho^2}
\left(a dt - \frac{r^2 + a^2}{\kappa} d\phi \right)^2
\end{equation}
where
\begin{eqnarray}
\label{zetrho}
\zeta & = & (r^2 + a^2)(1 - \frac{\Lambda r^2}{3}) - 2mr, \qquad \rho^2 = r^2 +
a^2 \cos^2 \theta \\
\label{kapchi}
\kappa & = & 1 + \frac{\Lambda a^2}{3}, \qquad \chi = 1 + \frac{\Lambda a^2 \cos^2 \theta}{3}
\end{eqnarray}
where $m\geq 0$ and $0\leq a^{2}\leq \Lambda^{-1}3(2-\sqrt{3})^{2}$.
As a function of $r$, $\zeta$ has one negative root and three positive roots (possibly counted with multiplicities).
The greatest root, $r_{ch}$, marks the cosmological horizon, while the second greatest, $r_{h}$,
marks the event horizon (from now on simply called ``horizon'').
The area of the horizon is
\begin{equation}
\label{AKdS}
A = \frac{4\pi \left(r_h^2 + a^2 \right)}{\kappa}
\end{equation}
and the induced metric on it reads
\begin{equation}
\label{ds}
ds^2 = \underbrace{\frac{\mu_h^2}{\kappa^2 \rho_h^2}}_{e^{\sigma}}
\bigg(\underbrace{\frac{\kappa^2 \rho_h^4}{\mu_h^2 \chi}}_{e^{2q}} d\theta^2 + \sin^2\theta d\phi^2
\bigg)
\end{equation}
where $\mu_h^2 = (r_h^2 + a^2)^2 \chi$ and $\rho_h =r_{h}^{2}+a^{2}\cos^{2}\theta$.
Hence
\begin{equation}
\label{spq}
e^{\sigma + q} = \frac{r_H^2 + a^2}{\kappa} = e^c = const. = \frac{A}{4\pi}
\end{equation}
and the metric is in the "canonical form" (\ref{can1}) of \cite{2011PhRvL.107e1101D}
\begin{equation}
\label{can2}
ds^2 = e^{\sigma}\left(e^{2q} d\theta^2 + \sin^2\theta d\phi^2 \right)
\end{equation}
with $\sigma + q = c = const.$.
We now calculate the twist potential $\omega(\eta)$ everywhere (not
only on ${\cal S}$), for $\eta^a = \partial/d \phi$.
Adapting a known calculation in the case $\Lambda = 0$ (cf. e.g.
Appendix A of \cite{Avila:2008te} and omitting some intermediate steps,
we find
\begin{eqnarray}
\omega' &=& \omega_{\theta} = \epsilon_{\theta\phi r t}g^{rr}g^{tt}\partial_r \eta_t +
\epsilon_{\theta\phi r t}g^{rr}g^{t\phi}\partial_r \eta_{\phi} =
- \frac{\zeta \sin \theta}{\kappa} \left(g^{tt}\partial_r g_{t\phi} +
g^{t\phi}\partial_r g_{\phi\phi} \right) = \\
{} & = & - \frac{\kappa}{\chi \sin \theta }
\left( g_{\phi\phi} \partial_r g_{t\phi} - g_{t\phi} \partial_r g_{\phi\phi}
\right) = - \frac{2ma \sin^3 \theta}{\kappa^2 \rho^2} \left[r^2 - a^2
+ \frac{2r^2}{\rho^2}\left(r^2 + a^2 \right) \right] = \label{om1} \\
{} & = & - \frac{2ma}{\kappa^2} \frac{\partial}{\partial \theta}
\left(\cos^3 \theta - 3 \cos\theta - \frac{a^2 \cos\theta \sin^4 \theta}{\rho^2} \right)
\label{om2}
\end{eqnarray}
It follows that
\be
\label{om3}
\omega = - \frac{2ma}{\kappa^2}
\left(\cos^3 \theta - 3 \cos\theta - \frac{a^2 \cos\theta \sin^4 \theta}{\rho^2} \right)
\ee
We note that compared to the case $\Lambda = 0$, $\omega$ just gets an extra
factor $1/\kappa^2$.
Integrating and using (\ref{Jom}) we obtain in particular that
\begin{equation}
\label{J}
J = am/{\kappa}^2
\end{equation}
which agrees with Equ. (2.10) of \cite{Cho:2008vr} and Equ. (18) of \cite{Caldarelli:1999xj}.
\subsection{Extreme horizons}
When at least two of the three non-negative roots
of $\zeta(r)$ coincide, (one of which is necessarily $r_{h}$), the horizon is called extremal.
When this happens the geometry near the horizon degenerates to a ``throat''. We refer to \cite{Cho:2008vr} for a further
discussion. In what follows we will just need the relation between the parameters $m,a,\Lambda, A$ and $J$
which we derive explicitly.
For extremal event horizons the radius of the limiting sphere $r_e$
satisfies, in addition to $\zeta (r_e) = 0$, the equation
\begin{equation}
\label{dp}
0 = \frac{1}{2} \frac{d \zeta}{dr}\bigg|_e = - \frac{2\Lambda r_e^3}{3} + r_e(1 - \frac{\Lambda a^2}{3}) - m.
\end{equation}
Here and henceforth a subscript $e$ indicates extremality.
Eliminating $m$ from $\zeta(r_e) = 0$ and (\ref{dp}) we obtain
\begin{equation}
\label{re}
\Lambda r_e^4 + r_e^2(\frac{\Lambda a^2}{3} - 1) + a^2 = 0.
\end{equation}
For $\Lambda \le 0$ this equation has just a single root which can be
called extremal horizon, while
for $\Lambda > 0$ there are two solutions
$r_{e} = r_{\pm}$ for given $J$
Explicitly, for $\Lambda > 0$,
\begin{equation}
\label{rEpm}
r_{\pm}^2 = \frac{1}{2 \Lambda}\left(1 - \frac{\Lambda a^2}{3} \right) \pm
\frac{1}{2 \Lambda} \sqrt{\left(1 - \frac{\Lambda a^2}{3} \right)^2 - 4 a^2 \Lambda}.
\end{equation}
When $r_{e}=r_{-}$, (and $r_{e}$ is not a triple root), the first two positive roots meet and $r_{e}<r_{ch}$, which means that a
cosmological horizon persists in spacetime. On the other hand when $r_{e}=r_{+}$, then the last two positive root
coincide and the event and the cosmological horizons become both extremal (and merge).
Using (\ref{re}) to eliminate $a^2$ from (\ref{AKdS}) we find
\begin{equation}
\label{ArE}
A = \frac{8\pi r_e^2}{1 + \Lambda r_e^2}.
\end{equation}
On the other hand, eliminating $r_e$ from (\ref{re}) and (\ref{AKdS}) gives
\begin{equation}
\label{aA}
a^2 = \frac{A}{4\pi} \frac{1 - \Lambda A/4\pi}
{ \left(1 - \Lambda A/8\pi \right) \left(1 - \Lambda A/12 \pi \right)}.
\end{equation}
In equation (\ref{re}) we eliminate now $m$ using (\ref{dp}), then $a^2$
using (\ref{re}) and finally $r_e^2$ using (\ref{re}). We obtain the
following simple relation between the angular momentum and the area for extreme K(a)dS
\begin{equation}
\label{J=A}
|J| = \mathcal{E}(A) := \frac{A}{8\pi} \sqrt{\left(1 - \frac{\Lambda A}{4\pi} \right) \left(1 - \frac{\Lambda
A}{12\pi} \right)}
\end{equation}
which after a trivial reformulation agrees with (2.32) of \cite{Cho:2008vr}.
In the case $\Lambda > 0$ and $J = 0$ the zeros of the parentheses correspond to the black hole horizon and the cosmological
horizon of Schwarzschild-deSitter, respectively.
For $\Lambda > 0$ we are only interested in the domain
$\Lambda A/4\pi < 1$ - recall that this bound can be shown for {\it all}~ stable MOTS
(irrespectively of spherical symmetry) \cite{Hayward:1993tt}.
In this range of $A$, (\ref{J=A}) takes on a maximal value
\begin{equation}
J_{max} = \frac{3 \sqrt{2}}{8 \Lambda \sqrt[4]{3}} \left(1 - \frac{1}{\sqrt{3}} \right) \approx \frac{0.17}{\Lambda}
\quad \mbox{at} \quad A_{max} = \frac{6\pi}{\Lambda} \left(1 - \frac{1}{\sqrt{3}} \right)
\end{equation}
which is the value stated in (\ref{A2}).
Moreover, for each $J$ with $|J| < J_{max}$ there are {\it two} values $A_-(J) < A_+(J)$ for the
area, cf Fig \ref{fig1}.
\begin{figure}[h!]
\centering
\begin{psfrags}
\psfrag{J}{$J$}
\psfrag{Jm}{$J_{max}$}
\psfrag{A-}{$A_-(J)$}
\psfrag{A+}{$A_+(J)$}
\psfrag{Am}{$A_{max}$}
\psfrag{A=}{$\frac{4\pi}{\Lambda}$}
\psfrag{J=E}{$J={\cal E}(A)$}
\psfrag{A}{$A$}
\includegraphics[width=10cm,height=7cm]{fig1.eps}
\end{psfrags}
\caption{The shaded region represents all points satisfying $|J|\leq \mathcal E(A)$.}
\label{fig1}
\end{figure}
We are now ready to describe the proof of Theorem \ref{main}.
\section{The structure and the proof of the main theorem}\label{sec:struc}
The main inequality
\be\label{MAININ}
|J| \leq \mathcal{E}(A)
\ee
with $\mathcal E$ given in \eqref{J=A} and $\Lambda>0$ will not be shown directly but it will follow from a related one.
This is explained in the following Theorem:
\begin{Theorem}
\label{hat}
For any given MOTS with area $A$, cosmological constant $\Lambda$ and angular momentum $J$,
there is a unique extreme KdS solution with area $\hat A$ constant $\Lambda$ and angular momentum $\hat J$ such that
\be\label{JJAA}
\frac{|J|}{A^2} = \frac{|\hat J|}{ \hat A^2},
\ee
and $\hat A \Lambda \le 4\pi$.
Moreover, the inequality $|J| \leq \mathcal{E}(A)$ is equivalent to the inequality
\be\label{AAHI}
\hat{A}\geq A.
\ee
\end{Theorem}
\begin{figure}[h!]
\centering
\begin{psfrags}
\psfrag{J}{$J$}
\psfrag{Jh}{$\hat J$}
\psfrag{Jj}{$J$}
\psfrag{Aa}{$A$}
\psfrag{Ah}{$\hat A$}
\psfrag{J=E}{$J = {\cal E}(A)$}
\psfrag{J=A}{$J = const. A^2$}
\psfrag{A}{$A$}
\includegraphics[width=10cm,height=7cm]{fig2.eps}
\end{psfrags}
\caption{The construction described in Theorem \ref{hat}}
\label{fig2}
\end{figure}
\begin{proof}[\bf Proof]
The first result, leading to equation \eqref{JJAA}, is intuitively clear from Fig \ref{fig2} since through any point
$(A,J)$ there is a unique parabola $J/A^2 = const.$, and any such parabola intersects the
``extreme'' curve $J = {\cal E}(A)$ precisely once apart from the trivial point $(0,0)$.
To state this rigorously, let $\lambda:=A/\hat{A}$ and hence $|\hat{J}|= \lambda^{2} |J|$ and $\hat A \Lambda \le 4\pi$. Then
the hatted version of (\ref{J=A}) gives a quadratic equation
for $\lambda(J,A)$. If $32 \pi^2 \sqrt{3} |J| > \Lambda A^2$ this
equation has a unique solution other than $(0,0)$. Otherwise, there are two
non-trivial solutions but only one of them lies in the region of interest $\hat A \Lambda \le 4\pi$.
To prove the equivalence between \eqref{MAININ} and \eqref{AAHI}, assume first that $\hat A\geq A$.
Then
\be\label{eq0}
A^2\leq \hat A^2=|\hat J|\frac{A^2}{|J|}=\mathcal{E}(\hat A)\frac{A^2}{|J|}=\frac{\mathcal{E}(\lambda A)}{\lambda^2}\frac{\hat A^2}{|J|}
\ee
where we have used \eqref{JJAA}, \eqref{J=A} and $\hat A= \lambda A$,
respectively.
We next use that the function $\frac{\mathcal{E}(\lambda A)}{\lambda^2}$ is monotonically decreasing
with $\lambda$ and therefore, as $\hat A\geq A$ we bound the last term as $\frac{\mathcal{E}(\lambda A)}{\lambda^2}\leq \mathcal{E}(A)$.
Putting this together with \eqref{eq0} we find
\be\label{eq2}
\hat A^2\leq \mathcal{E}(A) \frac{\hat A^2}{|J|}
\ee
which gives the desired result, that is, that \eqref{AAHI} implies \eqref{MAININ}.
To prove the converse assume $|J|\leq\mathcal E(A)$. Then $\hat J= \lambda^2 J$ and \eqref{J=A} give
\be
\mathcal E(\lambda A)= |\hat J|= \lambda^2 |J|\leq \lambda^2\mathcal E(A)
\ee
and therefore
\be
\frac{\mathcal E(\lambda A)}{\lambda^2}\leq \mathcal E(A).
\ee
Again, due to the monotonicity of the left hand side with respect to $\lambda$ we obtain $\lambda A\geq A$ which is \eqref{AAHI}.
\end{proof}
Having established the equivalence between the main inequality \eqref{MAININ} and \eqref{AAHI}, the next section will be devoted to proving
\eqref{AAHI} for a stable MOTS ${\cal S}$ with area $A$, angular momentum $J$ and data $(\sigma,\omega)$.
The proof consists of the same two steps as in the case $\Lambda = 0$. However, as we mentioned in the introduction and
as we will see below, when $\Lambda>0$ many new complications arise.
In general terms the basic steps can be described as follows.
\vs
{\rm\bf Step I.} We write the stability inequality (\ref{stab}) in terms of the data $(\sigma,\omega)$
and multiply it by an axially symmetric function $\alpha^2$
whose choice is motivated by the form of the data $(\sigma,\omega)$ of the extreme KdS horizon.
Then we integrate it on ${\cal S}$ to obtain a lower bound for $A$
in terms of the so-called mass functional $\fm$ depending on the dynamic variables $(\sigma,\omega)$.
The result is the following proposition:
\begin{Proposition}\label{PI} Let $(\sigma,\omega)$ be the data of a stable MOTS of area $A$ and angular momentum $J$.
Then, for any real number $a$ the following inequality holds
\be\label{ineq1}
\frac{A}{4\pi} \geq {\rm \eexp}^{\dfrac{{\cal M} (\sigma,\omega, A, a ) - \beta}{\ds 8\kappa}}
\ee
where the functional $\fm$ is given by
\be\label{defM}
\fm(\sigma, \omega, A, a):=\int_{0}^\pi
\bigg[ \sigma'^2 +
\frac{\omega'^2}{\eta^2}
+ 4 \sigma\frac{(1 + \Lambda a^2 \cos^2 \theta)}{\chi} + 4 \bigg(\frac{A}{4\pi}\bigg)^2\Lambda \eexp^{-\sigma} \bigg] \chi\sin\theta d\theta
\ee
where
\be
\label{beta}
\beta = \int_0^{\pi} \left(4 \chi + \frac{\chi'^2}{\chi} \right) \sin \theta d\theta
\ee
and where $\chi(a)$ has been defined in (\ref{kapchi}).
\end{Proposition}
At this stage the constant $a$ is arbitrary, but it will be fixed in the next step.
\vs
{\rm\bf Step II.}
The difficulty now is to choose $a$ conveniently and to show that, with such $a$, the r.h.s of (\ref{ineq1}) has the lower
bound $A^{2}/4\pi \hat{A}$. This would prove (\ref{AAHI}), (hence (\ref{MAININ}) by Theorem \ref{hat}).
We choose $a$ equal to the value that it would take for the extreme black hole of area $\hat{A}$. The explicit form is (\ref{aA}) with
$A$ replaced by $\hat{A}$. We will denote it by $\hat{a}$
and we denote by $\hat{\kappa}, \hat{\chi}$ and $\hat{\beta}$, the values of $\kappa$, $\chi$ and $\beta$ when $a$ is replaced by
$\hat{a}$ in (\ref{kapchi}, \ref{beta}).
Then, for the data $(\sigma, \omega)$ of the given MOTS define
\be
\label{sigom}
\hat{\sigma}:=\sigma+2\ln \lambda \qquad
\hat{\omega}= \lambda^{2}\omega,
\ee
where (again) $\lambda = A/\hat{A}$. With this change of variables we obtain
\be\label{ineq2}
\fm(\sigma,\omega,A,\hat{a})=\fm(\hat\sigma,\hat\omega,\hat{A}, \hat{a})-16\kappa\ln (\hat A/A).
\ee
Thus
\be
\frac{A}{4\pi}\geq \eexp^{\ \dfrac{\fm(\sigma,\omega,A,\hat{a})-\hat\beta}{8\hat\kappa}}=
\left(\frac{A}{\hat{A}}\right)^2\eexp^{\ \dfrac{\fm(\hat\sigma,\hat\omega,\hat{A}, \hat{a})-\hat\beta}{\ds 8\hat\kappa}}
\ee
and we need to prove
\begin{Proposition}\label{PII} In the setup explained above we have
\be\label{ineq3}
\eexp^{\ \dfrac{\fm(\hat\sigma,\hat\omega,\hat{A},\hat{a})-\hat{\beta}}{8\hat\kappa}}\geq
\frac{\hat{A}}{4\pi}.
\ee
\end{Proposition}
We wish to mention the following point here. (\ref{ineq3}) means that the
lower bound is obtained by minimising the functional $\fm(\hat\sigma,\hat\omega,\hat{A},\hat{a})$
among all pairs $(\hat\sigma,\hat\omega)$ of smooth functions with
$8\pi\hat{J}=-(\hat{\omega}(\pi)-\hat{\omega}(0))$.
A particular class of such functions has been constructed
above via (\ref{sigom}) from smooth data $(\sigma,\omega)$ on a smooth MOTS
of area $A$ and angular momentum $J$. However, this does {\it not} mean that
$(\hat\sigma,\hat\omega)$ will still form smooth data on
a smooth MOTS of area $\hat A$ and angular momentum $\hat J$. This can be seen as follows.
In order for the MOTS to be smooth (free of conical singularities), the coordinate function
$q$ must vanish at the poles,
i.e. $q(0) = q(\pi)=0$ which implies that $A = 4\pi e^{\sigma(0)} = 4\pi
e^{\sigma(\pi)}$. But inserting the scaling law (\ref{sigom}) in the latter relation contradicts the smoothness property
$\hat A = 4\pi e^{\hat \sigma(0)} = 4 \pi e^{\hat \sigma(\pi)}$ for the
hatted data, (except in the trivial case $\lambda = 1$).
Therefore, $\fm(\hat\sigma,\hat\omega,\hat{A},\hat{a})$ should
be considered as 'abstract' functional in the sense that its arguments are no longer
directly related to any MOTS. Nevertheless, extreme KdS is not only a critical point of
$\fm(\sigma,\omega,A,a)$ but also of $\fm(\hat\sigma,\hat\omega,\hat{A},\hat{a})$, and the properties of the
latter functional enable us to prove (\ref{ineq3}).
\vs
Next we present the proofs of Propositions \ref{PI} and \ref{PII}.
\section{Proof of the main propositions}
\subsection{Proof of Proposition \ref{PI}}
\begin{proof}[\bf Proof.]
The proof is analogous to the case $\Lambda = 0$ \cite{2011PhRvD..84l1503J}
to which it reduces by setting $\chi \equiv 1$. The starting point is the stability inequality (\ref{stab})
in which we take $\psi$ to be axially symmetric without loss of generality
(cf. the remarks after Definition \ref{MOTS}).
In terms of the quantities introduced in Sect. 2 we obtain, integrating
(\ref{stab}) against any axisymmetric function $\alpha:{\mathcal S}\rightarrow \mathbb{R}$,
\be
\label{PEP}
\int_{\cal S} ( |D \alpha|^2 + \frac{{}^2R}{2} \alpha^2 - \alpha^2 |\Omega^{(\ell)}|^{2} - \Lambda \alpha^2 ) dS \geq 0.
\ee
As mentioned in the previous section, we choose the trial function based on the form of
the extreme KdS geometry as
\be
\alpha=\chi^{1/2}\eexp^{-\sigma/2}.
\ee
In the coordinates (\ref{can2}) the scalar curvature takes the form
\be
{}^{2}R=\frac{\eexp^{\sigma-2c}}{\sin\theta}\bigg[-2\sigma' \cos\theta - \sin\theta \sigma'^{2}+2\sin\theta -
(\sin\theta \sigma')'\bigg].
\ee
Using this expression we obtain
\begin{align}
\label{PEQ}
\frac{1}{2\pi} \int_{\cal S} \bigg( |D \alpha|^2 +
\frac{{}^2R}{2} \alpha^2\bigg) dS = & \int_0^{\pi} \bigg(\frac{\chi \sigma'^2}{4} - \frac{\sigma' \chi' }{2} + \frac{\chi'^2}{4\chi} \bigg)\sin\theta d\theta \\
\label{SEQ} & \int_0^{\pi} \bigg(-\chi \sigma' \cos \theta - \frac{\chi \sigma'^2 \sin \theta}{2} + \chi \sin \theta
- \frac{\chi \left(\sin\theta \sigma' \right)'}{2} \bigg) d\theta.
\end{align}
Integration by parts and some rearrangement yields
\begin{align}
\label{PEQQ}
\frac{1}{2\pi} \int_{\cal S} \bigg( |D \alpha|^2 + \frac{{}^2R}{2} \alpha^2\bigg) dS = &
-\int_{0}^{\pi}\bigg[\frac{\sigma'^{2}}{4}+\sigma (1 + \frac{2\Lambda a^{2}}{3}\cos\theta)\bigg] \sin\theta d\theta\\
\label{PEQQQ}& + \int_{0}^{\pi}\bigg(\chi+\frac{\chi'^{2}}{4\chi}\bigg)\sin\theta d\theta - \chi\sigma\cos\theta\bigg|_{0}^{\pi}.
\end{align}
Using (\ref{spq}), the last term in line (\ref{PEQQQ}) above is equal to $2 \kappa \ln (A/4\pi)$. Finally, still following \cite{2011PhRvD..84l1503J}, we have
\be
-\frac{1}{2\pi} \int_{\cal S} (\alpha^2 |\Omega^{(\ell)}|^{2} + \Lambda \alpha^2) dS = -\frac{1}{4} \int_0^{\pi} \chi \frac{\omega'^2}{e^{2\sigma} \sin^4\theta}
\sin\theta d\theta - \Lambda e^{2c} \int_0^{\pi} \chi e^{-\sigma} \sin \theta d\theta.
\ee
Combining equations (\ref{PEP}), (\ref{PEQQ})-(\ref{PEQQQ}) and (\ref{PQIV})
we find
\be\label{PQIV}
2\kappa \ln \bigg(\frac{A}{4\pi}\bigg)\geq \frac{\fm - \beta}{4}
\ee
with $\beta$ as in (\ref{beta}). This expression is equivalent to (\ref{ineq1}) as wished.
\end{proof}
\subsection{Proof of Proposition \ref{PII}}
In this section we prove \eqref{ineq3} where the hatted variables $(\hat\sigma,\hat\omega)$ refer to the rescaled quantities introduced in \eqref{sigom}. To simplify the notation, for this section only, we omit the hats on these functions. With the new notation, inequality \eqref{ineq3} reads
\be\label{ineq4}
\eexp^{\ \dfrac{\fm(\sigma,\omega,\hat{A},\hat{a})-\hat{\beta}}{8\hat\kappa}}\geq
\frac{\hat{A}}{4\pi}.
\ee
As in the proof of the inequality in the $\Lambda=0$ case, this step is done by minimising the functional $\mathcal M$. We find first a minimum of $\fm$ for functions $\sigma,\omega$ defined on compact intervals $[\theta_a,\theta_b]\in(0,\pi)$ (in Prop.
\ref{LOCMIN} and \ref{GLOBAMIN}), and then take the limit
$[\theta_a,\theta_b]\to[0,\pi]$ to find \eqref{ineq4} (in Prop. \ref{Limit}). Recall that when $\Lambda=0$
the extreme Kerr geometry is the minimiser of the corresponding functional.
In this $\Lambda>0$ case, we find by a straightforward computation that extreme KdS data $(\sigma_e,\omega_e)$
is a critical point of $\fm$, that is, the explicit functions
\be
\label{EXTR}
e^{\sigma_e} = \frac{\hat \mu_e^2}{\hat \kappa^2 \hat \rho_e^2},\qquad
\omega_e'=\frac{-2 \hat \chi \hat a \hat r_e(\hat r_e^2+ \hat a^2)^2\sin^3\theta}{\hat \mu_e
\hat \rho_e^4}
\ee
satisfy the Euler-Lagrange equations of $\fm$:
\be\label{EL1a}
\frac{1}{\sin \theta}\frac{d}{d\theta}(2\hat \chi\partial_\theta\sigma\sin\theta)=-\frac{2
\hat \chi\omega'^2}{\eta^2} + 4(1 + \Lambda \hat a^2 \cos^2 \theta)-
\frac{\Lambda \hat \chi \hat A^2}{4\pi^2}e^{-\sigma}
\ee
\be\label{EL2a}
\frac{d}{d\theta}\left(\sin\theta\frac{\hat \chi \partial_\theta\omega}{\eta^2}\right)=0.
\ee
In (\ref{EXTR}), the quantities $\hat \rho_e$, $\hat \kappa_e$, $\hat \mu_e$
$\hat r_e$ and $\hat \chi_e$ were defined in (\ref{zetrho}), (\ref{kapchi}),
below (\ref{ds}) and in (\ref{re}) but carrying subscripts and
hats they refer here to the extreme KdS solution with parameter $\hat a$.
Using (\ref{re}) it is easy to
see that the above $\omega_e'$ indeed coincides with (\ref{om1}) and therefore with (\ref{om2}).
This property of extreme KdS geometry will play a fundamental role in the proof of \eqref{ineq4}, but before going into details, some
preliminary definitions are needed.
\vs
{\bf Preliminaries}. Let $0<\theta_{a}<\theta_{b}<\pi$ be fixed.
For any function $f:[\theta_{a},\theta_{b}]\rightarrow \mathbb{R}$ in $H^{1,2}$ define
\begin{align}\label{NOTAT}
& \|f\|^{2}_{2}:=\|f\|^{2}_{L^{2}}=\int_{\theta_{a}}^{\theta_{b}}f^{2}\ d\theta,\\
& \|f\|^{2}_{1,2}:=\|f\|_{H^{1,2}}=\int_{\theta_{a}}^{\theta_{b}} \left[(\partial_\theta f)^{2}+f^{2}\right]\ d\theta=\|\partial_\theta f\|^{2}_{2}+\|f\|_{2}^{2}.
\end{align}
Then, for any $\theta_{1}<\theta_{2}$, ($\theta_{a}<\theta_{1}$ and $\theta_{2}<\theta_{b}$), we have
\be
|f(\theta_{1})-f(\theta_{2})|^{2}\leq |\theta_{2}-\theta_{1}| \|f\|_{1,2}^{2}.
\ee
This says in particular that $f$ is uniformly continuous and we have
\be\label{REFBEL}
\|f-f_{a}\|^{2}_{\infty}:=\sup\big\{ (f-f_{a})^{2}(\theta): \theta\in [\theta_{a},\theta_{b}]\big\}\leq \pi \|\partial_\theta f\|^{2}_{2}\leq \pi \|f\|_{1,2}^{2}
\ee
where $f_{a}=f(\theta_{a})$.
We will use the affine space $\Gamma_{ab}$ of $H^{1,2}$ paths $\gamma:[\theta_{a},\theta_{b}]\rightarrow \mathbb{R}^{2}$, $\gamma=(\sigma,\omega)$, such that
\be\label{BOUNDAT}
(\sigma(\theta_{a}),\omega(\theta_{a}))=(\sigma_{e}(\theta_{a}),\omega_{e}(\theta_{a}))\quad \text{and}\quad (\sigma(\theta_{b}),\omega(\theta_{b}))=(\sigma_{e}(\theta_{b}),\omega_{e}(\theta_{b})),
\ee
where $(\sigma_{e},\omega_{e})$ are the data of extreme KdS of area $\hat A$.
In line with the notation (\ref{NOTAT}) we use the
shorthand $\| \gamma_{1}-\gamma_{2}\|^{2}_{1,2}:=\|\sigma_{1}-\sigma_{2}\|_{1,2}^{2}+\|\omega_{1}-\omega_{2}\|_{1,2}^{2}$.
Let $\fm_{ab}=\fm_{ab}(\gamma):\Gamma_{ab}\rightarrow \mathbb{R}$ be the functional given by
\be\label{FMAB}
\fm_{ab}(\gamma)=\int_{\theta_{a}}^{\theta_{b}} \bigg((\partial_{\theta}\sigma)^{2}+
4\sigma\frac{(1+\Lambda \hat{a}^{2}\cos^{2}\theta)}{\hat\chi}+
\frac{(\partial_{\theta}\omega)^{2}}{\eexp^{2\sigma}\sin^{4}\theta}+
4\Lambda \bigg(\frac{\hat A}{4\pi}\bigg)^{2}\eexp^{-\sigma}\bigg)\, \hat\chi\, \sin\theta d\theta.
\ee
Note that this functional is the same as the $\fm$ appearing in \eqref{ineq4} except that the integration is over $[\theta_{a},\theta_{b}]$ and that the arguments
$\gamma=(\sigma,\omega)$ vary in $\Gamma_{ab}$.
\vs
{\bf The functional $\olfm_{ab}$}. Consider the change of variables $(\theta,\sigma,\omega)\rightarrow
(\bar{\theta}, \bar{\sigma},\bar{\omega})$ given by
\be\label{CHANVAR}
\frac{d\bar{\theta}}{d\theta}=\frac{\sin\bar{\theta}}{\sin\theta\hat\chi(\theta)},\quad \bar{\sigma}=
\sigma+2\ln \frac{\sin\theta}{\sin\bar{\theta}},\quad \bar{\omega}=\omega.
\ee
Explicitly, $\bar \theta(\theta)$ reads, with a suitable choice of the integration
constant,
\begin{equation}
\label{tr}
\tan \frac{\bar \theta}{2} =
\left( \tan \frac{\theta}{2} \right)^{1/\hat \kappa}
\exp \left[ - \frac{\hat a}{\hat \kappa} \sqrt{\frac{\Lambda}{3}}
\arctan \left( \hat a \sqrt{\frac{\Lambda}{3}} \cos \theta \right) \right].
\end{equation}
It follows that the map $\theta\rightarrow \bar{\theta}$ is a diffeomorphism from $[0,\pi]$ into $[0,\pi]$ and that
$0<c_{1}<(\sin\theta/\sin\bar{\theta})<c_{2}<\infty$ for $c_{1}$ and $c_{2}$ depending only on $\hat{a}^{2}\Lambda$.
The transformation of the affine space $\Gamma_{ab}$ will be denoted by $\overline{\Gamma}_{ab}$. A straightforward computation shows
\begin{eqnarray}
\label{MM}
\fm_{ab}(\gamma)&=&\olfm_{ab}(\bar\gamma)+\\\nonumber
&+&\int_{\bar\theta_{a}}^{\bar\theta_{b}}\frac{4\cos^2\bar\theta}{\sin\bar\theta}d\bar\theta+4\bar\sigma\cos\bar\theta\bigg|_{\bar\theta_{a}}^{\bar\theta_{b}}-
\left.4\sigma(\cos\theta+\frac{\hat{a}^2\Lambda}{3}\cos^3\theta)\right|_{\theta_a}^ {\theta_b}-\int_{\theta_a}^ {\theta_b}\frac{4\hat\chi \cos^2\theta}{\sin\theta} d\theta
\end{eqnarray}
where the functional $\olfm_{ab}=\olfm_{ab}(\bar{\gamma}):\overline{\Gamma}_{ab}\rightarrow \mathbb{R}$ is given by
\be\label{OLFM}
\olfm_{ab}(\bar{\gamma})=\int_{\bar{\theta}_{a}}^{\bar{\theta}_{b}}\bigg((\partial_{\bar{\theta}}\bar{\sigma})^{2}+
4\bar{\sigma}+\frac{(\partial_{\theta}\bar\omega)^{2}}{\eexp^{2\bar\sigma}\sin^{4}\bar\theta}+
\bigg[4\Lambda \bigg(\frac{\hat A}{4\pi}\bigg)^{2}\frac{\sin^{4}\theta}{\sin^{4}\bar\theta}\hat{\chi}^{2}(\theta)\bigg]
\eexp^{-\bar\sigma}\bigg)\sin\bar\theta d\bar\theta.
\ee
Thus, the functionals $\fm_{ab}:\Gamma_{ab}\rightarrow \mathbb{R}$ and $\olfm_{ab}:\overline{\Gamma}_{ab}:\rightarrow \mathbb{R}$ differ by a constant and boundary terms.
This immediately implies that $\gamma$ is a critical point of $\fm_{ab}$ iff $\bar{\gamma}$ is a critical point of $\olfm_{ab}$. In particular as $\gamma_e$ is a critical point of
$\fm_{ab}$, $\bar\gamma_e$ is a critical point of $\olfm_{ab}$. As we will explain below, the nature of
critical points of the functional $\olfm_{ab}$ can be easily analysed via a crucial formula due to Carter. A similar simple formula to analyse the critical points of $\fm_{ab}$ is
unknown to us. For this reason we will continue working with $\olfm_{ab}$ rather than with $\fm_{ab}$.
\vs
{\bf The results}. The next three propositions together prove Proposition \ref{PII}.
Propositions \ref{LOCMIN} and \ref{GLOBAMIN} deal with the minimisation of the restricted functional $\olfm_{ab}$. Then,
Proposition ref{Limit} establishes the connection between
the minimisation of $\olfm_{ab}$ (or, equivalently, the minimisation of $\fm_{ab}$) and the minimisation of the original
functional $\fm$ that ultimately leads to \eqref{ineq4} and Proposition \ref{PII}. The angles $\theta_{a},\theta_{b}\in(0,\pi)$ defining $\olfm_{ab}$ are arbitrary.
\begin{Proposition}\label{LOCMIN}
For any critical point $\bar{\gamma}_{c}$ of $\olfm_{ab}$ there are constants $\epsilon>0$ and $c>0$,
such that if $\|\bar\gamma - \bar{\gamma}_{c}\|_{1,2}\leq \epsilon$ then
\be\label{PROPINEQ}
\olfm_{ab}(\bar\gamma)\geq \olfm_{ab}(\bar{\gamma}_{c})+c\|\bar\gamma-\bar{\gamma}_{c}\|^{2}_{1,2}.
\ee
In particular $\olfm_{ab}$ achieves a strict local minimum at any of its critical points.
\end{Proposition}
\begin{Proposition}\label{GLOBAMIN} $\olfm_{ab}$ has only one critical point $\bar{\gamma}_{c}=\bar{\gamma}_{e}$ and
$\olfm_{ab}(\bar{\gamma}_{e})$ is a global minimum, \textit{i.e.}
\be\label{eqprop5.2}
\olfm_{ab}(\bar\gamma)\geq \olfm_{ab}(\bar{\gamma_e}).
\ee
\end{Proposition}
\begin{Proposition}\label{Limit}
We have
\be\label{eqprop5.3a}
\fm_{ab}(\gamma,\hat{A},\hat{a})\geq \fm_{ab}(\gamma_e,\hat{A},\hat{a})
\ee
for functions $\gamma=(\sigma,\omega)$ having the boundary values $\gamma|_{\theta_a,\theta_b}=\gamma_e|_{\theta_a,\theta_b}$.
Moreover, taking the limit $[\theta_a,\theta_b]\to[0,\pi]$ we have
\be\label{eqlimit}
\fm(\gamma,\hat{A},\hat{a})\geq \fm(\gamma_e,\hat{A},\hat{a}).
\ee
The explicit form of $\fm(\gamma_e)$ gives \eqref{ineq4}.
\end{Proposition}
Note that taking the limit $(\theta_a,\theta_b)\to (0,\pi)$ is a very delicate issue as the limit boundary values of
$\sigma$ are not necessarily the same as those of $\sigma_e$. We will treat this problem following the ideas of \cite{2011CQGra..28j5014A}.
\begin{proof}[\bf Proof of Proposition \ref{LOCMIN}.] For given $\bar\gamma$ let $\tilde\gamma=(\tilde\sigma,\tilde\omega):=
\bar\gamma-\bar\gamma_{c}$ and define the path $\bar\gamma_{\tau}=\bar{\gamma}_{c}+\tau \tilde\gamma$ for $\tau$ in $[0,1]$. The Taylor expansion of $\olfm_{ab}(\bar\gamma_{\tau})$ at $\tau=0$ gives
\be\label{TAY}
\olfm_{ab}(\bar\gamma)=\olfm_{ab}(\bar{\gamma}_{c})+\frac{1}{2}\partial^{2}_{\tau}\olfm_{ab}(\bar{\gamma}_{\tau})\big|_{\tau=0}+\frac{1}{6}\partial^{3}_{\tau}\olfm_{ab}(\bar\gamma_{\tau})\big|_{\tau=\tau^*}
\ee
where $0\leq \tau^{*}\leq 1$. The proof of Proposition (\ref{LOCMIN}) comes from analysing the last two terms
on the right hand side of (\ref{TAY}). We do that separately.
To simplify notation set $\olfm_{ab}(\bar{\gamma}_{\tau})=\olfm_{ab}$.
Moreover, in the present proof primes on functions denote derivatives $\partial_{\bar \theta}$.
The first $\tau$-derivative of $\olfm_{ab}$ as a function of $\tau$ is
\be
\partial_{\tau}\olfm_{ab}=2\int_{\bar{\theta}_{a}}^{\bar{\theta}_{b}}
\left[\wD\tilde\sigma\cdot \wD\bar\sigma+2\bar\sigma+\frac{ \wD\tilde\omega\cdot
\wD\bar\omega-
\tilde\sigma(\wD\bar\omega)^2}{\bar\eta^2}- \frac{\tilde\sigma\fv\eexp^{-\bar\sigma}}{2}\right]\sin\bar\theta d\bar\theta,
\ee
where
\be
\fv:=4\Lambda \bigg(\frac{\hat A}{4\pi}\bigg)^{2}\frac{\sin^{4}\theta}{\sin^{4}\bar\theta}\hat{\chi}^{2}(\theta)
\ee
and the derivative operator $\wD$ and the dot products are taken with respect to the standard metric on
$S^2$. (Due to axisymmetry $\wD=\partial_{\bar{\theta}}$).
Evaluate at $\tau=0$, integrate by parts and use the boundary conditions to obtain the Euler-Lagrange equations
for $\olfm_{ab}$, namely
\begin{align}
\label{EL1} & \widehat\Delta\bar\sigma_c-2+\frac{(\wD\bar\omega_c)^2}{\bar\eta_c^2}=-\frac{\fv}{2}\eexp^{-\bar\sigma_c}, \\
\label{EL2} & \wD\bigg(\frac{ \wD\bar\omega_c}{\bar\eta_c^2}\bigg)=0,
\end{align}
where $\bar\eta_c=e^{\bar\sigma_c}\sin^2\bar\theta$, and $\widehat \Delta$ is the Laplace operator with respect to the
standard metric on $\mathbb{S}^2$. (Again, due to axisymmetry, $\widehat \Delta$ involves only derivatives with respect to $\bar\theta$).
The second $\tau$-derivative of $\olfm_{ab}$ reads
\be
\partial_{\tau}^{2}\olfm_{ab}=2\int_{\tilde\theta_{a}}^{\tilde\theta_{b}}\left[(\wD\tilde\sigma)^2+\frac{2\tilde\sigma^2(
\wD\bar\omega)^2-4\tilde\sigma \wD\bar\omega\cdot
\wD\tilde\omega+(\wD\tilde\omega)^2}{\bar\eta^2}+\frac{\tilde\sigma^2\fv \eexp^{-\bar\sigma}}{2}\right]\sin\bar\theta d\bar\theta.
\ee
Next, recall Carter's identity in the form (see \cite{Dain:2010qr})
\be
\label{Carter}
F+ \tilde\sigma G'_{\bar\sigma}+\tilde\omega G'_{\bar\omega}+2\tilde\sigma\tilde\omega G_{\bar\omega}-
{\bar\eta^{-2}}\tilde\omega^2G_{\bar\sigma}=H,
\ee
where
\begin{align}
& G_{\bar\sigma}(\tau)=\widehat \Delta\bar\sigma+\bar\eta^{-2}(\wD\bar\omega)^2-2,\\
& G_{\bar\omega}(\tau)= \wD\left(\bar\eta^{-2} \wD\bar\omega\right),
\end{align}
\begin{align}
& G'_{\bar\sigma}(\tau)=\widehat \Delta \tilde\sigma + \bar\eta^{-2}(2 \wD \tilde\omega .
\wD\bar\omega - 2\tilde\sigma (\wD \bar \omega)^{2}),\\
& G'_{\bar{\omega}}(\tau)=\wD(\bar\eta^{-2}(\wD\tilde\omega -2\tilde\sigma \wD\bar\omega)),
\end{align}
and
\begin{align}
& F(\tau)=(\wD\tilde\sigma+\tilde\omega\bar\eta^{-2}\wD\bar\omega)^2+(\wD(\tilde\omega \bar\eta^{-1}-
\bar\eta^{-1}\tilde\sigma \wD\bar\omega))^2+(\bar\eta^{-1}\tilde\sigma \wD\bar\omega-\tilde\omega \bar\eta^{-2}\wD\bar\eta)^2,\\
& H(\tau)= \wD(\tilde\sigma \wD\tilde\sigma+\tilde\omega \bar\eta^{-1}\wD(\tilde\omega \bar\eta^{-1})).
\end{align}
Now we can use the expressions for $G'_{\bar\sigma}$ and $G'_{\bar\omega}$ to obtain, after a simple integration by parts,
\be
\partial_{\tau}^{2}\olfm_{ab}=-2\int_{\bar\theta_{a}}^{\bar\theta_{b}}(\tilde\sigma G'_{\bar\sigma}+\tilde\omega G'_{\bar\omega}-
\frac{1}{2}\tilde\sigma^2 \fv \eexp^{-\bar\sigma})\sin\bar\theta d\bar\theta.
\ee
Using \eqref{Carter}, integrating by parts once again and using the boundary conditions $\tilde{\sigma}(\bar\theta_{a})=
\tilde{\sigma}(\bar\theta_{b})=0$, $\tilde{\omega}(\bar\theta_{a})=\tilde{\omega}(\bar\theta_{b})=0$ to get rid of $H$,
yields
\be
\partial^{2}_{\tau}\olfm_{ab}=2\int_{\bar\theta_{a}}^{\bar\theta_{b}}(F+2\tilde\sigma\tilde\omega G_{\bar\omega}-
\bar\eta^{-2}\tilde\omega^2 G_{\bar\sigma}+\frac{1}{2}\tilde\sigma^2 \fv \eexp^{-\bar\sigma})\sin\bar\theta d\bar\theta.
\ee
Evaluating at $\tau=0$ and using the Euler-Lagrange equations, we obtain
\be
\partial^{2}_{\tau}\olfm_{ab}\big|_{\tau=0}=2\int_{\bar\theta_{a}}^{\bar\theta_{b}} (F+\frac{1}{2}(\bar\eta_c^{-2}\tilde\omega^2 +
\tilde\sigma^2)\fv\eexp^{-\bar\sigma_{c}})\sin\bar\theta d\bar\theta,
\ee
which can be written in the form
\begin{align}
\partial_{\tau}^{2}\olfm_{ab}\big|_{\tau=0}=\label{LIN} 2\int_{\theta_{a}}^{\theta_{b}}
\bigg\{\ & \bigg(\tilde{\sigma}'+\bigg(\frac{{\omega_{c}}'}{\bar\eta_{c}}\bigg)\bigg[\frac{\tilde{\omega}}{\bar\eta_{c}}\bigg]\bigg)^{2}+\bigg(\bigg[\frac{\tilde{\omega}}{\bar\eta_{c}}\bigg]'-\bigg(\frac{\omega_{c}'}{\bar\eta_{c}}\bigg)\tilde{\sigma}\bigg)^{2} \\
\label{F1} & +\bigg(\frac{\tilde{\sigma}\omega_{c}'}{\bar\eta_{c}}-\frac{\tilde{\omega}\bar\eta_{c}'}{\bar\eta_{c}^{2}}\bigg)^{2}+
\frac{\fv}{2}\bigg(\bigg[\frac{\tilde{\omega}}{\bar\eta_{c}}\bigg]^{2}+\tilde{\sigma}^{2}\bigg)\eexp^{-\bar\sigma_{c}}\ \bigg\}\sin\bar\theta d\bar\theta.
\end{align}
We proceed by taking advantage of this formula.
\vs
First we note that because $\bar\gamma_c$ is a critical point we have $\omega_c'/\bar\eta_c^2=k/\sin\bar{\theta}$ where $k$ is a constant. Write
$\tilde{\tilde\omega}:=\tilde\omega/\bar\eta_c$ and disregard the first term in \eqref{F1}. We get
\begin{equation}
\label{F2}
\partial^{2}_{\tau}\olfm_{ab}\big|_{\tau=0}\geq
2\int_{\bar{\theta}_{a}}^{\bar{\theta}_{b}}\bigg\{ \bigg(\tilde{\sigma}'+
\bigg(\frac{k\bar\eta_c}{\sin\bar{\theta}}\bigg)\tilde{\tilde{\omega}}\bigg)^{2}
+\bigg(\tilde{\tilde{\omega}}'-\bigg(\frac{k\bar\eta_c}{\sin\bar{\theta}}\bigg)\tilde{\sigma}\bigg)^{2}
+\frac{\fv}{2}(\tilde{\tilde{\omega}}^2+\tilde{\sigma}^{2})\eexp^{-\bar{\sigma}_{c}}\ \bigg\}\sin\bar\theta d\bar\theta.
\end{equation}
Let $s:=\min\big\{(\sin\bar\theta)/\bar\eta_c\big\}$ and assume
\be\label{sup1}
\int_\Omega\tilde\sigma'^2\sin\bar\theta d\bar\theta>\frac{4k^2}{s^2}\int_\Omega\tilde{\tilde\omega}^2\sin\bar\theta d\bar\theta
\ee
Then the first term in \eqref{F2} can be bounded as
\begin{align}
\bigg[\int_{\bar\theta_{a}}^{\bar\theta_{b}}\left(\tilde\sigma'+\bigg(\frac{k\bar\eta_c}{\sin\bar\theta}\bigg)\tilde{\tilde{\omega}}\right)^2
& \sin\bar\theta d\bar\theta \bigg]^{1/2}\geq \\
& \geq \bigg[\int_{\bar\theta_{a}}^{\bar\theta_{b}}\tilde\sigma'^2\sin\bar\theta d\bar\theta\bigg]^{1/2}-
\bigg[\int_{\bar\theta_{a}}^{\bar\theta_{b}}\bigg(\frac{k^2\bar\eta_c^2}{\sin^2\bar\theta}\bigg)\tilde{\tilde\omega}^2\sin\bar\theta d\bar\theta\bigg]^{1/2}\\
& \geq\bigg[\int_{\bar\theta_{a}}^{\bar\theta_{b}}\tilde\sigma'^2\sin\bar\theta d\bar\theta\bigg]^{1/2}-
\frac{|k|}{s}\bigg[\int_{\bar\theta_{a}}^{\bar\theta_{b}}\tilde{\tilde\omega}^2\sin\bar\theta d\bar\theta\bigg]^{1/2}\\
\label{eq1} &\geq\bigg[\int_{\bar\theta_{a}}^{\bar\theta_{b}}\tilde\sigma'^2\sin\bar\theta d\bar\theta\bigg]^{1/2}-
\frac{1}{2}\bigg[\int_{\bar\theta_{a}}^{\bar\theta_{b}}\tilde\sigma'^2\sin\bar\theta d\bar\theta\bigg]^{1/2}\\
& =\frac{1}{2}\bigg[\int_{\bar\theta_{a}}^{\bar\theta_{b}}\tilde\sigma'^2\sin\bar\theta d\bar\theta\bigg]^{1/2}
\geq \min\{\frac{\sin^{1/2}\bar\theta}{2}\}\bigg[\int_{\bar\theta_{a}}^{\bar\theta_{b}}\tilde\sigma'^2 d\bar\theta\bigg]^{1/2},
\end{align}
where \eqref{eq1} has been obtained
using \eqref{sup1}. This bound together with the last term in \eqref{F2}
gives us
\be
\label{boundsigma}
\partial^{2}_{\tau}\olfm_{ab}\big|_{\tau=0}\geq c_{1}\|\tilde\sigma\|^2_{1,2}
\ee
for some constant $c_{1}>0$.
Now assume that the opposite to \eqref{sup1} holds, namely
\be\label{sup2}
\int_\Omega\tilde\sigma'^2\sin\bar\theta d\bar\theta\leq\frac{4k^2}{s^2}\int_\Omega\tilde{\tilde\omega}^2\sin\bar\theta d\bar\theta.
\ee
Then from \eqref{F2} we have
\begin{align}\label{F3}
\partial_{\tau}^{2}\olfm_{ab}|_{\tau=0}&\geq
\int_{\Omega}\fv(\tilde{\tilde{\omega}}^2+\tilde{\sigma}^{2})\eexp^{-\bar\sigma_{c}}\sin\bar\theta d\bar\theta\\
&\geq\min\{\fv\eexp^{-\bar\sigma_{c}}\}\int_{\bar\theta_{a}}^{\bar\theta_{b}}(\tilde{\tilde{\omega}}^2+\tilde{\sigma}^{2})\sin\bar\theta d\bar\theta\\
&\geq \min\{\fv\eexp^{-\bar\sigma_{c}}\}\int_{\bar\theta_{a}}^{\bar\theta_{b}}(\frac{s^2}{4k^2}\tilde\sigma'^{2}+\tilde\sigma^2)\sin\bar\theta d\bar\theta \\
&\geq \min\{\fv\eexp^{-\bar\sigma_{c}}\}\min\{1, \frac{s^2}{4k^2}\}\int_{\bar\theta_{a}}^{\bar\theta_{b}}(\tilde\sigma'^{2}+\tilde\sigma^2)\sin\bar\theta d\bar\theta
\end{align}
which again gives us an inequality $\partial^{2}_{\tau}\olfm_{ab}|_{\tau=0}\geq c_{2}\|\tilde\sigma\|^2_{1,2}$
for some constant $c_{2}>0$. Thus in either case we have
\be\label{boundsigma2}
\partial^{2}_{\tau}\olfm_{ab}\big|_{\tau=0}\geq c_{3}\|\tilde\sigma\|^2_{1,2}
\ee
for some constant $c_{3}>0$.
Now we can interchange the roles of $\tilde\sigma$ and $\tilde{\tilde\omega}$
(observing the symmetry in (\ref{F2})) to find again
\be\label{boundomega}
\partial^{2}_{\tau}\olfm_{ab}\big|_{\tau=0}\geq c_{3}\|\tilde{\tilde{\omega}}\|^2_{1,2}.
\ee
Using that $\tilde{\tilde{\omega}}=\tilde\omega/\eta_c$ and by an argument similar to the previous one we deduce from (\ref{boundomega}) that
\be\label{boundtomega}
\partial^{2}_{\tau}\olfm_{ab}\big|_{\tau=0}\geq c_{4}\|\tilde{\omega}\|^2_{1,2}
\ee
for some constant $c_{4}>0$. Collecting \eqref{boundsigma2} and \eqref{boundtomega} we get
\be\label{boundM2}
\partial^{2}_{\tau}\olfm_{ab}\big|_{\tau=0}\geq c_{5}\|(\tilde\sigma,\tilde{\omega})\|^2_{1,2}=c_{5}\|\bar\gamma-\bar\gamma_c\|^2_{1,2}
\ee
for some constant $c_{5}>0$.
Having treated the second term on the right hand side of (\ref{TAY}) we turn to the last one.
We claim that there is a constant $c_{6}>0$ such that if $\|\bar\gamma-\bar{\gamma}_{c}\|_{1,2}\leq 1$ then
\be\label{BOUND}
\partial^{3}_{\tau}\olfm_{ab}\big|_{\tau=\tau^{*}}\leq c_{6}\|(\tilde{\sigma},\tilde{\omega})\|_{1,2}^{3}.
\ee
Combined with (\ref{TAY}) this would show, as we want, that if $\|(\tilde{\sigma},\tilde{\omega})\|_{1,2}\leq \epsilon$
for $\epsilon$ sufficiently small, then (\ref{PROPINEQ}) holds for some constant $c>0$. The bound (\ref{BOUND}) is indeed
easily obtained. A direct computation gives
\be
\partial^{3}_{\tau}\olfm_{ab}=-2\int_{\bar\theta_{a}}^{\bar\theta_{b}} \bigg(\frac{6\tilde{\sigma}\tilde{\omega}'^{2}-
12\tilde{\sigma}^{2}\tilde{\omega}'\bar\omega'+
4\tilde{\sigma}^{3}\bar\omega'^{2}}{\bar\eta^{2}}+\frac{\fv}{2}{\tilde\sigma}^{3}\eexp^{-\tilde\sigma}\bigg)\sin\bar\theta d\bar\theta
\ee
Bounds for each term in this integral,
compatible with (\ref{BOUND}), are obtained by using that $\|\tilde{\sigma}\|_{\infty}\leq \sqrt{\pi}\|\tilde{\sigma}\|_{1,2}\leq
\sqrt{\pi} \|(\tilde{\sigma},
\tilde{\omega})\|_{1,2}$, and that if $\|\bar\gamma-\bar{\gamma}_{c}\|_{1,2}\leq 1$ then $\|\bar\sigma\|_{\infty}\leq c_{7}$ and
$\|\bar\omega'\|_{2}\leq c_{8}$ for constants $c_{7}>0$ and $c_{8}>0$. For instance the first term is bounded as
\be
\bigg|12\int_{\bar\theta_{a}}^{\bar\theta_{b}}\frac{\tilde{\sigma}\tilde{\omega}'^{2}}{\bar\eta^{2}}\sin\bar\theta d\bar\theta\bigg|\leq
12 \sup\big\{\frac{1}{\sin^{3}\theta}\big\}\, e^{2c_{7}}\|\tilde{\sigma}\|_{\infty}\|\tilde{\omega}'\|^{2}_{2}\leq
c_{9}\|(\tilde{\sigma},\tilde{\omega})\|_{1,2}^{3}
\ee
for some constant $c_{9}>0$. The other terms are bounded in the same way.
\end{proof}
\begin{proof}[\bf Proof of Proposition \ref{GLOBAMIN}.]
It will be more convenient to work with the functional $\fm_{ab}^{*}(\gamma^{*})$ of the arguments $\gamma^{*}=(u,\omega)$ with $u=-\ln\eta$, given by
\be\label{FMS}
\fm_{ab}^{*}(\gamma^{*})=\int_{\bar\theta_{a}}^{\bar\theta_{b}} (u'^{2}+\omega'^{2}e^{2u} + \fv^{*} e^{u})\,\sin\bar\theta d\bar\theta
\ee
where
\be
\fv^{*}=\fv \sin^{2}\bar{\theta}.
\ee
This functional is equal to $\fm_{ab}(\gamma)$ plus a constant independent of the
arguments. (Use $u=-\ln \eta$ in (\ref{OLFM})).
If $\fm^{*}_{ab}$ is shown to satisfy the Palais-Smale (PS) condition (see below), then a simple application of Proposition
\ref{LOCMIN} and the mountain pass theorem,
as explained in the {\it Corollary} on page 187 of \cite{MR766489},
shows that $\gamma_{e}^{*}=(\ln \eta_{e},\omega_{e})$ is the only critical point and that $\fm^{*}_{ab}(\gamma_{e}^{*})$
is the strict absolute minimum of $\fm^{*}_{ab}$.
We explain now how to verify the PS condition. Recall first that the PS condition holds iff
any sequence $\gamma^{*}_{i}$ for which
$\fm^{*}_{ab}(\gamma^{*}_{i})$ is bounded and for which $\|\delta \fm^{*}_{ab} (\gamma^{*}_{i})\|\rightarrow 0$ has a (strongly)
convergent subsequence.
Here $\|\delta \fm^{*}_{ab} (\gamma^{*}_{i})\|$ is the norm of the differential of $\fm^{*}_{ab}$ at
$\gamma^{*}_{i}$. Recall that this norm is $\|\delta \fm_{ab}^{*}(\gamma^{*})\| =
\sup\big\{|\delta_{X} \fm^{*}_{ab}(\gamma^{*})|: \|X\|_{1,2}=1\big\}$. Note from this definition
that if $\|\delta \fm_{ab}^{*}(\gamma^{*}_{i})\|\rightarrow 0$, then for any sequence $X_{i}$
with $\|X_{i}\|_{1,2}\leq K$ we have
\be\label{TOMEN}
|\delta_{X_{i}} \fm_{ab}^{*}(\gamma^{*}_{i})|\rightarrow 0.
\ee
Now, for any tangent vector $X=(\tilde{u},\tilde{\omega})$ to a point $\gamma^{*}=(u,\omega)$ we compute
\be\label{FVAR}
\delta_{X}\fm_{ab}^{*} (\gamma^{*}) = \int_{\bar\theta_{a}}^{\bar\theta_{b}} (2\tilde{u}'u'+2 \tilde{u}\omega'^{2} e^{2u} +
2\tilde{\omega}' \omega' e^{2u}+ \tilde{u} \fv \sin^{2}\bar{\theta} e^{u})\, \sin\bar\theta d\bar\theta.
\ee
This expression will be used below.
Let $\gamma^{*}_{i}$ be a sequence such that $\fm_{ab}^{*}(\gamma^{*}_{i})$ is uniformly bounded and such that
$\|\delta \fm_{ab}^{*}(\gamma_{i}^{*})\|\rightarrow 0$. From (\ref{FMS}) we deduce that $\|u'_{i}\|_{2}$ is uniformly bounded
\footnote{Note that there are constants $0<c_{1}<c_{2}<\infty$ such that $c_{1}<\sin\theta<c_{2}$.} and from
this and (\ref{REFBEL}) that $u_{i}$ is uniformly bounded and uniformly continuous. By
the theorem of Arzel\`a -Ascoli,
$u_{i}$ has a $C^{0}$-convergent subsequence (that we still index by `$i$'). As $\|u_{i}\|_{1,2}$ is
uniformly bounded we can assume that $u_{i}$ converges weakly in $H^{1,2}$ too.
Then, from the $C^{0}$-boundedness of $u_{i}$ and again from (\ref{FMS}),
we deduce in a similar fashion that $\omega_{i}$ has a subsequence converging in $C^{0}$ and weakly in $H^{1,2}$.
Assume then without loss of generality that for the above sequence $\gamma_{i}^{*}$
we have $u_{i}\rightarrow u_{\infty}$ and $\omega_{i}\rightarrow \omega_{\infty}$ weakly in $H^{1,2}$ and
strongly in $C^{0}$. Let $c>0$ be a constant such that $c<e^{2u_{i}}\sin\bar\theta$ for all $i$. Then,
\begin{align}
\nonumber & c\int_{\bar\theta_{a}}^{\bar\theta_{b}} (\omega'_{i}-\omega'_{\infty} )^{2}d\bar\theta\leq
\int_{\bar\theta_{a}}^{\bar\theta_{b}} (\omega'_{i}-\omega_{\infty}')^{2}e^{2u_{i}}\,\sin\bar\theta d\bar\theta=\\
\label{Z} & = \bigg( \int_{\bar\theta_{a}}^{\bar\theta_{b}} \omega_{i}'(\omega'_{i}-\omega_{\infty}') e^{2u_{i}}\,
\sin\bar\theta d\bar\theta -
\int_{\theta_{a}}^{\bar\theta_{b}} \omega_{\infty}'(\omega'_{i}-\omega_{\infty}') e^{2u_{i}}\,\sin\bar\theta
d\bar\theta \bigg)\rightarrow 0
\end{align}
where the first integral in (\ref{Z}) is seen to go to zero by taking $V_{i}=(0,\tilde{\omega}_{i})$ with
$\tilde{\omega}_{i}=\omega_{i}-\omega_{\infty}$ in (\ref{TOMEN}),
while the second integral in (\ref{Z}) tends to zero because $\omega_{i}\rightarrow \omega_{\infty}$
weakly in $H^{1,2}$ and $u_{i}\rightarrow u_{\infty}$ strongly in $C^{0}$ and weakly in $H^{1,2}$.
From (\ref{REFBEL}) and (\ref{Z}) we deduce that $\|\omega_{i}-\omega_{\infty}\|_{2}\rightarrow 0$,
which together with (\ref{Z}) again shows that $\omega_{i}\rightarrow \omega_{\infty}$ in $H^{1,2}$.
The convergence $u_{i}\rightarrow u_{\infty}$ in $H^{1,2}$ is shown in the same fashion. \end{proof}
\begin{proof}[\bf Proof of Proposition \ref{Limit}.]
Inequality (\ref{eqprop5.3a}) follows from Propositions \ref{LOCMIN} and \ref{GLOBAMIN},
together with the relation \eqref{MM} between the functionals $\fm_{ab}$ and $\olfm_{ab}$, as they imply
that extreme KdS data $(\sigma_e,\omega_e)$ are the unique global minimisers of $\fm_{ab}$ among functions $(\sigma,\omega)$ having the same
boundary conditions as $(\sigma_e,\omega_e)$ at $\theta_a,\theta_b$.
The proof of \eqref{eqlimit} is line by line identical to the proof when $\Lambda=0$ and which was obtained in \cite{2011CQGra..28j5014A}.
We will only sketch the argument here and refer the reader to \cite{2011CQGra..28j5014A} for details. It is important to remark that the
presence of the cosmological constant plays no important role in this step.
Divide the interval $[0,\pi]$ in three regions, $\Omega_I=\{\sin\theta\leq e^{(\ln t)^2}\}$, $\Omega_{II}=\{e^{(\ln t)^2}\leq\sin\theta\leq t\}$
and $\Omega_{III}=\{t\leq\sin\theta\}$. Note that when $t$ goes to zero, the regions $\Omega_I$ and $\Omega_{II}$ shrink toward the poles,
while $\Omega_{III}$ extends to cover the whole interval $[0,\pi]$. Then a specific partition function $f(\theta)$ (see eqs. (70)-(71) in
\cite{2011CQGra..28j5014A})
is used to interpolate between extreme KdS horizon data
in region $\Omega_I$ and general data in region $\Omega_{III}$. Define the auxiliary interpolating data $\gamma(t)=(\sigma(t),\omega(t))$ as
\be
\gamma(t)=f_t(\sin\theta)\gamma+(1-f_t(\sin\theta))\gamma_e,
\ee
then, as mentioned before, combining Propositions \ref{LOCMIN} and \ref{GLOBAMIN} on the region $[\theta_a,\theta_b]:=\Omega_{II}\cup\Omega_{III}$ for functions $\gamma(t):\Gamma_{ab}\to \mathbb R^2$
we find
\be\label{bound4}
\fm_{ab}(\gamma(t))\geq\fm_{ab}(\gamma_e).
\ee
Moreover, as $\gamma(t)|_{\Omega_I}=\gamma_e|_{\Omega_I}$, we can extend \eqref{bound4} to $[0,\pi]$ (recall that
$[0,\pi]=\Omega_I\cup[\theta_a,\theta_b]$) to obtain
\be\label{boundep}
\fm(\gamma(t))\geq\fm(\gamma_e).
\ee
The final step is to show that as $t$ goes to zero, the mass functional for the auxiliary data converges to the
mass functional for the original general data, that is
\be\label{boundm}
\lim_{t\to0}\fm(\gamma(t))=\fm(\gamma).
\ee
This is done in an identical manner as in \cite{2011CQGra..28j5014A} (with $\Lambda$ being irrelevant here),
by using that $\omega=\omega_e+\mathcal O(\sin^2\theta)$ near the poles and that
$\fm(\gamma)$ and $\fm(\gamma_e)$ are
well defined.
Inequalities \eqref{boundep} and \eqref{boundm} give \eqref{eqlimit}.
Moreover, using the explicit value
\be
\eexp^{\ \dfrac{\fm(\sigma_e,\omega_e,\hat{A},\hat{a})-\hat{\beta}}{8\hat\kappa}}=\frac{\hat A}{4\pi}
\ee
we find
\be
\eexp^{\ \dfrac{\fm(\sigma,\omega,\hat{A},\hat{a})-\hat{\beta}}{8\hat\kappa}}\geq \frac{\hat A}{4\pi}
\ee
which is inequality \eqref{ineq4}.
\end{proof}
\section{Possible generalisations}
\label{sec:disc}
We conclude discussing possible extensions of our main result to the case
with electromagnetic field and to the case $\Lambda < 0$.
In the former case we conjecture an inequality which, in addition to $A$, $J$ and $\Lambda$,
contains electric and magnetic charges $Q_E$ and $Q_M$ in the combination $Q^2=Q_E^2+Q_M^2$.
Such an extension is natural from the fact that all special cases are proven,
in particular we recall \cite{Clement:2012vb} the bound $A^2 \ge 16\pi^2 ( 4J^2 + Q^4)$ in the case $\Lambda
=0$. Moreover, extreme Kerr-Newman-deSitter saturates (\ref{J<}) and (\ref{Q<}).
\begin{Conjecture}
Under the assumptions of Theorem \ref{main} but under the
presence of an electromagnetic field with charges $Q_E$, $Q_M$ with
$Q^2=Q_E^2+Q_M^2$ and for any $\Lambda > 0$ we have
\begin{equation}
\label{J<}
J^2 \le \frac{A^2}{64\pi^2} \left[
\left( 1 - \frac{\Lambda A}{4\pi} \right) \left( 1 - \frac{\Lambda A}{12\pi} \right)
- \frac{2\Lambda Q^2}{3} \right] - \frac{Q^4}{4}
\end{equation}
or equivalently,
\begin{equation}
\label{Q<}
\left(Q^2 + \frac{\Lambda A^2}{48\pi^2} -
\sqrt{\frac{A^2}{16\pi^2} \left(1 - \frac{\Lambda A}{6\pi} \right)^2 - 4 J^2 } \right)
\left(Q^2 + \frac{\Lambda A^2}{48\pi^2} +
\sqrt{\frac{A^2}{16\pi^2} \left(1 - \frac{\Lambda A}{6\pi} \right)^2 - 4 J^2 } \right)
\le 0
\end{equation}
Moreover, (\ref{J<}) and (\ref{Q<})
are saturated precisely for
extreme Kerr-Newman-deSitter configurations.
\end{Conjecture}
As to the calculations leading to (\ref{J<}) and
(\ref{Q<}) we made use of Equ. (44) of Caldarelli et al. \cite{Caldarelli:1999xj},
where the temperature $T$ of a Kerr-Newman-anti-deSitter black hole is given in terms
of $l^2 = -3/\Lambda$, the mass $M$,
the entropy $S = A/4$, $Q$ and $J$. This calculation is insensitive to the
sign of $\Lambda$, and the requirement that $T \ge 0$ gives directly (\ref{J<}),
while (\ref{Q<}) is obtained via simple algebraic manipulations.
\vs
We finally comment on the prospects of proving the area inequalities
(\ref{A1}), (\ref{J<}) and (\ref{Q<})
for the case $\Lambda < 0$ along the lines described above.
We first remark that extreme Kerr-anti-deSitter saturates (\ref{A1}) which should be
clear from the discussion of Sect. 3, and extreme Kerr-Newman-anti-deSitter saturates
(\ref{J<}) and (\ref{Q<}).
Next, the first part of our proof of (\ref{A1}), namely the
lower bound for $A$ in terms on ${\cal M}$ as given in (\ref{ineq1})
carries over to $\Lambda < 0$ straightforwardly.
However, attempts of obtaining a lower bound for ${\cal M}$ analogously to
(\ref{ineq2}) seem to be in vain. The reason is that one can easily construct examples with sufficiently
small $\sigma$, (negative with large modulus), and suitably adjusted $\omega$
for which the last term in (\ref{defM}),
which is now negative, dominates the first two positive terms.
In fact these examples strongly suggest that ${\cal M}$ is even unbounded from
below unless the data are restricted appropriately.
Therefore, while it is still possible that (\ref{A1}), (\ref{J<}) and (\ref{Q<})
hold for $\Lambda < 0$ as well, our strategy which was successful for $\Lambda > 0$ is unlikely to carry over.
\bigskip
{\bf Acknowledgements.} We acknowledge helpful discussions with
Lars Andersson, Piotr Bizo\'n, Piotr Chru\'sciel, Sergio Dain, Jose Luis
Jaramillo, Marc Mars and Luc Nguyen.
M.E.G.C. is supported by CONICET (Argentina). W.S. was funded by the Austrian
Science Fund (FWF): P23337-N16 and by the Albert Einstein Institute
(Potsdam).
\bibliographystyle{plain}
|
\section{Introduction}
Magnetically shielded rooms (MSRs) are environments where external electromagnetic field distortions at low frequencies are strongly damped and the earth or ambient magnetic fields are strongly suppressed. MSRs are typically used in biomagnetic~\cite{trahms}, medical and fundamental physics applications~\cite{baker, tum_edm}. For these, both a strong suppression of time varying external fields of frequencies from mHz to kHz, as well as the resulting residual quasi static magnetic field ($<$~1~nT) and residual magnetic field gradients ($<$~1~nT/m) are important.
Shielding in the low frequency range of up to tens of Hz is usually achieved by using large amounts of highly permeable alloys (e.g. Permalloy). However, already at the 1~Hz regime the conductivity of the shielding material plays a significant role due to induced currents over large areas of the surface. For shielding of radio frequency (RF) disturbances, an additional layer of highly conducting material is added. The reference facility for magnetic shielding is the Berlin Magnetically Shielded Room 2 (BMSR-2) at the Physikalisch-Technische Bundesanstalt (PTB) Berlin~\cite{bmsr2}. With 7 layers of Mumetall\footnote{Mumetall is a brand name of Vacuumschmelze Hanau GmbH} and an additional aluminum layer for RF shielding it has a passive shielding factor (SF) of 75000 at 0.01~Hz. The new shield for the next-generation neutron electric dipole moment (nEDM) experiment at the Technische Universität München (TUM) reaches even higher SFs exceeding 1000000 at 1~mHz, without any active measures to compensate for low-frequency drifts~\cite{insert}.
Materials used for shielding have a high permeability and a non-zero remanence and therefore can be easily magnetized by arbitrary fields. Such fields are typically generated by (i) static external sources, (ii) internally applied fields for e.g.~in low-field NMR applications, (iii) magnetic fields emerging from adjacent shielding layers and (iv) magnetic distortions, e.g.~from construction materials of the shield itself. To achieve lowest possible residual fields and field gradients, the material has to be 'degaussed' repeatedly. 'Degaussing' is the commonly used term for the reduction of the remanent magnetization in magnetizable materials. It is usually performed by applying a sinusoidal current with decreasing amplitude to coils wound around the shielding material inducing magnetic flux in a closed loop of shielding material. The starting amplitude has to be large enough to reach saturation of the magnetization everywhere within the material. At the end of the cycle, the amlitude has to reach zero as exactly as possible. Any DC offset or noise can worsen the quality of the degaussing procedure. Degaussing of MSRs has been developed to an unprecedented quality~\cite{voigt, tum_msr}, with achieved residual fields of $<$~500~pT and gradients of $<$~0.2~nT/m over the center cubic meter volume inside walk-in MSRs. These small fields can be generated in a highly reproducible manner to within 10’s of pT, independent of the initial conditions.
The relationship between residual field and shielding has also been well investigated~\cite{tum_msr}. The shielding factor of an MSR is given as the frequency dependent attenuation of external disturbances
\begin{equation}
SF(\omega) = \frac{A_o \times \sin(\omega t)} {A_i \times \sin(\omega t + \phi)}.
\end{equation}
$A_o$ denotes the field magnitude at a position measured before the presence of the MSR, $A_i$ the magnitude at the same position with the MSR, $\omega$ is the frequency of the external disturbance and $\phi$ is a phase shift between the excitation field and the measured field inside the MSR. Assuming a \emph{sufficient} amount of shielding material, in the static case ($\omega=0$) the field within the MSR is dominated rather by the residual field from the shielding material, and \emph{not} by the effect of external fields.
Ultra-low residual fields are achieved by effectively offering the magnetic field a path of minimal magnetic resistance around the volume to be shielded. By minimizing the potential energy of the magnetic domains within the shielding material in their surrounding field an equilibrium state is formed, which is stable in time for a static environment. Therefore degaussing does not per se remove the magnetic field as the name suggests, but rather optimizes the magnetic field in the material with respect to the external condition. For these reasons, we prefer the usage of 'magnetic equilibration' instead of degaussing to refer to this process (naturally, spots of strongly magnetized material still have to be degaussed to reduce their magnetization).
One of the procedures to minimize residual magnetic fields inside MSRs is described in Ref.~\onlinecite{voigt}. Here, the coil setup for equilibration consists of 12 coils for each layer of the MSR with one along each edge of the cube. Four of these coils are connected in series to form a closed loop around one spatial direction (see also Fig~\ref{fig:coils}a). The three spatial directions are then degaussed sequentially. A sinusoidal current with a peak amplitude of 21$\times$5~Ampere$\times$turns and a frequency of 10~Hz is applied. The amplitude decreases over 2000 periods (‘cycles’) to a value as close to zero as technically achievable. Equilibration therefore takes 600~s for each layer. With this procedure residual fields below 1.5~nT and gradients smaller than 2~nT$/$m have been achieved.
In Ref.~\onlinecite{thiel} one of the main criteria for a reproducible procedure is the step size of the amplitude decrease. The difference in amplitude between two subsequent maxima of the sinusoidal current should be as small as possible as this delta corresponds to an error in the residual field. Therefore it is suggested to use several thousand cycles.\\
\indent In this work we describe a new procedure for magnetic equilibration, which is capable of producing extremely small residual fields, comparable to the best ever achieved. However, the result is achieved within a much shorter time constant of only 30-50~s per layer, with simpler wiring of equilibration coils and with simpler equipment. In the previous procedure with a duration of about 900~s, residual fields of $<$~0.7~nT and field gradients of 0.3~nT/m were achieved in the MSR at TU München~\cite{tum_msr}. The new configuration can produce the same or better values in only 150~s. Here, both layers of the MSR are equilibrated in a defined sequence.
For MSRs with only one layer of magnetizable material, such a sequence can even be shorter, on the order of only 10~s with a resulting residual field of $<$10~nT over a cubic meter inside, making single layer shielding a highly competitive technical solution at low cost.
Our finding has consequences for several applications, including in fundamental physics where the duty cycle of experiments is affected by frequent magnetic equilibration sequences (e.g.~required due to frequent reversal of applied NMR fields inside). Another example would be applications in biomagnetism, where the door to an MSR may be opened frequently.
\section{Apparatus}
\begin{figure}
\includegraphics[width=0.3\textwidth]{Fig1.eps}
\caption{The setup of the equilibration coils for the old configuration, shown in a), and the new configuration in b). In a) the 4 coils to create a magnetic flux around the Z direction and their connections are shown. The new configuration in b) has the old Y and Z direction connected in a L shape so that a magnetic flux is created in both directions simultaneously.}
\label{fig:coils}
\end{figure}
\begin{figure}
\includegraphics[width=0.35\textwidth]{Fig2.eps}
\caption{Allan deviation~\cite{allan} calculated for data recorded with the fluxgate inside the MSR for about 20000~s. The inset shows the corresponding time data, shifted to 0 at $t~=~0$ on the left axis. Bx and Bz have then been offset by 0.4 nT and -0.4 nT respectively for clarity. The smooth change of the magnetic field over time correlates with a smooth, monotonic drop of $0.2^\circ$~C in the environmental temperature. To attribute the correlation of field and temperature to a change of permeability is however more complex.}
\label{fig:fg}
\end{figure}
The MSR used for this measurement and also the previous configuration of equilibration coils are described in Ref.~\onlinecite{tum_msr} (see also Fig \ref{fig:coils}a). The two layers of the MSR are equilibrated individually, first the inner layer, then the outer one and after that the inner one again. For each layer, we start with the X direction followed by the Z direction and then the Y direction (see also Fig \ref{fig:coils}). It was found that equilibration of one direction is not sufficient: subsequent equilibrations in the second and third direction each improved the result. Note the door is in the XZ-plane at negative y coordinate.
Each equilibration sequence starts with an amplitude that saturates the magnetizable material, in the case of the inner layer that is $9 \times 7$ Ampere turns, and then linearly decreases over 1000~cycles to below noise level. A frequency of 10~Hz is used. The waveform for equilibration is generated in software on a PC, converted to an analog voltage by a 16~bit digital-analog converter (DAC) with a sampling frequency of 20~kHz. This voltage is passed through an analog voltage divider to use the full range of the DAC and a 100~Hz lowpass filter to smooth out DAC steps. It is converted to a current in a power amplifier\footnote{Custom power amplifier model 'Rohrer PA 2088B' provided by Rohrer GmbH Munich} and passed through a transformer to remove any DC offset before being distributed to the respective coil of the MSR by an automated relay switch. In this way magnetic equilibration can be done completely automated, in a process that takes about 15~min.
Field maps have been measured using a three-axis fluxgate sensor\footnote{Bartington Mag03-IEHV70}. The sensor heads at the end of flying leads are mounted in a nonmagnetic structure so that they form a right handed coordinate system. Their signal is read out with a 16-bit analog digital converter (ADC). Each data point in the maps is an average over a measuring time of 1~s, where data has been recorded with a sampling frequency of 1~kHz. Several measurements taken at the same point deviate less than 0.1~nT from each other. To correct for drifts of fluxgate offsets, a calibration is done before, during and after each field map.
The sensor is calibrated in the center of the room where the fields are smallest, to keep the influence of alignment errors as small as possible. To determine the offset of a sensor the field is measured at a position and then again at the same position with the sensor rotated by $180^{\circ}$ to measure in the opposite direction. Assuming a linear drift between calibrations the offsets drift rates were $<$~10~pT/min. All maps were corrected for these offset drifts.
Figure \ref{fig:fg} shows the performance of the fluxgate magnetometer using an Allan variance plot. To obtain the optimal performance, an integration time of about 100~s would be required. For shorter times, noise will have an increasing effect on the measured signals. For longer times the measurement will be dominated by drifts that could come from changing offsets or drifts of the readout electronics. To keep the measurement time per point reasonably short, an integration time of 1~s was chosen. Here the Allan deviation is already below 30~pT. With this and repeated offset calibration an overall accuracy of $\sim$~0.3~nT can be reached, so that low-noise fluxgate sensors are well suitable for this kind of measurements.
After magnetic equilibration in the earth field without active stabilization, field maps were recorded for the respective configuration. Each map consists of 108 points on a 6~by~6~grid in 3 different XY planes (for coordinate system see Fig.~\ref{fig:coils}) in the center of the MSR, 35~cm below and 35~cm above it. The distance between the grid points along each axis is 10~cm. The sensor was moved by hand from one point to the next with an estimated accuracy of about 0.5~cm and a tilt-precision of better than few degrees. The distance of the individual sensors in the mount is 2~cm, so the overall position accuracy of the measured points is about 2~cm.
\section{New coil configuration and measurements}
\begin{figure}
\includegraphics[width=0.5\textwidth]{Fig3.eps}
\caption{Maps of the magnitude of the magnetic flux density inside the MSR after magnetic equilibration in three XY planes at different heights. The points are 10~cm apart in each direction. The columns show: the previous coil configuration (a), and the L shaped configuration ( (b) and (c) ), where the later were taken with several days between the two measurements.}
\label{fig:maps}
\end{figure}
The equilibration scheme we describe here uses a different coil configuration: two coils sets of the previous configuration are reconnected so that the coils form an ‘L-shaped’ coil pattern along two edges of the MSR. Four of these Ls are again connected in series (see Fig. \ref{fig:coils} b). All connections between the coils are routed in a way that the fields produced by the connecting wires cancel. In this way, two spatial directions are equilibrated at the same time. The results show that the third direction can be omitted. Only the inner layer of the MSR was modified in this way. The equilibration sequence starts with the inner layer, followed by the unmodified outer layer and inner layer in sequence, using 50~s for the inner layer. This time can be further reduced, but below 30~s a notable decrease in the remanent field quality is observed.
Figure \ref{fig:maps} shows the recorded field maps. In column a) the residual field for the old configuration is shown. The maximum value is about 0.8 nT in the lower plane and near the door of the MSR. In the upper planes the field is about 0.5~nT. The higher values close to the door are likely caused by stray fields of the last equilibration step. It was previously~\cite{tum_msr} found that equilibrating the direction not including the door yielded the best result. However there stray field again magnetize the overlaps. Columns b) and c) show the residual field for the new configuration, measured twice with several days in between to show reproducibility. Here the residual field is slightly smaller, on the order of 0.5~nT or below. Also the prominent feature in the lower plane is not visible anymore. With only one equilibration cycle a re-magnetization of the adjacent wall is avoided. These are values that are achieved in typical daily operation, where non-magnetic furniture and tools, as well as people are inside the MSR.
\begin{figure}
\includegraphics[width=0.3\textwidth]{Fig4.eps}
\caption{Simulations of the magnetization in the adjacent face that is not part of the equilibration loop. The geometry of the room and placement of the coils is the same for both simulations and only shown in b). In a) a current is supplied to the four coils at the edges, in b) all eight coils carry a current. The color scale is the same for both plots and in arbitrary units. Both cases illustrate that the currents create a significant magnetization in the adjacent face.}
\label{fig:simcoils}
\end{figure}
To illustrate problems with the placement of equilibration coil a finite-elemente-method simulation with COMSOL is shown in Fig.~\ref{fig:simcoils}. In a) a current is applied to the four coils at the edges of the MSR, creating a flux loop around the four side walls of the MSR (around the Z direction). In Fig.~\ref{fig:simcoils}b) eight coils distributed equally along the path of flux are supplied with half the current used for a). Each coil has to penetrate the top wall of the MSR to form a loop with shielding material enclosed. The simulation shows the magnetization inside the top wall created by the current through the coils. Since this face is not included in the equilibration loop, its magnetization is not properly removed by the equilibration process. This dominates the error caused by a possible DC offset at the level of precision reached in our setup. Here a big number of equilibration cycles (as mentioned in Ref.~\onlinecite{thiel}) reduce this error. Additionally, when small volumes in the center of the MSR are considered, the error has a small influence on the residual field due to the symmetry of the MSR. However for extended-size volumes these contributions become more significant. Even when currents are better distributed and have a smaller amplitude the adjacent face retains a magnetization. By equilibrating all faces at once this problem does not occur.
\section{Conclusion}
The new scheme for equilibration coils presented here produces static remanent fields below 0.5~nT inside a two-layer MSR with less contributions from magnetizable features at the walls. This result is on the same order of magnitude as that from the previous equilibration setup, but achieved with significantly shorter cycle times. For the two-layered room at TUM, the time will be reduced from 900~s to 150~s when both layers are changed. In an MSR with more layers, the reduction can be even larger. For the outer layers perfect equilibration is not necessary and the time can be shorter than the 50~s used here. If for example the outer three layers of a 5 layer MSR are equilibrated for 20~s, and the inner two for 50~s, the whole procedure would take only 160~s instead of 1500~s. In the nEDM experiment at TUM the shield has to be equilibrated for example every 10~min. With the reduced time for the equilibration, the duty cycle of the experiment is effectively increased by a factor of 2, resulting in an improved sensitivity of the measurement by about 40\%.
In a single layer shield, where only remanent fields of tens of nT are targeted, equilibration can be done within about 10~s. Additionally, this setup only requires one set of coils per layer. This significantly reduces the amount of high current relays needed for automated switching between the equilibration coils.
Further, the simultaneous equilibration of all faces avoids accidental magnetization in adjacent faces due to stray fields.
\begin{acknowledgments}
We acknowledge the support at the FRM-II research reactor in Garching. This work was supported by DFG Priority Program SPP 1491 and the DFG Cluster of Excellence 'Origin and Structure of the Universe'.
\end{acknowledgments}
|
\section{Introduction}
One of the simplest extensions of the Standard Model (SM) consists of the
addition of scalar doublets to the SM spectrum. Multi-Higgs extensions arise
in a variety of frameworks, including supersymmetric extensions of the SM,
as well as models with family symmetries. A two Higgs doublet model (2HDM) was
first introduced by Lee \cite{Lee:1973iz}, in order to achieve spontaneous
breaking of the CP symmetry. If no extra symmetries are introduced, 2HDMs
lead to too large tree level scalar mediated
flavour-changing-neutral-currents (FCNC) \cite{Branco:2011iw}.
In order to avoid these
potentially dangerous currents, various schemes have been proposed: \newline
i) Glashow and Weinberg \cite{Glashow:1976nt} have pointed out that one can
avoid FCNC at tree level by introducing a $Z_2$ symmetry under which the two
Higgs doublets transform differently. The introduction of a $Z_2$
symmetry in 2HDMs prevents the generation of spontaneous
CP breaking \cite{Branco:1980sz}
unless the symmetry is softly broken \cite{Branco:1985aq}. \\
ii) Pich and Tuzon \cite{Pich:2009sp} have conjectured the existence of flavour
alignment of the two Yukawa matrices, thus avoiding FCNC at tree level. This
is an interesting suggestion, but it has the drawback of being an ad-hoc
assumption, not explained by any symmetry. Furthermore, it has been pointed out
that in general this scheme is not stable under the
renormalization group \cite{Ferreira:2010xe}. There have been attempts at obtaining
alignment in various extensions of the SM \cite{Serodio:2011hg}, \cite{Varzielas:2011jr},
\cite{Celis:2014zaa}.\\
iii) Another possibility has been proposed some time ago \cite{Branco:1996bq}
by Branco, Grimus and Lavoura (BGL) who have pointed out that there is a
symmetry which, when imposed on the Lagrangian, constrains the Yukawa
couplings in such a way that FCNC do arise at tree level, but are entirely
determined by the $V_{CKM}$ matrix, with no other free parameters.
In some of the
BGL models, one has a strong natural suppression of the most dangerous FCNC,
with, for example, the strangeness changing neutral currents,
proportional to $(V_{td} V_{ts} ^*)^2$, which implies a very strong natural
suppression of the contribution to the $K^0 - \bar{K^0}$ transition. With this
suppression, the neutral Higgs masses need not be too large. BGL models have
been extended to the leptonic sector \cite{Botella:2011ne}, their relation
to Minimal Flavour Violation models has been studied \cite{Botella:2009pq}
and their phenomenological implications have been recently analysed
\cite{Botella:2014ska}. \newline
In this paper, we reexamine the question of the stability of flavour alignment
under the renormalization group. Assuming that the Yukawa couplings of the
two Higgs doublets are aligned, i.e., proportional to each other, we study under
what conditions the alignment is maintained by the renormalization group.
Apart from the conditions already found in Ref.\cite{Ferreira:2010xe}, we find new
solutions which can be of great physical interest. One of these solutions,
corresponds to having all the Yukawa coupling matrices proportional to the
so-called democratic matrix \cite{Branco:1990fj}, \cite{Fritzsch:1994yj},
\cite{Branco:1995pw}.
This solution is rather unique, since on the one
hand it is stable under the renormalization group equations
(RGE) and on the other hand, it is the only stable solution
which provides a good starting point for reproducing the observed pattern of quark
masses and mixing.
We then point out that this flavour democratic solution
can be obtained as a result of a $Z_3 \times Z^\prime_3$ flavour symmetry.
In the framework that we propose, flavour alignment is exact in the limit
where only the third family acquires mass. Once the two light generations acquire
a mass, there are small deviations from alignment, which are suppressed by
the strong hierarchy of quark masses. As a result, one
obtains in this framework, a quasi-alignment of the Yukawa couplings, as a
result of the $Z_3 \times Z^\prime_3$ symmetry, together with the strong
hierarchy of quark masses. \newline
The paper is organised as follows. In the next section, we briefly describe
the general flavour structure of the 2HDM, in order to settle our notation.
In section 3 we derive all the solutions for the Yukawa couplings, leading
to alignment, stable under the renormalization group. In section 4 we show
that the flavour democratic solution can be obtained as a result of a
$Z_3 \times Z^\prime_3$ flavour symmetry and propose an
ansatz for the breaking of the $Z_3 \times Z^\prime_3$ symmetry.
In section 5, we examine the suppression of scalar mediated FCNC
in our framework. In section 6, we perform a numerical analysis,
showing how the pattern of quark masses and mixing can be obtained in
the framework of our ansatz. Finally our conclusions are contained
in section 7. In the Appendix we present a full study of the solutions of
the alignment conditions.
\section{Yukawa Couplings in the General Two-Higgs-Doublet-Model (2HDM)}
For completeness and in order to establish our notation we
briefly review the flavour structure of the 2HDM, when no extra symmetries
are introduced in the Lagrangian. The Yukawa couplings can be written:
\begin{equation}
\mathcal{L}_{Y}=-\overline{Q_{L}^{0}}\ \Gamma _{1}\Phi _{1}d_{R}^{0}-
\overline{Q_{L}^{0}}\ \Gamma _{2}\Phi _{2}d_{R}^{0}-\overline{Q_{L}^{0}}\
\Omega _{1}\tilde{\Phi}_{1}u_{R}^{0}-\overline{Q_{L}^{0}}\ \Omega _{2}\tilde{
\Phi}_{2}u_{R}^{0} -\overline{L_{L}^{0}}\ \Pi _{1}\Phi _{1}l_{R}^{0}-\overline{L_{L}^{0}}\
\Pi _{2}\Phi _{2}l_{R}^{0} + \text{h.c.}\
\label{Yuk}
\end{equation}
where $\Phi _{i}$ denote the Higgs doublets and $\widetilde{\Phi }_{i}\equiv
i\,\tau _{2}\,\Phi _{i}^{\ast }$ and $\Gamma _{i}$, $\Omega _{i}$, $\Pi_{i}$ are
matrices in flavour space. After spontaneous symmetry breaking, the
following quark mass matrices are generated:
\begin{equation}
M_{d}=\frac{1}{\sqrt{2}}(v_{1}\Gamma _{1}+v_{2}e^{i\alpha }\Gamma _{2})\
,\qquad M_{u}=\frac{1}{\sqrt{2}}(v_{1}\Omega _{1}+v_{2}e^{-i\alpha }\Omega
_{2}) \label{mumd}
\end{equation}
where $v_{i}/\sqrt{2}\equiv |<0|\phi _{i}^{0}|0>|$ and $\alpha $ denotes the
relative phase of the two vacuum expectation values (vevs) of the neutral
components $\phi _{i}^{0}$ of $\Phi _{i}$. The neutral and the charged
Higgs interactions with quarks are of the form:
\begin{eqnarray}
{\mathcal{L}}_{Y}(\mbox{quark, Higgs}) &=&-\overline{d_{L}^{0}}\,\frac{1}{v}%
\,[M_{d}\,H^{0}+N_{d}^{0}\,R+i\,N_{d}^{0}\,I]\,d_{R}^{0}+ \notag \\
&-&\overline{{u}_{L}^{0}}\,\frac{1}{v}\,[M_{u}\,H^{0}+N_{u}^{0}\,R+i%
\,N_{u}^{0}\,I]\,u_{R}^{0}+ \label{rep} \\
&&+\frac{\sqrt{2}H^{+}}{v}(\overline{{u}_{L}^{0}}N_{d}^{0}\,d_{R}^{0}-%
\overline{{u}_{R}^{0}}{N_{u}^{0}}^{\dagger }\,d_{L}^{0})+\text{h.c.} \notag
\end{eqnarray}%
where $v\equiv \sqrt{v_{1}^{2}+v_{2}^{2}}\approx 246\ GeV$ and $H^{0}$, $R$
are orthogonal combinations of the fields $\rho _{j}$, arising when one
expands \cite{Lee:1973iz} the neutral scalar fields around their vevs, $\phi
_{j}^{0}=\frac{e^{i\alpha _{j}}}{\sqrt{2}}(v_{j}+\rho _{j}+i\eta _{j})$,
choosing $H^{0}$ in such a way that it has couplings to the quarks which are
proportional to the mass matrices, as can be seen from Eq.~(\ref{rep});
similarly, $I$ denotes the linear combination of $\eta _{j}$ orthogonal to
the neutral Goldstone boson. The matrices $N_{d}^{0}$ and $N_{u}^{0}$ are
given by:
\begin{equation}
N_{d}^{0}=\frac{1}{\sqrt{2}}(v_{2}\Gamma _{1}-v_{1}e^{i\alpha }\Gamma
_{2}),\quad N_{u}^{0}=\frac{1}{\sqrt{2}}(v_{2}\Omega _{1}-v_{1}e^{-i\alpha
}\Omega _{2}) \label{ndnu}
\end{equation}
The quark mass matrices are diagonalized through
\begin{equation}
\begin{array}{c}
U_{dL}^{\dagger }\,M_{d}\,U_{dR}=D_{d}\equiv \mbox{diag}(m_{d},m_{s},m_{b}),
\\
\\
U_{uL}^{\dagger }\,M_{u}\,U_{uR}=D_{u}\equiv \mbox{diag}(m_{u},m_{c},m_{t}).
\end{array}
\label{mdmu}
\end{equation}
and the matrices $N_{d}^{0}$ and $N_{u}^{0}$ in the mass eigenstate basis
transform into:
\begin{equation}
U_{dL}^{\dagger }\,N_{d}^{0}\,U_{dR}=N_{d},\qquad U_{uL}^{\dagger
}\,N_{u}^{0}\,U_{uR}=N_{u}. \label{ndnu1}
\end{equation}
There are similar expressions for the leptonic sector. We do not introduce neutrino masses
since these are not relevant for our analysis.
\section{Stability of the aligned 2HDM under RGE}
The aligned two Higgs doublet model (A2HDM) is defined at tree level by the following relations involving the matrices introduced in Eq.~(\ref{Yuk}):
\begin{eqnarray}
\Gamma _{2} &=&d\cdot \Gamma _{1} \notag \\
\Omega _{2} &=&u\cdot \Omega _{1} \label{AlignementConditions} \\
\Pi _{2} &=&e\cdot \Pi _{1} \notag
\end{eqnarray}
where $d$, $u$, $e$ are constants.
In this section we analyse the stability of the A2HDM under renormalisation group
equations (RGE). The one loop renormalization group equations (RGE) for the Yukawa
couplings are \cite{Grimus:2004yh}, \cite{Ferreira:2010xe}:
\begin{eqnarray}
\mathcal{D}\Gamma _{k} &=&a_{\Gamma }\Gamma _{k}+ \nonumber \\
&&+\sum_{l=1}^{2}\left[ 3\text{Tr}\!\left( \Gamma _{k}\Gamma _{l}^{\dagger }+\Omega
_{k}^{\dagger }\Omega _{l}\right) +\text{Tr}\!\left( \Pi _{k}\Pi _{l}^{\dagger
}+\Sigma _{k}^{\dagger }\Sigma _{l}\right) \right] \Gamma _{l}+ \nonumber \\
&&+\sum_{l=1}^{2}\left( -2\Omega _{l}\Omega _{k}^{\dagger }\Gamma
_{l}+\Gamma _{k}\Gamma _{l}^{\dagger }\Gamma _{l}+\frac{1}{2}\Omega
_{l}\Omega _{l}^{\dagger }\Gamma _{k}+\frac{1}{2}\Gamma _{l}\Gamma
_{l}^{\dagger }\Gamma _{k}\right) \ , \label{RGE1}
\end{eqnarray}
\begin{eqnarray}
\mathcal{D}\Omega _{k} &=&a_{\Omega }\Omega _{k}+ \nonumber \\
&&+\sum_{l=1}^{2}\left[ 3\text{Tr}\!\left( \Omega _{k}\Omega _{l}^{\dagger }+\Gamma
_{k}^{\dagger }\Gamma _{l}\right) +\text{Tr}\!\left( \Sigma _{k}\Sigma _{l}^{\dagger
}+\Pi _{k}^{\dagger }\Pi _{l}\right) \right] \Omega _{l}+ \nonumber \\
&&+\sum_{l=1}^{2}\left( -2\Gamma _{l}\Gamma _{k}^{\dagger }\Omega
_{l}+\Omega _{k}\Omega _{l}^{\dagger }\Omega _{l}+\frac{1}{2}\Gamma
_{l}\Gamma _{l}^{\dagger }\Omega _{k}+\frac{1}{2}\Omega _{l}\Omega
_{l}^{\dagger }\Omega _{k}\right) \ , \label{RGE2}
\end{eqnarray}
\begin{eqnarray}
\mathcal{D}\Pi _{k} &=&a_{\Pi }\Pi _{k}+ \nonumber \\
&&+\sum_{l=1}^{2}\left[ 3\text{Tr}\!\left( \Gamma _{k}\Gamma _{l}^{\dagger }+\Omega
_{k}^{\dagger }\Omega _{l}\right) +\text{Tr}\!\left( \Pi _{k}\Pi _{l}^{\dagger
}+\Sigma _{k}^{\dagger }\Sigma _{l}\right) \right] \Pi _{l}+ \nonumber \\
&&+\sum_{l=1}^{2}\left( -2\Sigma _{l}\Sigma _{k}^{\dagger }\Pi _{l}+\Pi
_{k}\Pi _{l}^{\dagger }\Pi _{l}+\frac{1}{2}\Sigma _{l}\Sigma _{l}^{\dagger
}\Pi _{k}+\frac{1}{2}\Pi _{l}\Pi _{l}^{\dagger }\Pi _{k}\right) \ ,
\label{RGE3}
\end{eqnarray}
where $\mathcal{D}\equiv 16\pi ^{2}\mu \left( d/d\mu \right) $ and $\mu $ is
the renormalization scale. The coefficients
$a_{\Gamma }$, $a_{\Omega }$ and $a_{\Pi }$ are given by:
\begin{eqnarray}
a_{\Gamma }= -8 g_s^2 -\frac{9}{4} g^2 -\frac{5}{12} {g^\prime}^2 \\
a_{\Omega }= -8 g_s^2 -\frac{9}{4} g^2 -\frac{17}{12} {g^\prime}^2 \\
a_{\Pi } = -\frac{9}{4} g^2 -\frac{15}{4} {g^\prime}^2
\end{eqnarray}
where $g_s$, $g$ and $g^\prime$ are the gauge coupling constants of
$SU(3)_c$, $SU(2)_L$ and $U(1)_Y$ respectively. The alignment relations
given by Eq.~(\ref{AlignementConditions}) guarantee the absence of Higgs
mediated FCNC at tree level because both matrices $M_{d}$ and $N_{d}$
are proportional to $\Gamma _{1}$. Similarly both
$M_{u}$ and $N_{u}$ are proportional to $\Omega _{1}$ and
$M_{l}$, $N_{l}$ to $\Pi_{1}$. In general, these relations are
broken at one loop level. From Eqs.~(\ref{RGE1}),(\ref{RGE2}),(\ref{RGE3}) one can easily derive:
\begin{eqnarray}
\mathcal{D}\left( \Gamma _{2}\right) -d\cdot \mathcal{D}\left( \Gamma
_{1}\right) &=&\left( u^{\ast }-d\right) \left( 1+ud\right) \left\{
3Tr\left( \Omega _{1}^{\dagger }\Omega _{1}\right) -2\Omega _{1}\Omega
_{1}^{\dagger }\right\} \Gamma _{1}+ \notag \\
&&+\left( e-d\right) \left( 1+e^{\ast }d\right) Tr\left( \Pi _{1}^{\dagger
}\Pi _{1}\right) \Gamma _{1} \label{AlignmentBrokenDown}
\end{eqnarray}
\begin{eqnarray}
\mathcal{D}\left( \Omega _{2}\right) -u\cdot \mathcal{D}\left( \Omega
_{1}\right) &=&\left( d^{\ast }-u\right) \left( 1+ud\right) \left\{
3Tr\left( \Gamma _{1}^{\dagger }\Gamma _{1}\right) -2\Gamma _{1}\Gamma
_{1}^{\dagger }\right\} \Omega _{1}+ \notag \\
&&+\left( e^{\ast }-u\right) \left( 1+eu\right) Tr\left( \Pi _{1}^{\dagger
}\Pi _{1}\right) \Omega _{1} \label{AlignmentBrokenUp}
\end{eqnarray}
\begin{eqnarray}
\mathcal{D}\left( \Pi _{2}\right) -e\cdot \mathcal{D}\left( \Pi _{1}\right)
&=& 3\left( d-e\right) \left( 1+d^{\ast }e\right) Tr\left( \Gamma _{1}^{\dagger
}\Gamma _{1}\right) \Pi _{1} + \notag \\
&&+ 3\left( u^{\ast }-e\right) \left( 1+eu\right) Tr\left( \Omega _{1}^{\dagger
}\Omega _{1}\right)
\Pi _{1} \label{AlignmentBrokenLepton}
\end{eqnarray}
In order to enforce Eq.~(\ref{AlignementConditions}) at one loop level it is
easy to realize that it is sufficient to impose:
\begin{eqnarray}
\mathcal{D}\left( \Gamma _{2}\right) -d\cdot \mathcal{D}\left( \Gamma
_{1}\right) &\propto &\Gamma _{1} \label{RGEstableDown} \\
\mathcal{D}\left( \Omega _{2}\right) -u\cdot \mathcal{D}\left( \Omega
_{1}\right) &\propto &\Omega _{1} \label{RGEstableUp} \\
\mathcal{D}\left( \Pi _{2}\right) -e\cdot \mathcal{D}\left( \Pi _{1}\right)
&\propto &\Pi _{1} \label{RGEstableLepton}
\end{eqnarray}
in fact the proportionality constants on the r.h.s. are the running \footnote
{The autors of reference \cite{Ferreira:2010xe} impose the r.h.s. \ of Eqs.~(\ref
{AlignmentBrokenDown}),(\ref{AlignmentBrokenUp},(\ref{AlignmentBrokenLepton})
to be equal to zero. This amounts to imposing alignment at one loop level and
imposing additionally that there is no running of the parameters $u,d$ and $e$.}
of $d,u$ and $e $.
Therefore Eq.~(\ref{AlignmentBrokenLepton}) does not impose any constraint: at
one loop level the charged lepton sector remains aligned and there are no
FCNC in the leptonic sector. This result agrees with the findings of
references \cite{Braeuninger:2010td} and \cite{Jung:2010ik}.
In equations (\ref{AlignmentBrokenDown}), and (\ref{AlignmentBrokenUp}) the pieces
that can break the alignment in the quark sector are the terms: $\Omega
_{1}\Omega _{1}^{\dagger }\Gamma _{1}$ and $\Gamma _{1}\Gamma _{1}^{\dagger
}\Omega _{1}$ respectively$\footnote
{It can be readily seen that $\Omega _{1}\Omega _{1}^{\dagger }\Gamma
_{1}\propto M_{u}M_{u}^{\dag }M_{d}$ and $\Gamma _{1}\Gamma _{1}^{\dagger
}\Omega _{1}\propto M_{d}M_{d}^{\dag }M_{u}$. It is worth emphasizing
that these
structures are precisely the ones obtained in \cite{Braeuninger:2010td} and in
\cite{Jung:2010ik},
which produce FCNC at one loop level.}$. In order to have
alignment at one loop level - fulfilling Eqs(\ref{RGEstableDown}), (\ref
{RGEstableUp})- there are two types of solutions:
\begin{enumerate}
\item $\left( u^{\ast }-d\right) \left( 1+ud\right) =0$
\item $\Omega _{1}\Omega _{1}^{\dagger }\Gamma _{1}=\lambda _{\Gamma }\Gamma
_{1}$ and $\Gamma _{1}\Gamma _{1}^{\dagger }\Omega _{1}=\lambda _{\Omega
}\Omega _{1}$. With $\lambda _{\Gamma }$ and $\lambda _{\Omega }$ complex
numbers.
\end{enumerate}
Solutions of type 1 include the usual 2HDM with natural flavour conservation (NFC),
where the up and down quarks receive contributions from only one Higgs doublet. It
is well known that this can be achieved through the introduction of a $Z_{2}$ symmetry.
Here, we are not interested in this class of well known solutions. We are
interested in the class of solutions of type 2, and in the Appendix, we have studied
the complete set of matrices $\Omega _{1}$ and $\Gamma _{1}$ that
obey to the conditions required for this class of solutions. We have shown in
the Appendix, that if one requires stability under the RGE and at the same time
Yukawa structures which are, in leading order, in agreement with the observed
pattern of quark masses and mixing, then one is lead to a unique solution,
where the matrices $\Omega _{1}$ and $\Gamma _{1}$ are of the form:
\begin{equation}
\begin{array}{ccc}
\Omega _{1}=c_{1}^{d}\Delta & ; & \Gamma _{1}=c_{1}^{u}\Delta
\label{equ20}
\end{array}
\end{equation}
with $\Delta $ the democratic mass matrix:
\begin{equation}
\Delta =\left(
\begin{array}{ccc}
1 & 1 & 1 \\
1 & 1 & 1 \\
1 & 1 & 1
\end{array}
\right)
\end{equation}
This solution corresponds to the limit where only the top and bottom quarks acquire
mass, while the two first generations are massless. The up and down quarks are
aligned in flavour space, so the $V_{CKM}$ matrix equals the identity. The other
stable solutions of type 2 correspond to non realistic cases like for example
having all up or down quarks massless or two up or two down quark masses
degenerate or with a $V_{CKM}$ very far from the identity matrix.
It is remarkable to realise that the so called democratic mass matrix is
stable under RGE and that precisely this stability also enforces what could
be called "quark alignment" in the sense that we also have a proportionality
among $\Gamma _{i}$ and $\Omega _{i}$.
\section{Natural Quasi-Alignment of Yukawa couplings}
In this section we search for the minimal symmetry which when imposed on the
Lagrangian, leads to the stable solution described in the previous section,
corresponding to the democratic Yukawa couplings of Eq.~(\ref{equ20}). Before
describing this symmetry, it is worth analysing another type of alignment
which is verified experimentally, the so-called up-down alignment in the
quark sector.
\subsection{The up-down alignment in the quark sector}
In the quark sector, flavour mixing is small. This means that there is a
weak basis (WB) where both $M_u$, and $M_d$ are close to the diagonal form.
Experiment indicates that not only flavour mixing is small, but there is
also up-down flavour alignment in the quark sector in the following sense.
We can choose, without loss of generality, a WB where $M_u = \mbox{diag}
(m_u, m_c, m_t)$. Of course, this is just a choice of ordering, with no
physical meaning. Small mixing implies that in this WB $M_d$ is almost
diagonal. In principle, since the Yukawa couplings $Y_u$, $Y_d$ are not
constrained in the SM, there is equal probability of $M_d$ being close to $%
M_d = \mbox{diag} (m_d, m_s, m_b)$. corresponding to up-down alignment, or
being close for instance to $M_d = \mbox{diag} (m_b, m_s, m_d)$ in which
case there is up-down misalignment. It is clear that in the SM, assuming
small mixing and hierarchical quark masses, the probability of obtaining
up-down alignment is only 1/6. Given a set of arbitrary quark mass
matrices $M_u$, $M_d$, one can derive necessary and sufficient conditions to
obtain small mixing and up-down alignment, expressed in terms of WB
invariants \cite{Branco:2011aa}.
Since the experimentally verified up-down alignment is not
automatic in the SM, one may wonder whether there is a symmetry which leads
to up-down alignment. In the next subsection, we propose a symmetry which
leads to up-down alignment in the quark sector and when extended to a 2HDM
leads to a natural alignment of the two Higgs doublets in flavour space.
\subsection{$Z_3 \times Z^\prime_3$ symmetry and the two Higgs alignment}
We introduce the following $Z_{3}\times Z_{3}^{\prime }$ symmetry under
which the quark left-handed doublets $Q_{L_{i}}^{0}$ , the right-handed up
quarks $u_{L_{i}}^{0}$, and the right-handed down quarks $d_{L_{i}}^{0}$
transform in the following way:
\begin{equation}
\begin{aligned} Q^0_{L_i} &\longrightarrow P^{\dagger}_{ij}\,Q^0_{L_j},\\
u^0_{R_i} &\longrightarrow P_{ij}\,u^0_{R_j},\\ d^0_{R_i} &\longrightarrow
P_{ij}\,d^0_{R_j}, \end{aligned} \label{222}
\end{equation}
where $Z_{3}$ corresponds to $P=\mathbbm{1}+E_{1}$ and $Z_{3}^{\prime }$ to $
P=\mathbbm{1}+E_{2}$ with:
\begin{equation}
E_{1}=\frac{\omega -1}{2}\left(
\begin{array}{ccc}
1 & -1 & 0 \\
-1 & 1 & 0 \\
0 & 0 & 0
\end{array}
\right) ;\qquad E_{2}=\frac{\omega -1}{3}\left(
\begin{array}{ccc}
\frac{1}{2} & \frac{1}{2} & -1 \\
\frac{1}{2} & \frac{1}{2} & -1 \\
-1 & -1 & 2
\end{array}
\right) \label{e1e2}
\end{equation}
and $\omega =e^{\frac{2\pi i}{3}}$. The Higgs doublets transform trivially
under $Z_{3}\times Z_{3}^{\prime }$. The above symmetry leads to the
following form for the Yukawa matrices $\Gamma _{j}= c_{j}^{d}\Delta $,
$\Omega _{j} =c_{j}^{u}\Delta$, corresponding to the stable solution
of Eq.~(\ref{equ20}).
This can easily be checked since $\Delta E_{1}=\Delta E_{2}=0$.
We thus conclude that the symmetry of Eqs.~(\ref{222}), (\ref{e1e2})
leads to the alignment of the two Yukawa coupling matrices, with a democratic
flavour structure. Note that this solution also guarantees an up-down alignment
in the quark sector, as defined in the previous subsection. \\
In order to give mass to the first two quark generations, the
$Z_{3}\times Z_{3}^{\prime }$ symmetry has to be broken. This breaking
will also lead to Higgs mediated FCNC, but these couplings
will be suppressed by the smallness of the quark masses. In order to illustrate
how a realistic pattern of quark masses and mixing can be obtained, we
shall assume that the breaking of the $Z_{3}\times Z_{3}^{\prime }$ symmetry
occurs in two steps. In the first step the symmetry $Z_{3}\times Z_{3}^{\prime }$
is broken into just one of the $Z_{3}$ and the second generation acquires
mass and finally in the last step the mass of the quarks $u$, $d$ is
generated.
In the first step the symmetry $Z_{3}\times Z_{3}^{\prime }$ is broken to $
Z_{3}$ generated by $P=\mathbbm{1}+E_{1}$. One can check that:
\begin{equation}
\Gamma _{j}=c_{j}^{d}\left( \Delta +\varepsilon _{d}\,A\right) ;\qquad
\Omega _{j}=c_{j}^{u}\left( \Delta +\varepsilon _{u}\,A\right) ;\qquad
A=\left(
\begin{array}{ccc}
0 & 0 & 1 \\
0 & 0 & 1 \\
1 & 1 & 1
\end{array}
\right) \label{go}
\end{equation}
are invariant under this $Z_{3}$ symmetry. Note that $A\,E_{1}=0$. At this
stage the second generation acquires mass. Finally, the lightest quarks, $u$
and $d$ acquire mass through a small perturbation, proportional to $\hat{
\delta}_{d,u}$, which breaks this $Z_{3}$ symmetry. We assume that:
\begin{equation}
\Gamma _{2}=c_{2}^{d}\left( \Delta +\varepsilon _{d}\,A+\hat{\delta}%
_{d}\,B_{d}\right) \label{g1}
\end{equation}
while
\begin{equation}
\Gamma _{1}=c_{1}^{d}\left( \Delta +\varepsilon _{d}\,A\right) ; \label{g2}
\end{equation}
with equivalent expressions for the up sector. Here
\begin{equation}
B_{u}=\left(
\begin{array}{ccc}
0 & 0 & 1 \\
0 & 0 & 0 \\
1 & 0 & 1
\end{array}
\right);\qquad B_{d}=\left(
\begin{array}{ccc}
0 & 0 & 1 \\
0 & 0 & 0 \\
1 & 0 & \eta
\end{array}
\right) \label{bud}
\end{equation}
where $\eta $ is some complex number with modulus of order one. The symmetry
is broken, and neither $B_{u}$ nor $B_{d}$ are invariant under the $
Z_{3}\times Z_{3}^{\prime }$ symmetry.
\section{Suppression of scalar mediated FCNC}
In order to study the suppression of scalar mediated FCNC, it is useful to start
by analysing the parameter space in our framework.
\subsection{The Parameter Space}
From Eqs. (\ref{mumd}), (\ref{g1}) and (\ref{g2}) it follows that, in leading order
\begin{equation}
m_{b}=\frac{3}{\sqrt{2}}\left\vert c_{1}^{d}v_{1}+c_{2}^{d}v_{2}e^{i\alpha
}\right\vert \quad ;\quad m_{t}=\frac{3}{\sqrt{2}}\left\vert
c_{1}^{u}v_{1}+c_{2}^{u}v_{2}e^{-i\alpha }\right\vert \label{conv}
\end{equation}
Writing $v\equiv \sqrt{v_{1}^{2}+v_{2}^{2}}=v_{1}\sqrt{1+t^{2}}$, with
\begin{equation}
t=\frac{v_{2}}{v_{1}} \label{t}
\end{equation}
we obtain in leading order the following relations:
\begin{equation}
\frac{\left\vert c_{1}^{d}+c_{2}^{d}\ te^{i\alpha }\right\vert }{\sqrt{
1+t^{2}}}=\frac{\sqrt{2}}{3}\frac{m_{b}}{v}\quad ;\quad \frac{\left\vert
c_{1}^{u}+c_{2}^{u}\ te^{-i\alpha }\right\vert }{\sqrt{1+t^{2}}}=\frac{\sqrt{%
2}}{3}\frac{m_{t}}{v} \label{ci}
\end{equation}
which impose restrictions on the allowed parameter space.
A priori, we do not assume any conspiracy between parameters and take $
t=O(1)$.
It is then clear from Eq. (\ref{ci}), that the $c_{i}^{u}$ are generically
of order one, while $c_{i}^{d}$ are smaller, and may assume values of order $
O(\frac{m_{b}}{m_{t}})$. This is an important ingredient which, as we shall
see, will play a r\^{o}le in the evaluation of the strength of FCNC's and the allowed
parameter space for the Higgs masses.
Next we give the structure of the flavour changing neutral Yukawa couplings.
For that, it is useful
to express the quark mass matrices in Eq. (\ref{mumd}) in terms of the
perturbations given in Eqs. (\ref{g1}), (\ref{g2}):
\begin{equation}
\begin{array}{l}
M_{d}=\frac{v_{1}}{\sqrt{2}}\left( c_{1}^{d}+c_{2}^{d}te^{i\alpha }\right) \
\left[ \Delta +\varepsilon _{d}\ A+\delta _{d}\ B_{d}\right] \\
\\
M_{u}=\frac{v_{1}}{\sqrt{2}}\left( c_{1}^{u}+c_{2}^{u}te^{-i\alpha }\right)
\ \left[ \Delta +\varepsilon _{u}\ A+\delta _{u}\ B_{u}\right]%
\end{array}
\quad ;\quad
\begin{array}{c}
\delta _{d}\equiv \frac{c_{2}^{d}te^{i\alpha }}{c_{1}^{d}+c_{2}^{d}te^{i%
\alpha }}\ \widehat{\delta }_{d} \\
\\
\delta _{u}\equiv \frac{c_{2}^{u}te^{-i\alpha }}{c_{1}^{u}+c_{2}^{u}te^{-i%
\alpha }}\ \widehat{\delta }_{u}
\end{array}
\label{quark1}
\end{equation}
Then, we derive the expressions for the matrices which
couple to the Higgs scalars in Eqs. (\ref{rep}), (\ref{ndnu}). In the basis
where the up and down quark matrices are diagonal, the matrices $N_{d}$, $
N_{u}$ of Eq. ( \ref{ndnu1}) become
\begin{equation}
\begin{array}{c}
N_{d}=t\ D_{d}-\frac{v_{1}}{\sqrt{2}}\left( 1+t^{2}\right) e^{i\alpha }\
U_{d_{L}}^{\dagger }\ \Gamma _{2}\ U_{d_{R}} \\
\\
N_{u}=t\ D_{u}-\frac{v_{1}}{\sqrt{2}}\left( 1+t^{2}\right) e^{-i\alpha }\
U_{u_{L}}^{\dagger }\ \Omega _{2}\ U_{u_{R}}
\end{array}
\label{ns}
\end{equation}
where we have used Eqs.~(\ref{ndnu}) and (\ref{ndnu1}) with Eq.~(\ref{mumd}).
Finally, from Eq.~(\ref{ns}) combined with Eqs.~(\ref{g1}) and Eq. (\ref{quark1})
we find
\begin{equation}
\begin{array}{l}
N_{d}=\ \frac{c_{1}^{d}t-c_{2}^{d}e^{i\alpha }}{
c_{1}^{d}+c_{2}^{d}te^{i\alpha }}\ D_{d} - {\frac{v}{\sqrt{2}}}\
\frac{c_{1}^{d}\sqrt{1+t^{2}}}{t}\ \delta _{d}\
U_{d_{L}}^{\dagger }\ B_{d}\ U_{d_{R}} \\
\\
N_{u}=\frac{c_{1}^{u}t-c_{2}^{u}e^{-i\alpha }}{%
c_{1}^{u}+c_{2}^{u}te^{-i\alpha }}\ D_{u} - {\frac{v}{\sqrt{2}}} \frac{c_{1}^{u}\sqrt{%
1+t^{2}}}{t}\ \delta _{u}\ U_{u_{L}}^{\dagger }\ B_{u}\ U_{u_{R}}%
\end{array}
\label{ns1}
\end{equation}
where $D_{d}\equiv diag(m_{d},m_{s},m_{b})$, $D_{u}\equiv
diag(m_{u},m_{c},m_{t})$.
The crucial point is that in our scheme these matrices have an extra
suppression factor, proportional to $\delta _{d,u}$.
Using the expressions given in Eqs. (\ref{quark1}), (\ref{bud}) and
computing the trace, second invariant and determinant for the squared quark
mass matrices $H_{u,d}\equiv \left( MM^{\dagger }\right) _{u,d}$ , one can
find that in leading order:
\begin{equation}
\begin{array}{c}
\delta _{d}=\sqrt{3}\sqrt{\frac{m_{d}}{m_{s}}}\frac{m_{s}}{m_{b}}=O\left(
\lambda ^{3}\right) \\
\\
\delta _{u}=\sqrt{3}\sqrt{\frac{m_{u}}{m_{c}}}\frac{m_{c}}{m_{t}}=O\left(
\lambda ^{5}\right)
\end{array}
\label{dud}
\end{equation}
where $\lambda \equiv 0.2$ is of the order of the Cabibbo angle.
From Eq. (\ref{quark1}) it follows that in leading order $
U_{d_{L}}=U_{u_{L}}=U_{d_{R}}=U_{u_{R}}=F$, where
\begin{equation}
F==\left(
\begin{array}{ccc}
1/\sqrt{2} & 1/\sqrt{6} & 1/\sqrt{3} \\
-1/\sqrt{2} & 1/\sqrt{6} & 1/\sqrt{3} \\
0 & -2/\sqrt{6} & 1/\sqrt{3}%
\end{array}
\right) \label{f}
\end{equation}
is the matrix that diagonalizes the exact democratic limit $\Delta $. Thus,
taking into account Eq. (\ref{bud}), the matrix contributions from $U_{d_{L}}^{\dagger }\
B_{d}\ U_{d_{R}}$ and $U_{u_{L}}^{\dagger }\ B_{u}\ U_{u_{R}}$ are both of
order one. One can thus conclude that:
--for the down sector, with the assumptions made after Eq. (\ref{ci}), we
have a total suppression factor of $O(\frac{m_{b}}{m_{t}})\cdot O\left(
\lambda ^{3}\right) $
--for the up sector, we have a suppression factor of $O\left( \lambda
^{5}\right) $ or smaller depending on the value that we choose to assume for
$c_{1}^{u}$, but which, as explained, it is reasonable to take of order one.
\section{Numerical analysis}
The matrices of Eq.~\eqref{quark1} may be explicitly written as
\begin{equation}
M_{u}= c_{u} \begin{pmatrix}
1 & 1 & 1 + \varepsilon + \delta\\
1 & 1 & 1 + \varepsilon\\
1 + \varepsilon + \delta & 1 + \varepsilon & 1 + \varepsilon + \delta
\end{pmatrix}_{u},\quad
M_{d}= c_{d} \begin{pmatrix}
1 & 1 & 1 + \varepsilon + \delta\\
1 & 1 & 1 + \varepsilon\\
1 + \varepsilon + \delta & 1 + \varepsilon & 1 + \varepsilon + \eta\,\delta
\end{pmatrix}_{d},
\end{equation}
where we have introduced $c_d\equiv\frac{v_1}{\sqrt{2}}(c^d_1+c^d_2\,t\,e^{i\alpha})$, and $c_u \equiv \frac{v_1}{\sqrt{2}}(c^u_1+c^u_2\,t\,e^{-i\alpha})$. Although these two coefficients are in general complex, and since the physically meaningful matrices are those defined as $H=M\,M^\dagger$, both coefficients
may be taken as real for our numerical exercise. If one then parametrizes the remaining variables as
\begin{equation}
\varepsilon = \varepsilon_m \exp\left(i\,\varepsilon_f\right),\qquad \delta = \delta_m \exp\left(i\,\delta_f\right),\qquad \eta = \eta_m \exp\left(i\,\eta_f\right),
\end{equation}
one is left with twelve real parameters that compose the quark mass matrices in our scheme.
In order to check if this parameter space could accommodate the flavour sector, a numerical survey was made where we looked for one combination that could fit the observed values of the quark masses given at the scale of the Z boson mass \cite{Antusch:2013jca}, the moduli of the entries of the CKM matrix \cite{ckmfitter},
the strength of CP violation $I_{CP}$ and $\sin 2\beta$ and $\gamma$ \cite{ckmfitter}, with $\beta$ and $\gamma$ being two of the angles of the unitarity triangle. A simple run of all twelve
parameters produced a ``reference point'':
\par
\begin{center}
\begin{tabular}{ccc}
\hline\hline
& Up-sector & Down-sector\\
\hline
$c$ & 56.73 & 0.89\\
$\varepsilon_m$ & $1.6 \times 10^{-2}$ & 0.11\\
$\varepsilon_f$ &$ -5.6 \times 10^{-3}$ & 0.41\\
$\delta_m$ & $8.1 \times 10^{-4}$ &$ 2.2 \times 10^{-2}$\\
$\delta_f$ &$ \pi + 0.32$ & 2.26\\
$\eta_m$ & --- & 4.99\\
$\eta_f$ & --- & $\pi + 0.62$ \\ \hline \hline
\end{tabular}
\end{center}
\par
\noindent which yields the output values:
\begin{equation}
\begin{aligned}
D_d &= \textrm{diag}(0.00204,0.05824,2.85356) \ \mbox{GeV},\\
D_u &= \textrm{diag}(0.00114,0.61736,171.684) \ \mbox{GeV},\\
| V_{CKM}| &= \begin{pmatrix}
0.9745 & 0.2244 & 0.0036\\
0.2243 & 0.9737 & 0.0415\\
0.0087 & 0.0407 & 0.9991
\end{pmatrix},\\
|I_{CP}| &= 3.0 \times 10^{-5},\\
\sin 2\beta &= 0.69,\\
\gamma &= 69.3^\circ.
\end{aligned}
\end{equation}
It should be noted that the twelve parameters fix not only $V_{CKM}$
and the quark mass spectrum, but also the strength of all the FCNC couplings.
In order to evaluate the numerical stability of this reference point, we performed
a numerical check, varying the input parameters randomly around the values
that produced the reference point above;
the new results were then combined in the scatter plots shown in
Fig.~\ref{fig:scatter} where the reference point is highlighted.
\begin{figure}
\centering
\subfigure{
\centering
\includegraphics[width = 0.427\textwidth]{JS2b.pdf}}\quad
\subfigure{
\centering
\includegraphics[width = 0.4\textwidth]{S2bVub.pdf}}
\caption{We present scatter plots showing $|I_{CP}|$ versus $\sin 2 \beta$
and $\sin 2 \beta$ versus $|V_{ub}|$ obtained by varying randomly the
input parameters around the reference point.}
\label{fig:scatter}
\end{figure}
In order to obtain an estimate of the lower bound for the flavour-violating
Higgs masses, we consider the contribution to $K^0-\overline{K}^0$ mixing.
Apart from the SM box diagram
one now has a New Physics contribution arising from
the scalar-mediated FCNC tree-level diagrams thus making the total transition amplitude equal to $M_{12} = M_{12}^{\textrm{SM}} + M_{12}^{\textrm{NP}}$, \cite{Botella:2014ska} with:
\begin{equation}
M_{12}^{NP} = \sum_{H=R,I} \frac{f_M^2\,m_M}{96\,v^2\,m_H^2}\,\bigg\{\bigg[1+\bigg(\frac{m_M}{m_{q1}+m_{q2}}\bigg)^2\bigg]C_1(H) - \bigg[1+11\bigg(\frac{m_M}{m_{q1}+m_{q2}}\bigg)^2\bigg]C_2(H)\bigg\},
\end{equation}
where:
\begin{equation}
C_1(R) = \left[\left( N_{q2q1}\right)^{*} + N_{q1q2} \right]^2,\quad
C_2(R) = \left[\left( N_{q2q1}\right)^{*} - N_{q1q2} \right]^2,
\end{equation}
and
\begin{equation}
C_1(I) = -\left[\left( N_{q2q1}\right)^{*} - N_{q1q2} \right]^2,\quad
C_2(I) = -\left[\left( N_{q2q1}\right)^{*} + N_{q1q2} \right]^2.
\end{equation}
The indices $q_1$ and $q_2$ refer to the valence quarks of the meson $M$, and
$N$ is $N_u$ or $N_d$ depending on the meson system considered.
In this framework it is a good approximation to use the matrix $F$ for
both $U_{d_L}$ and $U_{d_R}$.
Using the values we obtained for $\delta_d$ and taking, as already discussed, $t\simeq 1$ and $c_1^d\simeq
\frac{\sqrt{2}}{3}\frac{m_b}{v}$, the new physics contribution to $M_{12}^K$
becomes solely dependent on $f_K$, $m_K$, $m_R$ and $m_I$. In $K^0-
\overline{K}^0$, both $M_{12}^K$ and $\Gamma_{12}^K$ are relevant for the
mass difference $\Delta m_K$. It is reasonable to impose the constraint
that $M_{12}^{\textrm{NP}}$ in the neutral kaon system does not exceed the
experimental value of $\Delta m_K$. Adopting as input values the PDG experimental determinations of $f_K$, $m_K$ and $\Delta m_K$ \cite{PDG:2014}, one is left with combinations of $m_R$ and $m_I$ where our model respects the inequality $M_{12}^{NP\,(K)}<\Delta m_K$. The region plot that we have obtained is presented in Fig.\ref{fig:region}. It is clear that in this framework, the masses of the flavour-violating
neutral Higgs can be below the TeV scale, so that they could be
discovered at the next run of the LHC.
\begin{figure}
\centering
\includegraphics[width = 0.4\textwidth]{quasi-2H.pdf}
\caption{Plot showing the allowed region for $m_I$ and $m_R$, taking
into account the constraint on $\Delta m_K$.}
\label{fig:region}
\end{figure}
\section{Conclusions}
We have studied in detail, in the framework of 2HDM, the question of stability
of alignment, under the renormalization group. It was shown that there are new
stable solutions, apart from those found in Ref.~\cite{Ferreira:2010xe}. Stability under the RGE puts very strict restrictions on the flavour structure of the Higgs
Yukawa couplings. If one imposes the stability conditions and at the same time
requires that the flavour structure is in agreement with the observed pattern
of quark masses and mixing, then one is lead to a unique solution, where
all Higgs flavour matrices are proportional to the so-called democratic matrix. We
have also shown that these flavour structures leading to stable alignment
can be obtained by imposing on the Lagrangian a $Z_{3}\times Z_{3}^{\prime }$
symmetry.
In the limit where this symmetry is exact, only the
third generation of quarks acquires a mass. Non-vanishing masses for the two
first generations are obtained through the breaking of the discrete symmetry
which in turn generates scalar mediated FCNC which are suppressed by the
smallness of the light quark masses. The scenario presented in this paper
establishes a possible intriguing link between stability of alignment in 2HDM
and the observed pattern of quark masses and mixing.
\section*{Appendix: Solutions to the alignment conditions}
The solutions to the alignment conditions
\begin{equation}
\begin{array}{ccc}
\Omega _{1}\Omega _{1}^{\dagger }\Gamma _{1}=\lambda _{\Gamma }\Gamma _{1} &
; & \Gamma _{1}\Gamma _{1}^{\dagger }\Omega _{1}=\lambda _{\Omega }\Omega
_{1}%
\end{array}
\label{A1}
\end{equation}
can be obtained by the following steps. First we define the Hermitian
matrices
\begin{equation}
\begin{array}{ccc}
H_{\Gamma }=\Gamma _{1}\Gamma _{1}^{\dag } & ; & H_{\Omega }=\Omega
_{1}\Omega _{1}^{\dagger }
\end{array}
\label{A2}
\end{equation}
It is easy to show that $\lambda _{\Gamma }$ and $\lambda _{\Omega }$ are
real. This can be achieved by multiplying the first equation by its
Hermitian conjugate an inserting the second equation ( and viceversa) to get
\begin{eqnarray}
\lambda _{_{\Omega }}H_{\Omega }^{2} &=&\left\vert \lambda _{\Gamma
}\right\vert ^{2}H_{\Gamma } \label{A3} \\
\lambda _{\Gamma }H_{\Gamma }^{2} &=&\left\vert \lambda _{_{\Omega
}}\right\vert ^{2}H_{\Omega } \label{A4}
\end{eqnarray}
it follows from these equations that both $\lambda _{_{\Omega }}$ and $\lambda _{\Gamma
}$ should be real since one has two identities among Hermitian matrices.
Now multiplying each of the Eqs.~(\ref{A1}) on the right
by $\Gamma _{1}^{\dagger }$ and
$\Omega _{1}^{\dagger }$ respectively we get
\begin{equation}
\begin{array}{ccc}
H_{\Omega }H_{\Gamma }=\lambda _{\Gamma }H_{\Gamma } & ; & H_{\Gamma
}H_{\Omega }=\lambda _{\Omega }H_{\Omega }
\end{array}
\label{A5}
\end{equation}
and taking Hermitian conjugates
\begin{equation}
\begin{array}{ccc}
H_{\Gamma }H_{\Omega }=\lambda _{\Gamma }H_{\Gamma } & ; & H_{\Omega
}H_{\Gamma }=\lambda _{\Omega }H_{\Omega }
\end{array}
\label{A5H}
\end{equation}
therefore
\begin{equation}
\lambda _{\Gamma }H_{\Gamma }=\lambda _{\Omega }H_{\Omega } \label{A6}
\end{equation}
and we conclude that
\begin{equation}
\left[ H_{\Gamma },H_{\Omega }\right] =0 \label{A7}
\end{equation}
implying that $V_{CKM}=I$ up to permutations of rows or columns. Denoting
the usual bi-unitary diagonalisation procedure by
\begin{equation}
\begin{array}{ccc}
\Gamma _{1}=V_{L}^{\Gamma }D_{\Gamma }V_{R}^{\Gamma \dag } & ; & \Omega
_{1}=V_{L}^{\Omega }D_{\Omega }V_{R}^{\Omega \dag }
\end{array}
\label{A8}
\end{equation}
from Eq.~(\ref{A7}) we conclude that we can always choose the unitary matrices $
V_{L}^{\Gamma }$ and $V_{L}^{\Omega }$ equal to each other
\begin{equation}
V_{L}^{\Gamma }=V_{L}^{\Omega } \label{A9}
\end{equation}
and the alignment conditions can be easily reduced to conditions among the
diagonal matrices $D_{\Gamma }$ and $D_{\Omega }$. From Eq.~(\ref{A5}),
it then follows that:
\begin{equation}
\begin{array}{ccc}
D_{\Omega }^{2}D_{\Gamma }=\lambda _{\Gamma }D_{\Gamma } & ; & D_{\Gamma
}^{2}D_{\Omega }=\lambda _{\Omega }D_{\Omega }
\end{array}
\label{A10}
\end{equation}
It can be checked that there are only two types of solutions. Those with $\lambda
_{\Gamma }$ and $\lambda _{\Omega }$ different from zero (solutions 1,2 and
3) and the remaining ones (solutions 4 and 5 )
\begin{enumerate}
\item $D_{\Gamma }=aP_{3}$ and $D_{\Omega }=\alpha P_{3}$ and changes of $%
P_{3}$ by $P_{2}$ or $P_{1}$
\item $D_{\Gamma }=a\left( I-P_{1}\right) $ and $D_{\Omega }=\alpha \left(
I-P_{1}\right) $ and changes of $P_{1}$ by $P_{2}$ or $P_{3}$
\item $D_{\Gamma }=aI$ and $D_{\Omega }=\alpha I$
\item $D_{\Gamma }=0$ and $D_{\Omega }$ arbitrary and viceversa.
\item $D_{\Gamma }=aP_{i}$ and $D_{\Omega }=\alpha \left( I-P_{i}\right) $
\end{enumerate}
where $P_{i}$ stand for the projection operators
\begin{equation}
\left( P_{i}\right) _{jk}=\delta _{ij}\delta _{ik} \label{A11}
\end{equation}%
Solutions 2,3 and 4 cannot be good approximations to the actual quark
spectra due to the implied degeneracy. Solution 5 gives rise to $V_{CKM}$ matrix very
different from the identity matrix. Only solution 1 provides, in leading approximation
the correct pattern of quark masses and mixing. In a suitable weak-basis, this solution
can be written as a democratic matrix $\Delta$.
\section*{Acknowledgments}
This work is partially supported by Spanish MINECO under grant FPA2011-23596,
by Generalitat Valenciana under grant GVPROMETEOII 2014-049
and by Funda\c{c}\~ao para a Ci\^encia e a
Tecnologia (FCT, Portugal) through the projects CERN/FP/123580/2011,
PTDC/FIS-NUC/0548/2012, EXPL/FIS-NUC/0460/2013, and CFTP-FCT Unit 777
(PEst-OE/FIS/UI0777/2013) which are partially funded through POCTI (FEDER),
COMPETE, QREN and EU.
|
\section{Introduction}
Blazars are active galactic nuclei (AGN) that are variable over the
entire electromagnetic spectrum. Their variability is enhanced due to
Doppler-beamed emission from a relativistic jet pointing close to the
line of sight. Blazars having featureless optical spectra or showing emission lines with
equivalent width $< 5$~{\AA} are historically classified as BL Lacs
\citep{stocke91}.
Markarian 421 (hereafter Mrk~421) is one of the best studied BL~Lac
objects. It is relatively nearby with a redshift of $z=0.031$ \citep{ulrich75}.
It is classified as a high synchrotron peaked blazar \citep{abdo10c} based on its
spectral energy distribution (SED). It was the
first blazar detected at TeV energies
\citep{punch92}, and since then it has been the subject of numerous multi-wavelength
campaigns \citep[e.g.,][]{aleksic12,abdo11,acciari11,horan09,donnarumma09,fossati08,rebillot06,
tosti98}. The broad-band SED of Mrk~421 can often be modeled
with a one-zone synchrotron self-Compton (SSC) model \citep{aleksic12,
acciari11, abdo11} or with lepto-hadronic models involving proton
synchrotron radiation and/or
photopion interactions \citep[e.g.,][]{bottcher13, mastichiadis13, dimitrakoudis14}.
Mrk~421 exhibits large variations in the TeV,
GeV, X-ray, and optical wavebands
\citep[e.g.,][]{acciari11,acciari09,horan09,blazejowski05}, with
correlated TeV and X-ray variations
\citep[e.g.,][]{giebels07,fossati08}. The connection between the TeV and
optical bands is less clear and more detailed time-dependent models
are being developed to study the complex correlations (see e.g.,
\citealt{chen11} for SSC modeling and \citealt{mastichiadis13} for
lepto-hadronic modeling).
In 2012 Mrk~421 underwent its largest radio flare ever observed at
15\,GHz \citep{hovatta12ATel}. The flare time-scale was
faster and its amplitude larger than
any other radio flare observed
from this source in the past 30 years of observations with the University of
Michigan Radio Astronomy Observatory monitoring program
\citep{richards13}.
The radio flare occurred about 40~d
after a major flare was observed in the $\gamma$-ray band
\citep{Dammando12ATel} by the {\it Fermi Gamma-Ray Space Telescope} (hereafter {\it
Fermi}). Unfortunately Mrk~421 was close to the sun at the
time of the radio flare so that only limited multi-wavelength coverage in
addition to the radio and $\gamma$-ray bands exists, apart from a
NuSTAR calibration observation that covered the early activity \citep{balokovic13b}. In the spring of 2013,
Mrk~421 underwent another major high-energy flaring event in the
X-ray to TeV energies \citep{cortina13,balokovic13,paneque13}. About 60~d later, it was
followed by fairly small amplitude radio flares in the radio and
millimeter bands \citep{hovatta13ATel}.
In this paper we present the radio data obtained with the Owens Valley
Radio Observatory (OVRO) 40-m telescope at 15\,GHz, and 95\,GHz data
obtained with Combined Array for Research in Millimeter-Wave Astronomy
(CARMA). We compare the radio variations with the $\gamma$-ray light
curves obtained by {\it Fermi}, by performing a cross-correlation
analysis on the full data sets available from 2008 to 2013. Based on
our findings and the coincidence of historically rare
extreme flares in the two bands, we assume that the events in the
radio and $\gamma$-ray bands are physically connected. Under this
assumption we aim to understand the extreme nature of the 2012 radio
event by adopting a reasonable physical model with the smallest
number of free parameters.
Our paper is organized as follows. We describe
the observations and data reduction in Sect.~\ref{sect:obs}. The light curves and
cross-correlations between the different bands are shown in
Sect.~\ref{sect:lc}. We model the flares in Sect.~\ref{sect:model}, discuss our results in
Sect.~\ref{sect:discussion}, and present our conclusions in Sect.~\ref{sect:conclusions}.
Throughout the paper we use a cosmology where $H_0 = 71~\mathrm{km}~\mathrm{s}^{-1}~\mathrm{Mpc}^{-1}$,
$\Omega_M = 0.3$, and $\Omega_\Lambda = 0.7$
\citep[e.g.,][]{komatsu09}.
\section{Observations}\label{sect:obs}
We present radio and $\gamma$-ray observations of the two flaring
events observed in 2012 and 2013. We are especially
concerned with the extreme 2012 radio event, whereas the 2013 flare is
more typical of the radio variability observed in Mrk~421.
\subsection{Radio observations}
Mrk~421 was observed as part of the blazar monitoring program\footnote{http://www.astro.caltech.edu/ovroblazars/} with the
OVRO 40-m telescope \citep{richards11}. In this program, a sample of
over 1800 AGN are observed twice per week at 15\,GHz. Mrk~421 has been
included since the beginning of the monitoring in late 2007.
The OVRO 40-m telescope uses off-axis dual-beam optics and a cryogenic high
electron mobility transistor (HEMT) low-noise amplifier with a
15.0~GHz center frequency and 3~GHz bandwidth. The two sky beams are Dicke switched using the
off-source beam as a reference, and the source is alternated between
the two beams in an ON-ON fashion to remove atmospheric and ground
contamination. A noise level of approximately 3--4~mJy in quadrature
with about 2 per cent additional uncertainty, mostly due to pointing errors,
is achieved in a 70~s integration period. Calibration is achieved
using a temperature-stable diode noise source to remove receiver gain
drifts and the flux density scale is derived from observations of
3C~286 assuming the \cite{baars77} value of 3.44~Jy at
15.0~GHz. The systematic uncertainty of about 5 per cent in the flux density
scale is not included in the error bars. Complete details of the
reduction and calibration procedure are given in Richards et
al. (2011).
For comparison and to fill gaps in the OVRO sampling,
we also consider the 37 GHz light curve obtained by the Mets\"ahovi
Radio Observatory. These observations were made with the
13.7-m radome-enclosed telescope using a 1\,GHz-bandwidth
dual beam receiver centered at 36.8\,GHz. A detailed description of
the data reduction and analysis is given in \cite{terasranta98}. As
the uncertainty in the data points is much larger than in the other
bands, we do not include the data in any subsequent modeling.
\subsection{Millimeter-band observations}
Since February 2013, Mrk~421 was observed $1-3$ times per week with
CARMA as a part of the MARMOT\footnote{Monitoring of
$\gamma$-ray Active galactic nuclei with Radio, Millimeter and
Optical Telescopes; http://www.astro.caltech.edu/marmot/} blazar
monitoring project. The observations were made using the
eight 3.5~m telescopes of the array with a central frequency of
95\,GHz and a bandwidth of 7.5\,GHz. The data were reduced
using the MIRIAD (Multichannel Image Reconstruction, Image Analysis
and Display) software \citep{sault95}, including standard
bandpass calibration on a bright quasar. The amplitude and phase
gain calibration was done
by self-calibrating on Mrk~421. The absolute flux calibration was
determined from a temporally nearby observation (within a day) of the planets
Mars, Neptune or Uranus, whenever possible. Otherwise the quasar
3C~273 was used as a secondary calibrator. The estimated absolute
calibration uncertainty of 10 per cent is not included in the error
bars.
\subsection{Gamma-ray data}
The $\gamma$-ray data were obtained with the Large Area Telescope
(LAT) aboard {\it Fermi}, which observes the entire sky every 3 hours at energies of
$0.1-300$\,GeV \citep{atwood09}. The publicly available reprocessed Pass
7 data\footnote{http://fermi.gsfc.nasa.gov/ssc/data/} were downloaded and analysed using the {\it Fermi} ScienceTools software
package version v9r32p5. The data were binned using the adaptive binning method of
\cite{lott12}. The uneven bin size was determined in such a way that
the statistical error
in each flux measurement is $\sim$15 per cent
We used the instrument response functions
P7REP\_SOURCE\_V15, Galactic diffuse emission model ``gll\_iem\_v05.fits'' and
isotropic background model ``iso\_source\_v05.txt''.\footnote{http://fermi.gsfc.nasa.gov/ssc/data/access/lat/BackgroundModels.html} Source class
photons (evclass=2) within 15$^\circ$ of Mrk~421 were selected, with
a zenith angle cut of 100$^\circ$ and a rocking angle cut of 52$^\circ$.
Once the bins were defined, the photon fluxes in the energy range of 0.1 -
200\,GeV were calculated using unbinned likelihood
and the tool {\it gtlike} with the Minuit optimizer. All sources within
15$^\circ$ of Mrk~421 were included in the
source model with their spectral parameters, except the flux, frozen to the values determined
in the 2nd {\it Fermi} LAT catalog (2FGL; \citealt{nolan12}). For sources more
than 10$^\circ$ from Mrk~421 we also froze the fluxes to the 2FGL
value. The 10 per cent systematic uncertainty below 100 MeV, decreasing
linearly in Log(E) to 5 per cent in the range between 316~MeV and 10~GeV
and increasing linearly in Log(E) up to 15 per cent at 1~TeV \citep{ackermann12c} is not included in the error bars.
For the purpose of the modeling presented in
Sect.~\ref{sect:model}, we also obtained the energy fluxes, with a power-law spectral
model where the index is frozen to the 2FGL catalog value of
$\Gamma=1.77$.
\section{Light curves}\label{sect:lc}
The radio, millimeter and $\gamma$-ray light curves are shown in
Fig.~\ref{fig:lc}. They cover the time range since the beginning of
the {\it Fermi} mission in August 2008 (MJD~54688) until the end of
October 2013 (MJD~56610). The light curves illustrate the unusual nature of the 2012
flare (flare 1) in both radio and $\gamma$ rays, that lasted from May
2012 until October 2012 (MJD 56060 - 56225). This is a unique
event, especially in the radio band where such fast and prominent
flares have not been observed before in Mrk~421. The 2013 flare
(flare 2) from March 2013 until November 2013 (MJD 56350-56610), although prominent, is
much broader and of lower amplitude.
\begin{figure}
\includegraphics[scale=0.4]{f1.eps}
\caption{From top to bottom: light curves of Mrk~421 in $\gamma$ rays from {\it Fermi} LAT, radio 15\,GHz from OVRO, 37\,GHz from Mets\"ahovi, and 95\,GHz from CARMA. The light curves span a time period from August 2008 until
November 2013, except for the CARMA light curve that starts in February 2013. The solid lines indicate the time range used to model
the 2012 flare from 13 May 2012 (MJD 56060) until 25 October 2012 (MJD
56225). The dashed lines indicate the time range for the 2013 flare
from 9 March 2013 (MJD 56360) until 14 November 2013 (MJD
56610). Because of the larger uncertainties in the data, the
37\,GHz light curve is not used in any subsequent modeling.}\label{fig:lc}
\end{figure}
A full cross-correlation analysis between more than four years of OVRO and
LAT data was done by \cite{max-moerbeck14}. They used weekly binned
$\gamma$-ray light curves and data from 2008 until November 2012,
including the rapid 2012 flare. They found a peak in the
discrete correlation function (DCF), with the $\gamma$ rays leading
the radio by $40\pm9$~d. The significance of this correlation was
between 96.16 and 99.99 per cent depending on the power spectral density (PSD) model used for the
light curves, with a best-fitting value of 98.96 per cent. For details of the significance estimation, see
\cite{max-moerbeck14a}.
We repeated the cross-correlation
analysis using the extended light curves considered here, and found that the DCF
shows
a broad peak ($\sim$ 30~d). In particular, the time delay ranges from 40 to 70~d with $\gamma$ rays leading,
consistent with the estimate from \cite{max-moerbeck14}. The
significance of the peak was from 91.90 to 99.99 per cent, with a best fit
value of 97.36 per cent, depending on the PSD model. The difference
compared to the exact value derived by \cite{max-moerbeck14} is
because they only considered the maximum of the
DCF, while the peak itself is broad.
Recently, \cite{emmanoulopoulos13}
introduced a method for simulating light curves, which also accounts
for their flux distribution. This is more appropriate in the $\gamma$-ray
light curves, as the light curves have a non-Gaussian photon
flux distribution. Using this
method, the significance of the correlation increases to 99.82 per cent, when
using the best-fitting PSD. In the rest of the paper, we will thus assume that the major flares in radio and
$\gamma$ rays are physically connected.
The $\gamma$-ray flares in both 2012 and 2013 have
significant sub-structure, as already shown in Fig.~\ref{fig:lc}
(top panel). If we also consider the broadness of the DCF peak, it
becomes unclear which $\gamma$-ray spike of the overall flare is
most plausibly associated with the radio
peaks. Therefore, we consider both alternatives in our
modeling.
\subsection{The flare of 2012}\label{sect:2012}
On 16 July 2012 (MJD 56124) the $\gamma$-ray flux reached the highest
value since the start of the {\it Fermi} mission \citep{Dammando12ATel}. The
$\gamma$-ray flux increased by a factor of three from $(3.4 \pm 0.6)
\times10^{-7}$\,ph\,cm$^{-2}$\,s$^{-1}$ to $(1.1 \pm 0.2) \times10^{-6}$\,ph\,cm$^{-2}$\,s$^{-1}$ within seven days. The flare
appears double peaked with the second flare peaking on 15 August 2012 (MJD 56154) at a
flux of $(9.5 \pm 0.1)\times10^{-7}$\,ph\,cm$^{-2}$\,s$^{-1}$. The total duration of the flaring event was about
60~d. The OVRO 15\,GHz light curve exhibits a fast rise which leads
to an increase of the flux density by a factor of about two, i.e. from 0.6 to 1.1 Jy, in the period
6 August 2012 -- 21 September 2012.
This flux density is higher than any flux density measured at 15\,GHz
in the OVRO program or during the preceding 30-plus years of monitoring with the University of Michigan Radio
Astronomy Observatory \citep{richards13}.
At the time of the first $\gamma$-ray peak there is a gap in the
OVRO 15\,GHz light curve. Thus, a double peaked radio flare cannot
be excluded {\sl a priori}. However, a higher frequency radio light
curve observed at 37\,GHz at Mets\"ahovi Radio Observatory (third
panel from the top in Fig.~\ref{fig:lc}) shows only a
single flare during 2012. Given the spectral proximity of the two radio
bands and the similar features in the two light curves, it is safe to assume
that the OVRO sampling does not miss a peak at the beginning of the
event.
Because of the larger statistical uncertainties in the 37\,GHz
data, we do not include them in our subsequent analysis as they would
not add additional constraints on the modeling.
In order to obtain a better estimate of the time delay between the
$\gamma$-ray and 15\,GHz light curves for this flare only, we use the
DCF method over the time period of the flare ( Fig.~\ref{fig:cc1}, bottom panel). There are two peaks in the DCF which
simply correspond to the delays between each of the spikes in
the double-peaked $\gamma$-ray flare and the single radio
flare. The DCF peaks are both consistent with the
delay measurement we obtained for the full light curves.
Although the correlation is stronger for the $\sim$40~d lag, the
amplitude of the DCF peak at about $-70$~d is not low
enough to justify exclusion of a possible association of
the radio flare with first $\gamma$-ray spike. Thus, we will test
both possibilities in Sect.~\ref{sect:model}.
\begin{figure}
\includegraphics[scale=0.52]{f2}
\caption{$\gamma$-ray flux (top) and 15\,GHz flux density (middle)
during the 2012 flare. The discrete correlation function between
the two light curves is shown in the bottom panel. A negative time
delay means that $\gamma$ rays are leading the radio. The (blue)
solid line shows the exponential fit that is used to estimate the
variability brightness temperature and Doppler factor. The residuals
of the fit are on average less than 20~mJy. The time range in the
figure corresponds to the solid lines in
Fig.~\ref{fig:lc}}.\label{fig:cc1}
\end{figure}
We can estimate the Doppler boosting factor of the radio flare by assuming
that the rise time of the flare corresponds to the light travel time
across the emission region \citep{lahteenmaki99b, hovatta09}. We fit
an exponential function of the form
\begin{equation}
S(t) = \Delta S e^{(t-t_{max})/t_{rise}},
\end{equation}
to the light curve, where $\Delta S$ is the amplitude of the flare,
$t_{max}$ is the peak location of the flare and $t_{rise}$ is the rise
time of the flare. The decay time of the flare has been frozen to 1.3 times the rise time,
as in \cite{lahteenmaki99b}.
The fitting is done using the MultiNest Markov-Chain Monte Carlo (MCMC) algorithm
\citep{feroz08,feroz09,feroz13}. As shown
in Fig.~\ref{fig:cc1} middle panel, an exponential function fits the 15\,GHz flare
fairly well. From the fit we obtain the rise time of the flare
$t_{rise}=10.6\pm0.5$~d and amplitude $\Delta S = 0.52\pm0.01$~Jy,
where the uncertainties are estimated from the MCMC analysis.
Assuming an emission region with the geometry of a uniform disk, we
can estimate the lower limit of the variability brightness temperature
\citep{lahteenmaki99,hovatta09}
\begin{equation}\label{eq:tb}
T_{\rm var} = 1.548 \times 10^{-32} \frac{\Delta S d_L^2}{\nu^2t_{rise}^2(1+z)},
\end{equation}
where $d_L$ is the luminosity distance to the object in meters (here, $4.0\times10^{24}~{\rm m}$),
$z$ is the redshift, $\nu$ is the
frequency in GHz, $t_{rise}$ is in days, and $\Delta S$ is in janskys. This gives us $T_{\rm var} =
(5.2\pm0.5)\times10^{12}$K.
The variability brightness temperature is related to the Doppler
factor, $\delta$, as $\delta = (T_{\rm var}/T_{\rm b,int})^{1/3}$, where $T_{\rm
b,int}$ is the intrinsic brightness temperature
\citep[e.g.,][]{lahteenmaki99}. As the variability brightness
temperature estimate is a lower limit, the Doppler factor estimates
are also lower limits. The largest uncertainty in the Doppler factor
estimate comes from the uncertainty in the intrinsic brightness
temperature and what method is used to estimate it. Assuming
equipartition between the particles and magnetic field, $T_{\rm b,int}
= 10^{11}$K \citep{readhead94}. This results in a Doppler factor
$\delta > 3.7$. If we use a value of $T_{\rm b,int} =
5\times10^{10}$~K, as determined by \cite{lahteenmaki99}, the Doppler
factor is $\delta > 4.7$.
We can further constrain the intrinsic brightness temperature by
comparing the variability brightness temperature with the brightness
temperature obtained via simultaneous Very Long Baseline Array (VLBA)
observations \citep{lahteenmaki99, hovatta13}. In order to study the
parsec-scale jet structure after the 2012 flare we conducted 5 epochs
of target-of-opportunity VLBA observations of Mrk~421 at several
frequency bands (J. Richards et al. in preparation). The first one of
these was taken on 12 October 2012 when the radio flare was
already decaying. The brightness temperature estimate from the 15\,GHz
data is $T_{\rm VLBI} = 5.2\times10^{10}$~K (for a uniform disk), which depends on the
Doppler factor as $\delta = T_{\rm VLBI} / T_{\rm b,int}$. We can then
solve for the intrinsic brightness temperature by calculating $T_{\rm
b,int} = \sqrt{T_{\rm VLBI}^3/T_{\rm var}} = 5.2\times10^{9}$~K. If
we use this value for the intrinsic brightness temperature and the
variability brightness temperature, we obtain $\delta > 10$.
Thus, we conclude that the lower limit of the Doppler factor is
$\delta \sim 3-10$.
We note that the intrinsic brightness temperature obtained using
the VLBA data is about $10-20$ times below the equipartition
limit. We think the peak brightness temperature is likely higher than our estimate because
by the time of the first VLBA epoch,
the single-dish flux density had already declined by 30 per cent
from the peak. Therefore it is likely that the true simultaneous
brightness temperature from the VLBA observations is at least 30
per cent higher, because the core could have also been more compact,
increasing the brightness temperature even further. Because of the
strong dependence of $T_{\rm b,int}$ on $T_{\rm VLBI}$, any
uncertainties in the VLBA parameters are magnified in the estimate
of $T_{\rm b,int}$. A slightly higher $T_{\rm VLBI}$ would also
agree with estimates from \cite{lico12}, who found the core
brightness temperature of
Mrk~421 to be of the order of few times $10^{11}$~K, in agreement
with equipartition arguments.
\subsection{The flare of 2013}\label{sect:2013}
In April 2013, Mrk~421 was again flaring in the X-ray to TeV bands \citep{balokovic13,paneque13},
reaching the highest levels ever observed at TeV energies \citep{cortina13}.
Triggered by this activity, we began monitoring the source more
frequently at CARMA.
The appearance of this flare was very different from the 2012
flare, both in the $\gamma$ rays and radio. In $\gamma$ rays the
activity began in March 2013 reaching the highest peak on
14 April 2013 (MJD 56397).
The flux
increased from about $2\times10^{-7}$\,ph\,cm$^{-2}$\,s$^{-1}$ to
($6.7 \pm 0.9)\times10^{-7}$\,ph\,cm$^{-2}$\,s$^{-1}$ over about 30~d and the flaring
period consisted of several peaks over a total duration of about 100~d. At 15\,GHz, the flux density began increasing in April 2013 from
about 0.6~Jy to its peak of 0.77~Jy on 18 June 2013 (MJD
56461). Similarly, at 95\,GHz the flux density increased from about
0.5~Jy to a peak of 0.71~Jy on 8 June 2013 (MJD 56451).
The DCFs between all the bands are shown in
Fig.~\ref{fig:cc2}. The highest correlation is found for the 15
and 95\,GHz data, with a peak at a delay of about $-10$ to $-20$~d, with 95\,GHz leading. The time delays between the
CARMA (OVRO) and LAT data can only be estimated at about $-60$~d
($-70$~d) because of the broad peaks of the DCF. These estimates
are, however, compatible with the broad peak obtained from our
cross-correlation analysis of the full light curves, and thus we
assume that the events are physically connected in the radio and
$\gamma$-ray bands.
\begin{figure*}
\includegraphics[scale=0.6]{f3.eps}
\caption{Top: Light curves at $\gamma$ rays (left), 95\,GHz (middle),
and 15\,GHz (right) during the 2013 flare. Bottom: Discrete correlation functions between the
light curves. Negative time delay means that higher frequency
emission ($\gamma$-ray or 95\,GHz) is leading. The time range
corresponds to the dashed lines in Fig.~\ref{fig:lc}.}\label{fig:cc2}
\end{figure*}
\section{Conditions in the flaring region}\label{sect:model}
In this section we attempt to explain the rough features of the 2012
and 2013 flaring periods, e.g., time delays between various energy bands, peak fluxes and pulse profiles,
by adopting the simplest possible theoretical framework (the single zone
SSC model), together with a minimum set of
free parameters.
In particular, we restrict our modeling to the
major flares observed in the periods of MJD 56060-56225 and MJD
56360-56610, which are denoted as ``flare 1'' and ``flare 2'''
in Sect.~\ref{sect:lc} (see Fig.~\ref{fig:lc}), while modeling
of the smaller amplitude variability seen in both bands lies out of the scope of this work.
As the `goodness' of the fits was not the main point of interest here, we did not attempt
a detailed parameter space search for finding the set with the best
$\chi^2$ value.
Based on the long-term radio light curve shown in Fig.~\ref{fig:lc},
the 2013 radio and millimeter-band flares are more typical of blazar radio emission than the 2012 radio flare:
they are less sharp, have a decay timescale larger than their rise timescale, and
the flux density increases by no more than a factor of $\sim 1.4$ in
both the radio and millimeter bands. We start our analysis with the 2013 flaring events in the context of a typical SSC
model. Then, we highlight the differences between the 2012 and 2013 radio flares and discuss possible modifications within
the same framework that may explain the 2012 data.
Our main goal is to
investigate what conditions are required to produce this extreme and unique radio
flare.
As already discussed in Sect.~\ref{sect:lc}, we assume that the events in the
radio and $\gamma$-ray bands are physically connected.
We note that, in what follows, we ignore the small differences between quantities measured in the observer frame and the source frame
since $1+z \simeq 1$.
\subsection{The $\gamma$-ray and radio flares of 2013}
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{f4a.eps}
\includegraphics[width=0.45\textwidth]{f4b.eps}
\includegraphics[width=0.45\textwidth]{f4c.eps}
\caption{{\sl Fermi}-LAT light curve (top), CARMA 95~GHz light
curve (middle) and OVRO 15~GHz light curve (bottom).
The results of the one-zone SSC model described in the text are shown
with blue lines in all panels. The time range of the fit corresponds
to the 2013 flare indicated by the dashed lines in Fig.~\ref{fig:lc}.}
\label{lc}
\end{figure}
Motivated by the fact that the CARMA and OVRO fluxes peak $\sim 60$
and 70~d after the $\gamma$-ray flare, respectively
(Sect.~\ref{sect:2013}), and by the fact that the decaying part of
their light curve is approximately exponential, we applied the
simple scenario where both the $\gamma$-ray and radio flares originate
in the same region, which does not necessarily have the
same physical conditions as the region responsible for the quiescent
emission. We note that we do not attempt to model the
sub-structure of the flares, but we only consider the highest
peak in each band. In particular, we focus on
the largest $\gamma$-ray flare, which peaks at MJD
$\sim$56400. We also note that the DCF only shows a
single peak for this flare, unlike in 2012 where we will
consider multiple associations between flares.
In our scenario the $\gamma$-ray flare is produced
by an instantaneous injection event of electrons having Lorentz factor
$\gamma_0$ and emitting in $\gamma$ rays through Compton scattering of
synchrotron photons (SSC). In particular, electrons are injected at
$\tau=0$, which is set to be the time of the $\gamma$-ray flare, and
then they are left to evolve via synchrotron and SSC
cooling. In this context, the observed delays $t_{\rm h} \sim 60$~d
and $t_{\ell} \sim 70$~d between the $\gamma$-ray and the radio
flares at $\nu_{\rm h}=95$~GHz and $\nu_{\ell}=15$~GHz, respectively,
correspond to the cooling timescale of electrons that have been
injected at $\tau=0$ with Lorentz factor $\gamma_0$.
\subsubsection{Analytical estimates}
To define the size $R$ of the emitting region we choose
as a typical variability timescale ($t_{\rm v}$) the one dictated by
the $\gamma$-ray light curve.
For the purpose of our analysis, we
choose a variability timescale of $t_{\rm v}=20$~d based on the light
curve. We note that the exact value is not critical as the estimate
will be refined in the next section where we model the flares numerically.
We find that $R \approx 2 \times 10^{17} \ {\rm cm}\ \left(\delta/4 \right) \left(t_{\rm v}/ 20 \ {\rm d}\right)$,
where we normalized the Doppler factor to 4 (see Sect.~\ref{sect:2012}).
In this framework, one can derive the magnetic field strength $B$ of the emission region as a function
of the radio frequency $\nu_{\rm h}$, the observed time delay $t_{\rm h}$ and the Doppler factor $\delta$.
Electrons that have been injected at $\tau=0$ with Lorentz factor $\gamma_0$ cool due to synchrotron losses
and reach a Lorentz factor $\gamma_{\rm h}$. This can be found by solving the characteristic equation for synchrotron cooling
and is given by
\begin{eqnarray}
\gamma_{\rm h} = \frac{1}{1/\gamma_0+ \alpha \delta t_{\rm h}} \approx \left(\alpha \delta t_{\rm h} \right)^{-1},
\label{gamma}
\end{eqnarray}
where $\alpha = (4/3)\sigma_{\rm T} c U_{\rm B}/ m_{\rm e} c^2$ and
$U_{\rm B}=B^2/8\pi$ is the energy density of the magnetic field. The
approximation holds as long as $\gamma_0 \gg \gamma_{\rm h}$.
Combining Eq.~(\ref{gamma}) with the characteristic synchrotron
frequency $\nu_{\rm h}=\delta m_{\rm e} c^2 b \gamma_{\rm h}^2/ h$,
where $b=B/B_{\rm cr}$, $B_{\rm cr}=4.4 \times 10^{13}$~G, and
$h=6.63\times10^{-27}$~erg s is the Planck constant, we find that the required magnetic field strength of the region is
\begin{eqnarray}
B \simeq 0.5 \ {\rm G} \left(\frac{\nu_{\rm h}}{95 \ {\rm GHz}}\right)^{-1/3} \left(\frac{\delta} {4}\right)^{-1/3}
\left(\frac{t_{\rm h}}{60\ {\rm d}}\right)^{-2/3},
\label{B-delay}
\end{eqnarray}
which depends most strongly on the time delay between the $\gamma$-ray and
radio flares. Notice also that lower values of the Doppler factor favour stronger magnetic fields. The expected time delays for the two frequencies should satisfy the relation
$t_{\ell}/t_{\rm h}=(\nu_{\rm h}/ \nu_{\ell})^{1/2}$.
This is roughly consistent with the observed values, since
$\sqrt{\nu_{\rm h}/ \nu_{\ell}}=2.5$ and $t_{\ell}/t_{\rm h}$ is in
the range of $1 - 1.7$ based on the broad peaks in the time
delays (see Fig.~\ref{fig:cc2}).
The duration of a flare is also related to the magnetic field strength, since it depends on the electron cooling timescale.
For a given frequency $\nu_{\rm h} \propto B \gamma^2_{\rm h}$ the full width at half maximum (FWHM) of a flare
will scale roughly as $\propto B^{-2} \gamma_{\rm h}^{-1} \propto B^{-3/2} \nu_{\rm h}^{-1/2}$.
Thus, for strong enough magnetic fields the duration
of a flare even at low frequencies may be shortened significantly.
Assuming that the inverse Compton scattering occurs in the Thomson
limit, which will be checked {\sl a posteriori}, the typical
energy of up-scattered photons is given by $E_{\gamma} \simeq
(4/3)\delta m_{\rm e} c^2 b \gamma_0^4$. Using the value of $B$ derived above,
we find that the injection Lorentz factor of electrons is
$\gamma_0 \simeq 2\times 10^{4} \left(E_{\gamma}/ 4 \ {\rm GeV}\right)^{1/4} \left(\delta/4 \right)^{-1/4} \left(B / 0.5 \ {\rm G}\right)^{-1/4}$,
where we chose $E_{\gamma}=4$~GeV as a representative value for the
energy of $\gamma$-ray photons.
Because of the weak dependence of
$\gamma_0$ on $E_{\gamma}$, a different choice of the $\gamma$-ray energy would
not affect the derived value for the injection Lorentz factor. Increasing $E_{\gamma}$ by a factor of 50, for example, would increase $\gamma_0$ by only a factor of $\sim 2$.
In principle, electrons could have been injected with $\gamma > \gamma_0$
and still produce the GeV flare, since they would very quickly cool to the
value $\gamma_0$. In this case a
contemporaneous TeV flare would also be expected, and indeed one was seen by
the MAGIC and VERITAS instruments \citep{cortina13}.
Having derived the expression for $\gamma_0$, which does not depend strongly on the parameters,
we can now verify our earlier assumption that the Thomson limit applies. Since $b \gamma_0^3 = 0.09 \left(\frac{B}{0.5 \ {\rm G}}\right)^{1/4}
\left(\frac{E_{\gamma}}{4 \ {\rm GeV}} \right)^{3/4} \left(\frac{\delta}{4} \right)^{-3/4} < 3/4$, we are safely in the Thomson regime.
The derived values for the magnetic field strength and the Lorentz factor of injected
electrons are reasonable and consistent with typical SSC models of
high frequency peaked BL~Lac objects, such as Mrk~421 \citep[e.g.,][]{abdo11,aleksic12}. Note, however, that
here we adopted a low value for the Doppler factor implied by radio observations, in contrast to typical
SED modeling where much larger values, e.g. $20-50$, are usually used
\citep[see e.g.,][]{maraschi99}
\subsubsection{Numerical results}
The required electron injection luminosity is determined numerically
by fitting the observed energy fluxes. For this, we employed
the numerical code described in \cite{mastichiadis95, mastichiadis97}. Using as a stepping stone
the values determined analytically in Sect. 4.1.1., we derive the following set of parameters:
$R=10^{17}$~cm, $B=0.1$~G, $\delta=2.2$, $\gamma_0=2\times 10^4$ and the electron injection compactness $\ell_{\rm e}^{\rm inj}=1.6\times 10^{-2}$. This is
defined as $\ell_{\rm e}^{\rm inj} = \sigma_{\rm T} R U_{\rm e}/ m_{\rm e}c^2$ where $U_{\rm e}$ is the energy density of electrons
at injection time as measured in the rest frame of the emission region and $\sigma_T=6.65\times10^{-25}$~cm$^2$ is the Thomson cross section.
Although our analytical estimates were based upon the hypothesis
of instantaneous injection, here we find that an injection episode lasting $\sim R/c$, or equivalently $20$~d,
is better in reproducing the observed flares.
Our results are illustrated in Fig.~\ref{lc}, where we
use energy fluxes to facilitate comparison between the different
bands. In the radio band, flux densities are converted to
energy fluxes by multiplying by the observing frequency and
converting to cgs units. The energy fluxes in the
$\gamma$-ray band are obtained as
described in Sect. 2. The model light curves can describe
the radio observations fairly well, although the model
$\gamma$-ray flare is slightly broader than the actual
data. Still, the results of Fig.~\ref{lc} are
satisfactory given the small number of free parameters and the fact
that we have not attempted to find the best fit (i.e., with the
lowest $\chi^2$ value). The radio flares in both frequencies have
wide pulse profiles and exhibit a slow exponential decay after their
peak. Both of these features contrast with the sharpness
and the symmetry of the 2012 flare. In the
present scenario, the width of each flare is related to the
cooling timescale of electrons emitting at the particular energy band
and, thus, the radio flares are wider than the one in $\gamma$ rays
(see Sect. 4.1.1.). The asymmetry of the pulse
profiles at lower energies is a strong prediction of this
model. Possible expansion of the flaring region and/or
decay of the magnetic field would increase the predicted
asymmetry.
For the derived parameters, the region is initially\footnote{The
electron energy density will actually decrease with time because of
(i) cooling and (ii) no replenishment of particles in the region.}
particle dominated with $U_{\rm e}\approx0.18$~erg cm$^{-3}$ and
$U_{\rm B}=4\times 10^{-4}$~erg cm$^{-3}$. The value of the Doppler
factor derived here is approximately half of that obtained in
Sect.~\ref{sect:2012} assuming equipartition. However, the large ratio $U_{\rm e}/U_{\rm B} \simeq
450$ derived from the values mentioned above would
also imply a higher brightness temperature \citep[e.g.,][]{readhead94}
by a factor of $450^{1/8}\sim 2$ and thus a lower value of $\delta$ by
a factor of $2^{-1/3} \sim 0.7$. Given that the fit shown in Fig.~\ref{lc} is not unique, one could
find parameters that would decrease the initial ratio $U_{\rm
e}/U_{\rm B}$ and bring the emission region closer to an
equipartition state. By assuming that a fraction $\eta_{\rm rad}$ of
the injected luminosity in electrons is radiated as $\gamma$ rays,
i.e. $\eta_{\rm rad} \ell_{\rm e}^{\rm inj} = L_{\gamma}^{\rm obs}
\sigma_{\rm T}/4 \pi R \delta^4 m_{\rm e}c^3$, and using the
definition of $\ell_{\rm e}^{\rm inj}$ as well as the equations for
$R$ and $B$ in Sect.~4.1.1., we find that
\begin{eqnarray} \frac{U_{\rm e}}{U_{\rm
B}} \simeq F_0
\left(\frac{\eta_{\rm
rad}}{0.1} \right)^{-1}\!\! \left(\frac{\delta}{4}
\right)^{-16/3} \!\! \left( \frac{\nu_{\rm h}}{95 \ {\rm
GHz}}\right)^{2/3} \left(\frac{t_{\rm h}}{60 \ {\rm d}}
\right)^{4/3} \left( \frac{t_{\rm v}}{20 \ {\rm d}}\right)^{-2}, \end{eqnarray}
where $F_0 \equiv F_{\gamma}^{\rm obs} / 2\times 10^{-9}$~erg
cm$^{-2}$ s$^{-1}$. The ratio of energy densities depends
strongly on the Doppler factor. Thus, searching for possible fits
with a slightly higher value of the Doppler factor, e.g., $\delta
\simeq 7$, would bring down the ratio from 450 to close to unity and
would ensure rough equipartition between the particles and magnetic
field at injection.
\subsection{The $\gamma$-ray and radio flares of 2012}
\begin{figure*}
\centering
\includegraphics[width=0.45\textwidth]{f5a.eps}
\includegraphics[width=0.45\textwidth]{f5b.eps}
\caption{{\sl Fermi}-LAT light curve (left) and OVRO 15~GHz light
curve (right) in energy flux units along with the light curves (blue lines) obtained
from the model. Solid and dashed lines
demonstrate the model results, if the radio flare is associated with the first and second spikes of the $\gamma$-ray
flare, respectively. A different x-axis range is used in the right panel to show the extended model light curve more clearly.
}
\label{2012}
\end{figure*}
The 2012 radio flare is a unique event not only because of its large
flux increase but also because of its symmetric pulse profile, which
resembles flares at higher energies where the corresponding
radiating particles cool efficiently. According to the
cross-correlation analysis presented in Sect. 3.1, the 2012 radio
flare may be associated with either the first or second spike of the
$\gamma$-ray flare (see Figs.~1 and 2). We investigate both
possibilities under the assumption that there is indeed a physical
connection between the extreme radio and $\gamma$-ray flares. For
this we begin with the parameters we derived for the 2013
``typical'' flare, and examine two alternative scenarios. These
include changes of the Doppler factor and magnetic field strength,
since both parameters are related to the time delay (see
Eq.~\ref{B-delay}) and to the pulse profile (see Sect. 4.1.1.). We
emphasize that in both scenarios we attempt to introduce minimal
changes to the parameters. As before, we do not consider
the small-amplitude sub-structure of the flares.
\subsubsection{Association of the radio flare with the first
$\gamma$-ray spike} We first investigate the
possibility that the radio flare is associated with the
narrower first spike of the $\gamma$-ray flare, which peaks at
$\sim$56133 MJD. In order to do this, we
use the same parameter set as derived for the 2013 flare, but
add variations in the Doppler beaming to the
model. Such changes could occur if the emission region moves on a
curved trajectory with a changing viewing angle, or if the bulk
Lorentz factor of the emission region changes. Varying Doppler factors
have been previously used to explain radio spectral variations in, for
example, S5~0716+714 \citep{rani13}. The Doppler factor $\delta$ is
modeled as \begin{eqnarray} \delta(\tau) =\delta_0 \left(1+ g (\tau)
\cos(2\pi \tau / P) \right), \end{eqnarray} where $\delta_0=2.1$, $\tau$ is
the time in the comoving frame in $R/c$ units\footnote{Time is
measured with respect to the time of the first spike seen in the
2012 $\gamma$-ray flare.}, $P=5.4$ is the period of the
variation in $R/c$ units and $g (\tau)$ is the function
\begin{eqnarray}
g (\tau) & = & 0.3, \ {\rm for} \ \tau < \tau_{\rm br} \\
g (\tau) & = & 0.3 e^{-(\tau-\tau_{\rm br})/T} , \ {\rm for}
\ \tau \ge \tau_{\rm br}, \end{eqnarray} with $\tau_{\rm br}=1.2P$
and $T=0.2 P$.
The model-derived pulse profiles are shown in Fig.~\ref{2012} with blue solid lines.
Although the cooling timescale of electrons emitting at radio
frequencies is long compared to those emitting initially at $\gamma$
rays, the shape of the radio flare (right panel in Fig.~\ref{2012}) is
well reproduced here, because of the change in the Doppler factor,
which does not allow the observer to see the typical wide pulse
profile (see e.g., Fig.~\ref{lc}). The effect of the variable
$\delta$ on the $\gamma$-ray light curve is
negligible because the radiating electrons at the time of
injection have short cooling timescales. Note that the exponentially
decaying amplitude in the variation of $\delta$ is required in order
to avoid any excess of the radio emission following the extreme flare.
Assuming that the change in the Doppler factor is caused by small variations
in the viewing angle $\theta$ while the Lorentz factor
remains constant at $\Gamma=10$, no extreme variations
in either $\delta$ or the angle $\theta = \cos^{-1} \left(\beta^{-1}
\left(1-(\Gamma \delta)^{-1} \right) \right)$ are
required.\footnote{This Lorentz factor is approximately five times
higher than the value derived in Sect.~\ref{sect:discussion} for
the bulk Lorentz factor of the jet ($\Gamma_{\rm var}=1.9$). If
we had used this value instead, only the angle $\theta$ would be
affected, i.e. it would be larger by a factor of 2. However, it
is still possible that the Lorentz factor of the emission region
is higher than that of the jet \citep[see
e.g.,][]{nalewajko14}.}
In particular, $\delta$ and $\theta$
change by less than $28$ and $22$ per cent, respectively, relative to their
average values.
The fact that the necessary parameter changes are small
make this scenario plausible and attractive.
However, if we attempt to interpret the second spike of the
$\gamma$-ray flare in the same framework, then we cannot avoid the
appearance of a radio flare $\sim$~70~d later,
i.e. at MJD~56220. Its absence implies one (or more) of
the following: (i) a faster onset of the exponential decay of
$\delta$; (ii) a different functional form for $\delta(\tau)$; (iii)
different conditions in the emission region which suppress
the 15~GHz emission, such as different magnetic field and/or
size; (iv) a different framework for the second $\gamma$-ray spike.
Finally, the choice of a damped oscillator for modeling $\delta(\tau)$, albeit plausible,
was not based on a physical picture.
According to the discussion of Sect. 4.1.1., the FWHM of a flare depends on the magnetic
field strength. Therefore, another alternative might be to adopt
a stronger magnetic field for the emission region.
Although a stronger magnetic field would
lead to a shorter duration of the flare, an even lower value of the Doppler factor would be required in order
to explain the observed time delay between the $\gamma$-ray and radio flares (see Eq.~\ref{B-delay}).
Given that the value of the Doppler factor already lies at the low end of values consistent with observations, we do not examine this possibility in more detail.
\subsubsection{Association of the radio flare with the second $\gamma$-ray spike}
We now consider the scenario where the radio
flare is associated with the second spike of the $\gamma$-ray flare.
This scenario would arise naturally if the first $\gamma$-ray spike occurred
when the emission region was still optically thick to radio emission.
In this case, the time-delay between the $\gamma$-ray and radio flares is shorter than before, which, in our framework, suggests
faster electron cooling.
We searched, therefore, for reasonable fits using a stronger magnetic field than
that adopted in Sect. 4.2.1, and obtained the
following parameter set: $B=0.25$~G, $\delta=2.3$ and $\ell_{\rm e}^{\rm inj}=1.3\times 10^{-2}$. All other parameters, including the injection profile, are the same as before.
The resulting pulse profiles are shown in Fig.~\ref{2012}
with dashed blue lines. We find that the light curves obtained by the
model are in rough agreement with the observations, with the radio
light curve having a slightly longer decay timescale than is
observed. In this scenario, even without Doppler factor
variations, we obtain a sharper radio flare than
the one in 2013 (see e.g. Fig.~4). Comparing these parameter
values with the first scenario, we see that although we
adjusted the numerical values for three parameters, only the
magnetic field is significantly altered. It has increased by a
factor of 2.5. We can compare the FWHM of the 15~GHz radio flares in
2012 and 2013 as obtained by our model (Figs.~\ref{lc} and
\ref{2012}). We find their ratio to be $\Delta T_{2012}/\Delta
T_{2013} \sim 0.16$, where $\Delta T$ denotes the duration at
FWHM. This is comparable to the analytical
estimate we can obtain from Eq.~\ref{B-delay}, $\Delta T_{2012}/\Delta T_{2013} \simeq \left(B_{2012}/B_{2013}\right)^{-3/2} \simeq
\left(0.25/0.1\right)^{-3/2}= 0.25$.
We note that this model is subject to tight constraints on
its parameters. These must be adjusted to ensure: (i) the electron
cooling timescale results in the observed delay between the
$\gamma$-ray spike and the radio flare (see Eq.~(\ref{B-delay})); (ii)
the cooling timescale of electrons radiating at the radio frequency
is such that it explains the observed width of the pulse; and
(iii) the synchrotron self-absorption frequency, which depends on
$\ell_{\rm e}^{\rm inj}$, $B$ and $R$, is below the radio
frequency where the extreme flare is observed. From the above it
becomes clear that fine tuning of the parameters is required for this
scenario to be viable.
\section{Discussion}\label{sect:discussion}
Unlike in many other blazars, prominent radio flares are rare in Mrk~421 and there are few
examples where the radio flares have been modeled in detail. In 1997 a
22\,GHz flare was observed with the Mets\"ahovi Radio Observatory 14-m
telescope \citep{tosti98}. The flare lasted about 200~d during
which the flux density increased by a factor of two. The radio flare occurred
60~d after a large optical flare that also included very high
optical polarization. Due to the high optical polarization during the
flare, the authors interpreted the flare within the shock-in-jet model
of \cite{marscher85}. The delay between the optical and radio bands was
explained by opacity effects. While this model may explain
the general features of the flares and the delay between optical and
radio bands, more detailed modeling was not presented by \cite{tosti98}.
In 2001 a simultaneous flare was observed at TeV, X-ray and 5\,GHz
radio bands \citep{katarzynski03}. The radio flux density increased by
a factor of 1.5 and the flare lasted only for 15~d. The absence of
time delays between the bands indicates that the emission region was
optically thin even at radio frequencies. An instant injection of particles into the
jet base was unable to explain the radio flares as the
electrons would have cooled too much to generate significant flux
density changes before reaching the optically thin regime for radio
emission. Instead,
detailed modeling of the
flare suggested that it was caused by in-situ acceleration of electrons in a
blob in the jet. The expansion of the blob explains the decay of the
flare at all bands.
The 2013 flaring event was a more typical example of radio flares in
Mrk~421, with a broad flare in the millimeter and centimeter bands
(see Fig.~\ref{fig:lc} for the long-term behavior). Assuming that the
events in the different bands are physically connected, we modeled the
most prominent features of this event with a simple one-zone SSC model
where an instantaneous injection of electrons produces a $\gamma$-ray
flare. Then, as electrons cool because of synchrotron and SSC energy
losses, they radiate at longer wavelengths and eventually produce the
radio flares with a delay relative to the $\gamma$-ray flare. Within
this context, we estimated the required magnetic field strength, the
size of the emission region, and the Lorentz factor of electrons at
injection. While the simple model presented here can describe the
main features of the data satisfactorily, more information from
simultaneous multi-wavelength observations is needed in order to
discriminate between other possible models. For example, one could
postulate a scenario where the delay between the $\gamma$-ray and
radio flares is related to opacity effects, such as if the source were
optically thick in radio frequencies at the time of the GeV flare but
eventually became optically thin due to the expansion of the emission
region or decay of the magnetic field \citep[e.g.,][]{fuhrmann14}.
Although the coincident unprecedented events between the LAT
and OVRO data suggests a physical connection of the high- and
low-energy emission, models where two spatially separated emission
regions are invoked \citep{blazejowski05, petropoulou14} are still
viable alternatives. We note that the 2013 flare occurred during a
planned multi-wavelength campaign and Mrk~421 was observed regularly
with numerous instruments at various frequency bands. Therefore we
anticipate several dedicated studies of the behavior of Mrk~421 during
this flare.
Even though the 2012 radio flare was more extreme, the
Doppler factor inferred from the radio variability time-scale is
fairly low, only about $3-10$. This is in accordance with the low observed
apparent speeds obtained through radio interferometric observations
where only subluminal speeds
have been detected \citep[e.g.,][]{piner99,lico12, lister13}. Based on
the jet brightness asymmetry, low observed apparent speeds, and low
brightness temperature of the core component, \cite{lico12} estimate
the radio Doppler beaming factor to be about 3, consistent with our
estimate. If simultaneous X-ray or TeV data were available, we could
estimate a minimum Doppler factor by demanding the source to be optically thin to
the highest energy emitted photon
\citep[e.g.,][]{dondi95}. From previous TeV flares of Mrk~421 it has
been estimated that the minimum Doppler factor required for the TeV
emission to be optically thin is $\sim 10$
\citep{celotti98}, which is marginally higher than the value obtained
from our radio data when equipartition is assumed.
Preliminary analysis of our follow-up VLBA observations indicate that
the component speeds were consistent with subluminal motion even after
the major flare \citep{richards13}. If we take the fastest observed
jet speed, $\beta_\mathrm{app}=0.28c$, obtained
from several years of monitoring at 15\,GHz with
the VLBA~\citep{lister13}, we can estimate the jet Lorentz factor as
\begin{equation}
\Gamma_\mathrm{var} = \frac{\beta_\mathrm{app}^2 +
\delta_\mathrm{var}^2 + 1}{2\delta_\mathrm{var}},
\end{equation}
where $\delta_\mathrm{var}$ is the Doppler factor inferred from the
variability time-scales. This results in a Lorentz factor of
$\Gamma_\mathrm{var} \sim 1.7-5$. Similarly, we can estimate the viewing
angle of the jet to be
\begin{equation}
\theta_\mathrm{var}=
\arctan\frac{2\beta_\mathrm{app}}{\beta_\mathrm{app}^2 +
\delta_\mathrm{var}^2 - 1},
\end{equation}
resulting in $\theta_\mathrm{var} \sim 0.3-4^\circ$. As
noted by \cite{lico12}, it is unlikely that the component speeds
resemble the flow speed in Mrk~421 because unreasonably small
viewing angles would be required to explain the beaming estimates
from the jet/counter-jet ratio. In this case the above equations
would not be valid for estimating the true Lorentz factor and
viewing angles. However, they find a viable scenario for
Mrk~421 with a structured jet, where the viewing angle is between
$2^\circ$ and $5^\circ$, and the Lorentz factor of the radio
emitting region about 1.8, in agreement with our estimates.
One of the long-standing problems in modeling of the high synchrotron
peaked blazars is the large discrepancy
between the Doppler factor values inferred from radio observations and those
obtained by SED modeling, with the former being usually $\lesssim 10$
\citep[e.g.,][]{piner99,lister13} and the latter
lying in the range $\sim 20-50$ \citep[e.g.,][]{maraschi99,abdo11}.
Several alternatives have been explored to explain
this well-known ``Doppler factor crisis'', such as a structured jet
\citep{ghisellini05}, a decelerating jet \citep{georganopoulos03},
and a jet-in-jet model \citep{giannios09}. In the structured jet model
the high-energy emission comes from a fast spine of the jet while the
radio emission is produced in a slower sheath
\citep{ghisellini05}. This model is favored by observations of limb
brightening of the Mrk~421 jet at 43\,GHz \citep{piner10}.
In the present work, where we have adopted a single-zone emission model,
we attempted to avoid a large discrepancy between the Doppler factor values
inferred from the observations and those
used in the modeling of the flares.
However, one could relax this condition and search for possible fits to both the radio and $\gamma$-ray flares
with $\delta \gtrsim 20$. One can estimate the effect of a higher $\delta$ on
our results by inspection of the analytical expressions derived in
Sect. 4.1.1. In particular,
we find that $R\propto \delta$, $B\propto \delta^{-1/3}$, $\gamma_0 \propto \delta^{-1/6}$, $U_{\rm e}/ U_{\rm B}\propto \delta^{-16/3}$
and FWHM$\propto \delta^{1/2}$, where we assumed that the radio observing frequencies and their time-delays
with respect to the $\gamma$-ray flare are fixed. On the one hand, choice of a higher $\delta$ would require
a weaker magnetic field, a larger emission region and would reduce the ratio of particle
to magnetic energy densities (see discussion in Sect. 4.1.2).
On the other hand, the model light curves would be wider than those presented in Fig.~4, since the FWHM would increase. This is a direct
result of the longer electron cooling timescale due to the weaker magnetic field.
We conclude, therefore, that values $\gtrsim 20$ are less plausible for the modeling of both the radio and $\gamma$-ray flares.
Another possibility is to assume a high Doppler factor value for the $\gamma$-ray flare alone and
a low value for the radio flares.
Because a single-zone model only contains a single Doppler factor,
this parameter would need to change rapidly between the two flares.
This scenario would require the Doppler factor to drop by $\gtrsim
10$ between the $\gamma$-ray and radio flare within a period
of $t_{\rm h}=60$~d, which is much larger than the modest variations
applied in the modeling of the 2012 flares.
The 2012 $\gamma$-ray flare had two prominent spikes, and we used
these to investigate the physical conditions required to produce the
extreme radio event. By using the 2013 model parameters as a
starting point and introducing as few modifications as possible, we
considered two possible scenarios. These result
from the association of the radio flare with the first or second
$\gamma$-ray spikes. Under the assumption of a physical connection
between the radio flare and the first $\gamma$-ray spike, we showed
that the addition of a varying Doppler beaming factor
to a one-zone SSC model with fairly typical parameters can
explain the observed sharp radio flare. The SSC model would
otherwise produce too broad a radio flare, similar to the 2013
event (see Figs.~\ref{lc}-\ref{2012}). Although this
scenario succeeds in explaining the radio flare, it results in a
wider $\gamma$-ray pulse profile than is observed (solid
lines in Fig.~\ref{2012}). Moreover, it is difficult to
explain the lack of a second peak in the radio light curve, as
discussed in the end of Sect.~4.2.1. Nonetheless, a
varying Doppler beaming factor is not unreasonable. An adequate
variation could result from, e.g., a modest change in the Lorentz
factor or the viewing angle. In the latter case, the sharp flare
would occur when the direction of motion of the emission
region crosses very near to the line of sight to the observer. Similar models with
curved emission region trajectories have been suggested for other
blazars as well
\citep[e.g.,][]{marscher08,marscher10,abdo10,rani13,aleksic14,molina14}.
These models assume that either the emission feature moves on a
helical trajectory due to an ordered helical magnetic field
\citep{marscher08,marscher10,molina14}, or that there is an actual
bend in the jet \citep{abdo10,aleksic14}.
In our second scenario, we considered the possibility
that the radio flare is associated with the second,
broader $\gamma$-ray spike. In this case, the short duration of
the radio flare and the shorter time delay between the
radio and $\gamma$-ray flares imply faster electron cooling, and
thus a stronger magnetic field than the one used in the
first scenario. We found indeed that a viable model is achieved by
increasing the magnetic field strength by a factor of 2.5.
Fig.~\ref{2012} shows that the model light curve (dashed lines)
describes well the $\gamma$-ray data, although the
modeled radio flare is now not quite as sharp as the observed one.
In this scenario, the first $\gamma$-ray spike would have to be
unassociated with the radio event, perhaps by occurring closer to
the black hole where the emission region is optically thick for
radio emission.
\section{Conclusions}\label{sect:conclusions}
We have obtained radio, millimeter and $\gamma$-ray light curves of
Mrk~421 during the flaring activity in 2012 and 2013. In July 2012,
Mrk~421 exhibited the largest $\gamma$-ray flare observed by {\it
Fermi} since the beginning of its mission. About
$40-70$~d later, the largest ever 15\,GHz flare was
observed. The flare rise time determined from an exponential fit was
just $10.6\pm0.5$~d, which is extreme compared to
previous radio flares observed in the source. This
implies a variability Doppler factor
$\delta\sim3-10$. In 2013 Mrk~421 underwent major
$\gamma$-ray flaring again, followed by radio and millimeter flares
about 60~d later. Under the assumption that the
events in the different bands are physically connected we
have modeled the variations with the simplest possible theoretical
model, with the main goal of explaining the extreme radio flare in
2012. Starting with the less extreme 2013 flare, we
obtained a one-zone SSC model that could explain the main
features of the flares reasonably well.
The 2012 $\gamma$-ray flare was double peaked, and by
modeling the radio flare as connected to each peak in
turn, we were able to reproduce the radio flare with either a
varying Doppler factor, or an increased magnetic field strength. In
both cases several specific conditions in the jet need to be
fulfilled for the models to be viable, showing the extreme and unique
nature of the 2012 event.
\section*{Acknowledgments}
We thank Bindu Rani, Jeremy Perkins, Dave Thompson, and the anonymous
referee for their comments and suggestions that greatly improved the paper.
T.\,H. was supported by the Academy of Finland project number 267324.
M.\,P.: Support for this work was provided by NASA
through Einstein Postdoctoral
Fellowship grant number PF3 140113 awarded by the Chandra X-ray
Center, which is operated by the Smithsonian Astrophysical Observatory
for NASA under contract NAS8-03060.
M.\,B. acknowledges support from the International Fulbright Science and Technology Award.
The OVRO 40-m monitoring program is supported in part by NASA grants
NNX08AW31G and NNX11A043G, and NSF grants AST-0808050 and
AST-1109911. The National Radio Astronomy Observatory is a facility of
the National Science Foundation operated under cooperative agreement
by Associated Universities Inc. Support for CARMA construction was
derived from the states of California, Illinois, and Maryland, the
James S. McDonnell Foundation, the Gordon and Betty Moore Foundation,
the Kenneth T. and Eileen L. Norris Foundation, the University of
Chicago, the Associates of the California Institute of Technology, and
the National Science Foundation. Ongoing CARMA development and
operations are supported by the National Science Foundation under a
cooperative agreement, and by the CARMA partner universities. The \textit{Fermi} LAT Collaboration acknowledges generous ongoing support
from a number of agencies and institutes that have supported both the
development and the operation of the LAT as well as scientific data analysis.
These include the National Aeronautics and Space Administration and the
Department of Energy in the United States, the Commissariat \`a l'Energie Atomique
and the Centre National de la Recherche Scientifique / Institut National de Physique
Nucl\'eaire et de Physique des Particules in France, the Agenzia Spaziale Italiana
and the Istituto Nazionale di Fisica Nucleare in Italy, the Ministry of Education,
Culture, Sports, Science and Technology (MEXT), High Energy Accelerator Research
Organization (KEK) and Japan Aerospace Exploration Agency (JAXA) in Japan, and
the K.~A.~Wallenberg Foundation, the Swedish Research Council and the
Swedish National Space Board in Sweden.
Additional support for science analysis during the operations phase from the following agencies is also gratefully acknowledged: the Istituto Nazionale di Astrofisica in Italy and the Centre National d'\'Etudes Spatiales in France.
\footnotesize{
\bibliographystyle{mn2eb}
|
\section{Introduction}
Due to the unstable nature of the weakly decaying hyperons there are no scattering experiments with hyperons in the initial state. However, they naturally occur in the final state, for instance in baryon-antibaryon pair production via electron-positron annihilation $e^+e^-\longrightarrow \bar B B$, in deeply virtual exclusive meson electroproduction $\gamma^* p \longrightarrow K^+ \Lambda,\ K^+ \Sigma^0,\ K^0 \Sigma^+$, and in decays of heavy quarkonia to baryon-antibaryon pairs like $J/\Psi,\ \Upsilon \longrightarrow \bar BB$. The standard way to parametrize the nonperturbative information contained in such exclusive processes are (transition) generalized parton distributions or ordinary form factors. At high momentum transfer the contributions from Fock states containing more than the minimal number of partons are power-suppressed and the process can be approximated by a convolution of the involved distribution amplitudes (DAs) with the process-dependent hard scattering kernel. The requirement of large momentum transfer, the instability of the final state hadrons and the fact that distribution amplitudes only occur in convolutions require high luminosity and high granularity detectors to extract information on the hyperon DAs from experiment.\par%
Another type of process where hyperon DAs are involved are the exclusive rare decays of $b$-baryons, like $\Xi_b$, $\Lambda_b$, $\Sigma_b$ and $\Omega_b$, into octet baryons (plus $\gamma$, $l^+ l^-$, \dots). Due to the large mass difference one can hope that higher order Fock states are sufficiently suppressed to allow for a description by three-quark DAs. Since the bottom baryons are produced with increasing rates at LHC and at B-factories worldwide, we have to expect that ever more precise experimental results will be available in future, even for rare decays containing flavor-changing neutral currents, which are sensitive to new physics. Notwithstanding the fact that $b$-baryons are produced at much lower rates than $b$-mesons, they are not less interesting since they allow for an examination of the helicity structure of the $b \longrightarrow s$ transition and thus complement the measurements in the meson sector~\cite{Aaij:2013qta}. As shown in refs.~\cite{Aliev:2002tr,Aliev:2002ww} there are possible scenarios where deviations from the standard model are not seen in the branching ratio of $\Lambda_b\longrightarrow\Lambda l^+l^-$ but only in the $\Lambda$ baryon polarization. It is therefore mandatory to establish a theoretical basis for the description of such decays, and the knowledge of hyperon DAs is one important ingredient. Even the higher twist components can yield relevant contributions~\cite{Wang:2008sm}. Note that constraining the shape of wave functions by calculating the moments of the DAs with lattice QCD plays an even more important role for hyperons than for nucleons, since experimental bounds are less strict than in the nucleon sector.\par%
A first parametrization of the leading twist contributions in hyperon wave functions was already presented in ref.~\cite{Chernyak:1987nu}. A complete parametrization (including all contributions from higher twist) of baryon-to-vacuum matrix elements was first performed for the case of the nucleon in ref.~\cite{Braun:2000kw}, where it turned out that higher twist contributions can yield substantial effects in the baryon sector, since the corresponding normalization constants $\lambda^N_1$ and $\lambda^N_2$ are large compared to the leading twist wave function normalization constant $f^N$. The same procedure has later on been reused in refs.~\cite{Liu:2008yg,Liu:2009uc} to give similar parametrizations for matrix elements of the hyperons in the baryon octet, namely $\Sigma^\pm$, $\Sigma^0$, $\Xi^-$, $\Xi^0$ and $\Lambda$. Our work unifies these different approaches and we find relations between the distribution amplitudes for different baryons even if $\operatorname{SU}(3)_f$ symmetry is broken. The obtained relations are exact including terms up to first order in the quark masses. In this sense we call our results model-independent. However, one should keep in mind that higher order contributions which lie beyond the accuracy of our analysis are model-dependent indeed, since they are affected by the neglection of higher order terms during operator construction and by the choice of the regularization scheme.\par%
As shown in refs.~\cite{Liu:2008yg,Liu:2009uc} one has to introduce six additional DAs if one extends the formalism from the nucleon doublet to the complete baryon octet. Our results show that these additional DAs are determined by the eight independent DAs already known from the nucleon sector. I.e., if one knows the eight standard DAs (and their dependence on the mass splitting between light and strange quarks) for the $\Lambda$ and for at least two types of octet baryons with nonzero isospin, one can predict all the rest. Using the parametrization given in refs.~\cite{Anikin:2013aka,Anikin:2013yoa,Braun:2008ia}, where contributions of Wandzura-Wilczek type~\cite{Wandzura:1977qf} are taken into account, and applying the approximation advocated in ref.~\cite{Anikin:2013aka}, where contributions that can mix with four-particle operators are systematically neglected, we need only $43$ parameters to describe the complete set of baryon octet DAs, including their dependence on the splitting between light and strange quark mass. For details see section~\ref{sect_example_of_application}. This amounts to a significant reduction of parameters compared to an ad~hoc linear extrapolation without the knowledge of $\operatorname{SU}(3)_f$ symmetry breaking, which would require $72$ parameters for the given setup. Therefore our results are useful for the extrapolation of lattice data. In a first step it can be checked whether the nontrivial relations between the different DAs that we have obtained are realised in lattice simulations. If this is the case to a satisfactory degree, one can perform a simultaneous fit to all DAs, which, owing to the significant reduction of parameters mentioned above, has much higher precision. Note that the parameters occurring in the approximation described above are determined by the zeroth, first and second moments of the leading twist DAs and by the zeroth and first moments of twist $4$ DAs, which are, apart from the first moments of the higher twist amplitudes, within reach of state of the art lattice simulations (see ref.~\cite{PhysRevD.89.094511}).\par%
Let us note that $\operatorname{SU}(3)_f$ breaking effects can be of considerable strength: e.g.\ sum rule estimations of the symmetry breaking for leading twist wave function normalization constants range from $\sim 10 \%$, see ref.~\cite{Chernyak:1987nu}, to $\sim 50 \%$. The latter is obtained if one takes the values for $f^\Sigma$ and $f^\Xi$ from refs.~\cite{Liu:2008yg,Liu:2009uc}. In ref.~\cite{Chernyak:1987nu} it is stated that the impact on the shape of the wave function is even larger (at a scale of $\unit{1}{\giga\electronvolt}$). One therefore has to expect substantial effects also at intermediate scales which are relevant for phenomenological computations. We expect lattice QCD calculations to provide quantitative results for $\operatorname{SU}(3)_f$ breaking effects at physical mean quark mass in the near future. These will provide the means to determine the strength of the symmetry breaking terms in our formalism, allowing us to draw more definite conclusions.\par%
At this point we want to highlight a conclusion that can be drawn from our results, which is of conceptual importance and also affects the nucleon sector: we find that the nonanalytic chiral behaviour of moments of DAs does not depend on the twist of the amplitude. Instead, the leading chiral logarithms in the chiral-odd sector are determined by the type of amplitude to which the corresponding moment contributes. The ones occurring in $\Phi^B_{+,i}$ ($\Phi^B_{-,i}$) amplitudes, which will be defined in eq.~\eqref{definition_superior_DAs}, have the same chiral logarithms as $f^B$ ($\lambda_1^B$). The odd moments of the leading twist DA therefore behave like $\lambda_1^B$ instead of $f^B$, which is quite contrary to the intuitional expectation. The shape parameters occurring in ref.~\cite{Anikin:2013aka} can all be assigned uniquely to one of the two classes, which means that the destinction between moments described above is to some extent already present in currently used parametrizations.\par%
This work is organized as follows: In section~\ref{sect_fundamental_definitions} we present some fundamental definitions to lay the base for the parametrization of the nonlocal three-quark operators in terms of baryon and meson fields, which is performed in section~\ref{sect_operator_construction}. A sketch of the leading one-loop baryon chiral perturbation theory (BChPT) calculation is given in section~\ref{sect_calculation}, where we also explain how we have matched our results to the standard DAs given in ref.~\cite{Braun:2000kw}. In section~\ref{sect_result} we present our main results. We provide a definition for DAs that do not mix under chiral extrapolation and naturally embed the $\Lambda$ baryon. The result section is to the most part self-contained such that the reader can skip the details of the derivation at will. We summarize in section~\ref{sect_summary}
\section{Fundamental definitions}\label{sect_fundamental_definitions}
There exist various possible realizations of chiral symmetry, which all lead to equal results. In the following we only present the definitions we use in this work. For a detailed treatment of the effective field theory framework we refer to~\cite{Weinberg:1978kz,Gasser:1983yg,Gasser:1984gg,Gasser:1987rb,Krause:1990xc,Bernard:2007zu}. The pseudoscalar fields are contained in%
\begin{align}
u&=\sqrt{U}=\exp \biggl( \frac{i}{2 F_0} \lambda^a \phi^a \biggr) = \exp \biggl( \frac{i}{2 F_0} \phi \biggr) \ ,
\end{align}
where $\lambda^1$, \dots, $\lambda^8$ are Gell-Mann matrices and $F_0$ is the pion decay constant in the three-flavor chiral limit, which corresponds to the convention where $F_\pi = F_0 + \mathcal{O}(m_\pi^2,m_K^2,m_\eta^2) \approx \unit{92}{\mega\electronvolt}$. The matrix $\phi$ can be written in terms of meson fields%
\begin{align}
\phi &= \sqrt{2}
\begin{pmatrix}
\frac{1}{\sqrt{2}} \pi^0 + \frac{1}{\sqrt{6}} \eta & \pi^+ & K^+ \\
\pi^- & -\frac{1}{\sqrt{2}} \pi^0 + \frac{1}{\sqrt{6}} \eta & K^0 \\
K^- & \overbar{K}^0 & -\frac{2}{\sqrt{6}} \eta
\end{pmatrix} \ .
\intertext{The $3 \times 3$ matrix $B$ contains the baryon octet:}
\begin{split}
B &=
\begin{pmatrix}
\frac{1}{\sqrt{2}} \Sigma^0 + \frac{1}{\sqrt{6}} \Lambda & \Sigma^+ & p \\
\Sigma^- & -\frac{1}{\sqrt{2}} \Sigma^0 + \frac{1}{\sqrt{6}} \Lambda & n \\
\Xi^- & \Xi^0 & -\frac{2}{\sqrt{6}} \Lambda
\end{pmatrix} \\
&\equiv \kappa_p p + \kappa_n n + \kappa_{\Sigma^-} \Sigma^- + \kappa_{\Sigma^0} \Sigma^0 + \kappa_{\Sigma^+} \Sigma^+ + \kappa_{\Xi^-} \Xi^- + \kappa_{\Xi^0} \Xi^0 + \kappa_\Lambda \Lambda \ ,
\end{split}
\end{align}
where the second line defines the matrices $\kappa_B$. Let us from now on use $X\in\{L,R\}$ and, as a convenient notation, $\bar{L}=R$ and $\bar{R}=L$. Where they are not used as an index, $L$ and $R$ are meant to be elements of $\operatorname{SU}(3)_{L/R}$. Defining $u_R = u$ and $u_L = u^\dagger$ the transformation properties of meson and baryon fields under chiral rotations read%
\begin{subequations}
\begin{align}
u_{X} &\overset{\hat{\chi}}{\longrightarrow} X u_{X} K^\dagger = K u_{X} \bar{X}^\dagger \ , \\
B &\overset{\hat{\chi}}{\longrightarrow} K B K^\dagger \ ,
\end{align}
\end{subequations}
with the so-called compensator field $K$, which is a common, nonlinear realization of chiral symmetry~\cite{Coleman:1969sm,Gasser:1987rb}. The covariant derivative acting on a baryon field is defined as%
\begin{align}
D_\mu B &= \partial_\mu B + [\Gamma_\mu,B] \ ,
\end{align}
where $\Gamma_\mu$ is called the chiral connection and is given by%
\begin{align}
\Gamma_\mu &= \frac{1}{2} \bigl( u^\dagger \partial_\mu u + u \partial_\mu u^\dagger \bigr) \ .
\end{align}
The chiral vielbein $u_\mu$ and the quark mass insertions $\chi_\pm$ are defined as%
\begin{subequations}
\begin{align}
u_\mu &= i \bigl( u^\dagger \partial_\mu u - u \partial_\mu u^\dagger \bigr) \ , \\
\chi_\pm &= u^\dagger \chi u^\dagger \pm u \chi^\dagger u \ ,
\end{align}
\end{subequations}
where $\chi=2 B_0 \mathcal{M}$ includes the quark mass matrix, and transform under chiral rotations as follows:%
\begin{subequations}
\begin{align}
u_\mu &\overset{\hat{\chi}}{\longrightarrow} K u_\mu K^\dagger \ , \\
\chi_\pm &\overset{\hat{\chi}}{\longrightarrow} K \chi_\pm K^\dagger \ .
\end{align}
\end{subequations}%
\begin{table}[t]
\centering%
\begin{tabular}{l r r r r}
\toprule
$\Gamma$ & $\eta^P_\Gamma$ & $\eta^C_\Gamma$ & $\eta^h_\Gamma$ & $\eta^5_\Gamma$ \\ \midrule
$\mathds{1}$ & $1$ & $-1$ & $1$ & $1$\\
$\gamma_5$ & $-1$ & $-1$ & $-1$ & $1$\\
$\gamma_\mu$ & $(-1)_\mu$ & $1$ & $1$ & $-1$\\
$\gamma_\mu \gamma_5$ & $-(-1)_\mu$ & $-1$ & $1$& $-1$ \\
$\sigma_{\mu\nu}$ & $(-1)_\mu (-1)_\nu$ & $1$ & $1$& $1$ \\ \bottomrule
\end{tabular}%
\caption{\label{symmetry_properties_clifford_algebra} The constants $\eta^P_\Gamma$, $\eta^C_\Gamma$, $\eta^h_\Gamma$ and $\eta^5_\Gamma$ characterizing the symmetry properties of the elements of the Clifford algebra, where $(-1)_\mu$ is $1$ for $\mu=0$ and $-1$ for $\mu=1,2,3$.}
\end{table}%
Finally we define for the elements of the Clifford algebra in a unitary representation%
\begin{subequations}
\begin{align}
\Gamma &= \eta^P_\Gamma \gamma_0 \Gamma \gamma_0 \ , \\
\Gamma^T &= \eta^C_\Gamma C \Gamma C \ , \\
\Gamma^\dagger &= \eta^h_\Gamma \gamma_0 \Gamma \gamma_0 \ , \\
\Gamma &= \eta^5_\Gamma \gamma_5 \Gamma \gamma_5 \ ,
\end{align}
\end{subequations}
where $C=i\gamma^2 \gamma^0$ is the charge conjugation matrix. The different $\eta$'s are collected in table~\ref{symmetry_properties_clifford_algebra}.%
\section{Operator construction} \label{sect_operator_construction}
In this section we will construct the light-cone ($n$ is a lightlike four-vector) three-quark operator%
\begin{align}
q^a_\alpha (a_1 n) q^b_\beta (a_2 n) q^c_\gamma (a_3 n) \ ,
\end{align}
in terms of baryon octet and meson octet fields. The antisymmetrization in color indices (which makes the operator a color singlet) and the Wilson lines connecting the quark fields (providing gauge invariance) are not written out explicitly. $a$, $b$, $c$ are flavor and $\alpha$, $\beta$, $\gamma$ Dirac indices. Note that there are many possible parametrizations owing to the freedom of choice one has by neglecting higher order effects. The task is therefore not only to find a parametrization, but to find one that is most convenient for the loop calculation to be performed and can be easily matched to the standard decomposition given in ref.~\cite{Braun:2000kw}. For the parametrization of the nonlocal operator one needs functions, where the moments of the functions play the role of low energy constants (LECs). For the parametrization presented below these functions can be easily matched to standard distribution amplitudes.%
\subsection{Symmetry properties}\label{sect_sym_3qopp}
To perform the construction of an operator within the effective theory we have to know its symmetry properties. To make use of chiral symmetry it is convenient to split the quark fields in left- and right-handed parts%
\begin{align} \label{3q_op_decompRL}
\begin{split}
q^a_\alpha (a_1 n) q^b_\beta (a_2 n) q^c_\gamma (a_3 n) &=\mathcal O^{abc}_{RR,\alpha\beta\gamma}(a_1,a_2,a_3) + \mathcal O^{abc}_{LL,\alpha\beta\gamma}(a_1,a_2,a_3) \\
&\quad+ \mathcal O^{abc}_{RL,\alpha\beta\gamma}(a_1,a_2,a_3) + \mathcal O^{abc}_{LR,\alpha\beta\gamma}(a_1,a_2,a_3) \\
&\quad+ \mathcal O^{cab}_{RL,\gamma\alpha\beta}(a_3,a_1,a_2) + \mathcal O^{cab}_{LR,\gamma\alpha\beta}(a_3,a_1,a_2) \\
&\quad+ \mathcal O^{bca}_{RL,\beta\gamma\alpha}(a_2,a_3,a_1) + \mathcal O^{bca}_{LR,\beta\gamma\alpha}(a_2,a_3,a_1) \ ,
\end{split}
\end{align}
where the operators $\mathcal O_{XY}$ for $X$, $Y$ $\in \{L,R\}$ are given by%
\begin{align} \label{3qOp_definition_chiral}
\mathcal O^{abc}_{XY,\alpha\beta\gamma}(a_1,a_2,a_3) = q^a_{X,\alpha}(a_1 n) q^b_{X,\beta}(a_2 n) q^c_{Y,\gamma}(a_3 n) \ ,
\end{align}
where the left-/right-handed quark fields are defined as $q_{L/R}=\gamma_{L/R} \, q$ with the projection matrices $\gamma_{L/R}=(1\mp \gamma_5)/2$. These operators can be characterized by their transformation properties under parity transformation ($\hat p$), charge ($\hat c$) and hermitian ($\dagger$) conjugation and chiral rotations ($\hat \chi$):%
\begin{subequations}
\begin{align}
\mathcal O^{abc}_{XY,\alpha\beta\gamma}(a_1,a_2,a_3) &\overset{\hat p}{\longrightarrow} (\gamma_0)_{\alpha \alpha^\prime} (\gamma_0)_{\beta \beta^\prime} (\gamma_0)_{\gamma \gamma^\prime} \mathcal O^{abc}_{\bar X \bar Y,\alpha^\prime \beta^\prime \gamma^\prime}(a_1,a_2,a_3) \label{sym_P} \ , \\
\mathcal O^{abc}_{XY,\alpha\beta\gamma}(a_1,a_2,a_3) &\overset{\mathclap{\hat c \ \dagger \ \hat p}}{\longrightarrow} -C_{\alpha \alpha^\prime} C_{\beta \beta^\prime} C_{\gamma \gamma^\prime} \mathcal O^{abc}_{XY,\alpha^\prime \beta^\prime \gamma^\prime}(a_1,a_2,a_3) \ , \label{sym_hPC} \\
\mathcal O^{abc}_{XY,\alpha\beta\gamma}(a_1,a_2,a_3) &\overset{\hat \chi}{\longrightarrow} X_{a a^\prime} X_{b b^\prime} Y_{c c^\prime} \mathcal O^{a^\prime b^\prime c^\prime}_{XY,\alpha\beta\gamma}(a_1,a_2,a_3) \ ,
\end{align}
\end{subequations}
where in eq.~\eqref{sym_hPC} charge conjugation is performed first. Additionally we know that each operator transforms under a translation in $n$-direction as%
\begin{align}
\begin{split}
\mathcal O^{abc}_{XY,\alpha\beta\gamma}(a_1+\delta a,a_2+\delta a,a_3+\delta a) &= \exp \left\{i \, \delta a \, n \cdot \hat P \right\}\mathcal O^{abc}_{XY,\alpha\beta\gamma}(a_1,a_2,a_3)\exp \left\{-i \, \delta a \, n \cdot \hat P \right\} \ ,
\end{split}
\end{align}
where $\hat P$ is the momentum operator which acts as a generator of translations. Another symmetry of the three-quark operators defined in eq.~\eqref{3qOp_definition_chiral} is the invariance under the exchange of the quark in the first and the second position or even an invariance under exchange of all three quarks in case of the operators containing right-handed or left-handed fields exclusively. On top of this the operator is invariant if one simultaneously rescales $a_i \longrightarrow \lambda a_i$ and $n_\mu \longrightarrow n_\mu/\lambda$, which we will call scaling property.%
\subsection{Low energy operators}
Using the previously defined fields $u_R$ and $u_L$ we can write down the operators, which contribute to baryon-to-vacuum matrix elements of three-quark currents at leading one-loop level and have correct transformation properties under chiral rotations in the following compact form:%
\begin{align}\label{eff_opp}
\begin{split}
\mathcal O^{abc}_{XY,\alpha\beta\gamma}(a_1,a_2,a_3) &= \int [dx] \sum \limits_{i,j} \sum \limits_{k=1}^{k_j} \mathcal{F}^{i,j,k}_{XY}(x_1,x_2,x_3) \, \Gamma^{i,XXY}_{\alpha \beta \gamma \delta} \, B^{j,k,XXY}_{\delta,abc}(z) \ ,
\end{split}
\end{align}
where the correct transformation behaviour under translations in $n$-direction is ensured by $z_\mu= n_\mu \sum x_i a_i$ and the constraint that $x_1+x_2+x_3=1$ in%
\begin{align}
\int [dx]&= \int dx_1 dx_2 dx_3 \, \delta \bigl(1-\sum x_i\bigr) \ ,
\end{align}
where the integrations run from $0$ to $1$. The $\mathcal{F}$'s are functions of $x_1$, $x_2$, $x_3$ only and $k_j$ is given in table~\ref{symmetry_properties_B}. The $\Gamma$'s are defined as%
\begin{align}
\Gamma^{i,XYZ}_{\alpha \beta \gamma \delta} &= (\gamma_X \Gamma^i_A \gamma_Y C)_{\alpha \beta} (\gamma_Z \Gamma^i_B (i \slashed{\partial})^{d^m_i})_{\gamma \delta} (n \cdot \partial)^{d^n_i} \ , \label{definition_Gamma_structure}
\end{align}
\begin{table}[t]
\centering%
\begin{tabular}{c c c c c c c}
\toprule
$i$ & $\qquad\mathllap{\Gamma^i_A } \otimes \mathrlap{ \Gamma^i_B}\qquad $ & $\eta^h_{\Gamma,i}=\eta^C_{\Gamma,i}$ & $\eta^{C}_{\Gamma^A,i}$ & $\eta^{5}_{\Gamma^B,i}$ & $d^m_i$ & $d^n_i$ \\ \midrule
$1$ & $ \mathllap{i \mathds{1} } \otimes \mathrlap{ \slashed{n}} $ & $-1$ & $-1$ &$-1$ & $2$ & $-1$\\
$2$ & $ \mathllap{\mathds{1} } \otimes \mathrlap{ \mathds{1}} $ & $\phantom{+}1$ & $-1$ &$\phantom{+}1$ & $1$ & $\phantom{-}0$ \\
$3$ & $ \mathllap{\sigma^{\partial n} } \otimes \mathrlap{ \slashed{n}} $ & $\phantom{+}1$ & $\phantom{+}1$ &$-1$& $2$ & $-2$ \\
$4$ & $ \mathllap{\sigma^{\mu n} } \otimes \mathrlap{ \gamma_\mu} $ & $\phantom{+}1$ & $\phantom{+}1$ &$-1$& $2$ & $-1$\\
$5$ & $ \mathllap{\sigma^{\mu \nu} } \otimes \mathrlap{ \sigma_{\mu \nu}} $ & $\phantom{+}1$ & $\phantom{+}1$ &$\phantom{+}1$& $1$& $\phantom{-}0$ \\
$6$ & $ \mathllap{i \sigma^{\partial n} } \otimes \mathrlap{ \mathds{1}} $ & $-1$ & $\phantom{+}1$ &$\phantom{+}1$& $1$ & $-1$ \\
$7$ & $ \mathllap{\sigma^{\mu \partial} } \otimes \mathrlap{ \sigma_{\mu n}} $ & $\phantom{+}1$ & $\phantom{+}1$ &$\phantom{+}1$& $1$ & $-1$ \\
$8$ & $ \mathllap{\sigma^{\mu \partial} } \otimes \mathrlap{ \gamma_\mu} $ & $\phantom{+}1$ & $\phantom{+}1$ &$-1$& $0$ & $\phantom{-}0$ \\
$9$ & $ \mathllap{\sigma^{\mu n} } \otimes \mathrlap{ \sigma_{\mu n}} $ & $\phantom{+}1$ & $\phantom{+}1$ &$\phantom{+}1$& $3$ & $-2$ \\ \bottomrule
\end{tabular}
\caption{\label{symmetry_properties_Gamma} List of $\Gamma^i_A \otimes \Gamma^i_B $. $\eta^P_{\Gamma,i}=1$ by choice (see comment in the text). We have multiplied structures $1$ and $6$ with a factor of $i$ such that $\eta^h_{\Gamma,i}=\eta^C_{\Gamma,i}$ for all structures and, thus, $\eta^{hPC}_{\Gamma,i}=1$. In cases where four-vectors are used in the place of Lorentz indices the notation means that the corresponding Lorentz index is contracted with the index of the vector; e.g.\ $\sigma^{\partial n}=\sigma^{\mu\nu} \partial_\mu n_\nu$.}
\end{table}%
where $\Gamma^i_A$, $\Gamma^i_B$, $d^m_i$ and $d^n_i$ can be taken from table~\ref{symmetry_properties_Gamma}. The occurring derivatives act on the $B$'s. We have introduced adequate powers of $i \slashed{\partial}$ to have functions $\mathcal{F}$ of mass dimension $2$, which is compatible with the standard mass dimension of distribution amplitudes. Using $i \slashed{\partial}$ (which leads to a factor $m_B$ in the final result) instead of the baryon mass in the chiral limit $m_0$ (which would be the standard choice) has the advantage that it allows for a straightforward matching of our parametrization to the general decomposition given in ref.~\cite{Braun:2000kw} and to refs.~\mbox{\cite{Liu:2008yg,Liu:2009uc}} (see also section~\ref{sect_matching}). The power of $(n \cdot \partial)$ is chosen such that the scaling property is fulfilled. Note that in the chiral-odd sector one can actually write down more structures, which have the form $\Gamma^{i,XYX}_{\beta \gamma \alpha \delta}$ or $\Gamma^{i,YXX}_{\gamma \alpha \beta \delta}$. However, these structures are not independent. They can be rewritten in terms of $\Gamma^{i,XXY}_{ \alpha \beta \gamma \delta}$ using Fierz transformation. In order to reduce the $\Gamma$'s to the minimal set given in table~\ref{symmetry_properties_Gamma} one has to use the identity $\sigma^{\mu\nu} \gamma_5 = \frac{i}{2} \varepsilon^{\mu\nu\rho\sigma}\sigma_{\rho\sigma}$ and the fact that it is sufficient to construct structures of positive parity (see explanation below eq.~\eqref{parity_constraint}). Additionally one has to use that multiplying both structures $\Gamma^i_A$ and $\Gamma^i_B$ with a $\gamma_5$ does not lead to a new, independent structure owing to the projection with $\gamma_{L/R}$ in eq.~\eqref{definition_Gamma_structure}.\par
The $B$'s in eq.~\eqref{eff_opp} are defined as%
\begin{align}
B^{j,k,XYZ}_{\delta,abc} &= (u_X)_{a a^\prime} (u_Y)_{b b^\prime} (u_Z)_{c c^\prime} B^{j,k}_{\delta,a^\prime b^\prime c^\prime} \ ,
\end{align}%
\begin{table}[t]
\centering%
\begin{tabular}{c c c c c c}
\toprule
$j$ & $B^j_{1,\delta}$ & $B^j_2$ & $B^j_3$ & trace$^j$ & $k_j$ \\ \midrule
$1$ & $B_\delta$ & $\mathds{1}$ & $\mathds{1}$ & $1$ & $3$ \\
$2$ & $B_\delta$ & $\mathds{1}$ & $\mathds{1}$ & $\tr{\chi_+} m_0^{-2} $ & $3$ \\
$3$ & $B_\delta$ & $\tilde\chi_+ m_0^{-2} $ & $\mathds{1}$ & $1$ & $6$ \\ \bottomrule
\end{tabular}
\caption{\label{symmetry_properties_B} In this table we list only terms which contribute to the one-loop calculation of baryon-to-vacuum matrix elements of the operator. $\tilde \chi_+$ is defined as $\chi_+ -\tr{\chi_+}/3$. This is a convenient choice since this combination (in a leading one-loop calculation) vanishes along the symmetric line, where $m_u=m_d=m_s$.}
\end{table}%
where%
\begin{subequations}
\begin{align}
B^{j,1}_{\delta,a b c} &= (B^j_{1,\delta})_{a a^\prime} (B^j_2)_{b b^\prime} (B^j_3)_{c c^\prime} \varepsilon_{a^\prime b^\prime c^\prime} \times \text{trace$^j$} \ , \\
B^{j,2}_{\delta,a b c} &= (B^j_3)_{a a^\prime} (B^j_{1,\delta})_{b b^\prime} (B^j_2)_{c c^\prime} \varepsilon_{a^\prime b^\prime c^\prime} \times \text{trace$^j$} \ , \\
B^{j,3}_{\delta,a b c} &= (B^j_2)_{a a^\prime} (B^j_3)_{b b^\prime} (B^j_{1,\delta})_{c c^\prime} \varepsilon_{a^\prime b^\prime c^\prime} \times \text{trace$^j$} \ , \\
B^{j,4}_{\delta,a b c} &= (B^j_2)_{a a^\prime} (B^j_{1,\delta})_{b b^\prime} (B^j_3)_{c c^\prime} \varepsilon_{a^\prime b^\prime c^\prime} \times \text{trace$^j$} \ , \\
B^{j,5}_{\delta,a b c} &= (B^j_{1,\delta})_{a a^\prime} (B^j_3)_{b b^\prime} (B^j_2)_{c c^\prime} \varepsilon_{a^\prime b^\prime c^\prime} \times \text{trace$^j$} \ , \\
B^{j,6}_{\delta,a b c} &= (B^j_3)_{a a^\prime} (B^j_2)_{b b^\prime} (B^j_{1,\delta})_{c c^\prime} \varepsilon_{a^\prime b^\prime c^\prime} \times \text{trace$^j$} \ .
\end{align}
\end{subequations}
For cases where $B^j_2=B^j_3$ we only use $ B^{j,1}_{\delta,a b c}$, $ B^{j,2}_{\delta,a b c}$ and $ B^{j,3}_{\delta,a b c}$ and thus $k_j=3$. The different possible combinations of $B$'s can be taken from table~\ref{symmetry_properties_B}. All baryon and meson fields which are connected to each other (by a summation over a shared flavor index) have to be at the same spacetime position, owing to the fact that the compensator field $K$ is a local transformation. However, chiral symmetry actually also allows for the possibility that the trace term in $B$ is situated at a different spacetime position as the rest of the operator. We consider this possibility in appendix~\ref{app_no_cov_der} and show that such a parametrization only differs in higher order terms. Note that no structures of the form $[B_\delta,\tilde\chi_+]$, $\{B_\delta,\tilde\chi_+\}$, or $\tr{B_\delta \tilde\chi_+}$ occur in table~\ref{symmetry_properties_B}, since they can be reexpressed in terms of the third structure, which means that we have only one second order structure ($j=3$) that is responsible for $\operatorname{SU}(3)_f$ breaking. Also the operators which describe the behaviour along the $\operatorname{SU}(3)_f$ symmetric line ($j=1,2$) are not linearly independent, but the situation is more complicated in this case: since operators of the same class (i.e.\ same $j$ but different $k$) are related to each other (see eq.~\eqref{lin_dep_j12}) one has to take care that the symmetry properties of the operator under quark exchange are respected. Therefore, we postpone this discussion to section~\ref{sect_elim}. \par
There are no covariant derivatives acting on the baryon field within the $B$'s. In appendix~\ref{app_no_cov_der} we show that they can always be traded for derivatives acting on the whole structure plus higher order contributions, which can be neglected. This fact will turn out to be very convenient for calculating loop contributions, since the derivatives acting on the complete structure do not lead to additional loop momenta in the integrals. \par
The effective operator given in eq.~\eqref{eff_opp} already transforms correctly under chiral rotations and translations along the light-cone vector $n$. It also fulfills the scaling property. The remaining symmetry properties given in section~\ref{sect_sym_3qopp} will now be implemented by constraining the functions $\mathcal{F}$. We consider%
\begin{subequations}
\begin{align}
B^{j,k,XYZ}_{\delta,abc} &\overset{\hat p}{\longrightarrow} (\gamma_0)_{\delta \delta^\prime} B^{j,k,\bar{X}\bar{Y}\bar{Z}}_{\delta^\prime,abc} \ , \label{P_B} \\
B^{j,k,XYZ}_{\delta,abc} &\overset{\mathclap{\hat c \ \dagger \ \hat p}}{\longrightarrow} - C_{\delta \delta^\prime} B^{j,k,XYZ}_{\delta^\prime,abc} \ , \label{hPC_B}
\end{align}
\end{subequations}
and%
\begin{subequations}
\begin{align}
\Gamma^{i,XYZ}_{\alpha \beta \gamma \delta} &\overset{\hat p}{\longrightarrow} -\eta^P_{\Gamma,i} \ (\gamma_0)_{\alpha \alpha^\prime} (\gamma_0)_{\beta \beta^\prime} (\gamma_0)_{\gamma \gamma^\prime} \ \Gamma^{i,\bar{X}\bar{Y}\bar{Z}}_{\alpha^\prime \beta^\prime \gamma^\prime \delta^\prime}\ (\gamma_0)_{\delta^\prime \delta} \ , \label{P_Gamma} \\
\begin{split}
\Gamma^{i,XYZ}_{\alpha \beta \gamma \delta} &\overset{\mathclap{\hat c \ \dagger \ \hat p}}{\longrightarrow} - \eta^{hPC}_{\Gamma,i} \ C_{\alpha \alpha^\prime} C_{\beta \beta^\prime} C_{\gamma \gamma^\prime} \ \Gamma^{i,XYZ}_{\alpha^\prime \beta^\prime \gamma^\prime \delta^\prime} \ C_{\delta^\prime \delta} \ . \label{hPC_Gamma}
\end{split}
\end{align}
\end{subequations}
Eqs.~\eqref{hPC_B} and \eqref{hPC_Gamma} yield (together with eqs.~\eqref{sym_hPC} and \eqref{eff_opp} and since $\eta^{hPC}_{\Gamma}=1$)%
\begin{align} \label{phase_1}
\bigl(\mathcal{F}^{i,j,k}_{XY}\bigr)^* &= \mathcal{F}^{i,j,k}_{XY} \ ,
\end{align}
which would mean that the $\mathcal{F}$'s are real-valued. However this argument relies on the assumption that one gets no additional phases from charge conjugation of quarks and baryons, which is not necessarily true. If we allow for such additional phases the above equation has to be generalized to%
\begin{align} \label{phase_2}
\bigl(\mathcal{F}^{i,j,k}_{XY} e^{i \theta}\bigr)^* &= \mathcal{F}^{i,j,k}_{XY} e^{i \theta} \ ,
\end{align}
where we have an additional overall phase which is equal for all distribution amplitudes. However, this additional phase is unphysical and can be dropped. Eqs.~\eqref{P_B} and \eqref{P_Gamma} yield (together with eqs.~\eqref{sym_P} and \eqref{eff_opp} and since $\eta^P_{\Gamma}=1$)%
\begin{align} \label{parity_constraint}
{\mathcal{F}^{i,j,k}_{\bar X \bar Y}} &= -\mathcal{F}^{i,j,k}_{XY} \ .
\end{align}
Therefore we only have to differentiate between chiral-even $\mathcal{F}^{i,j,k}_{RR}=-\mathcal{F}^{i,j,k}_{LL}\equiv \mathcal{F}^{i,j,k}_{\text{even}}$ and chiral-odd $\mathcal{F}^{i,j,k}_{LR}=-\mathcal{F}^{i,j,k}_{RL}\equiv \mathcal{F}^{i,j,k}_{\text{odd}}$. Notice that we have chosen to only construct structures $\Gamma_A \otimes \Gamma_B$ which have positive parity. The negative parity structures, which one can obtain by multiplying all $\Gamma_B$ with a $\gamma_5$, would lead to the same operators since eq.~\eqref{parity_constraint} then would yield an extra minus sign.%
\subsection{Symmetry under exchange of quark fields}
In this section we use the symmetry of the original three-quark operators under exchange of quark fields with the same handedness to reduce the number of amplitudes. Using the constraint that the operators have to be equal under exchange of the first and the second quark yields%
\begin{subequations} \label{symmetry_constraints_12}
\begin{align}
&\text{\underline{$j=1,2$:}} & \mathcal{F}^{i,j,1}_{XY}{(x_1,x_2,x_3)} &= -\eta^C_{\Gamma^A,i} \mathcal{F}^{i,j,2}_{XY} {(x_2,x_1,x_3)} \ , \label{symmetry_constraints_12_kj3_1}\\
& & \mathcal{F}^{i,j,3}_{XY}{(x_1,x_2,x_3)} &= -\eta^C_{\Gamma^A,i} \mathcal{F}^{i,j,3}_{XY} {(x_2,x_1,x_3)} \ , \label{symmetry_constraints_12_kj3_2} \\[0.5cm]
&\text{\underline{$j=3$:}} & \mathcal{F}^{i,3,1}_{XY}{(x_1,x_2,x_3)} &= -\eta^C_{\Gamma^A,i} \mathcal{F}^{i,3,4}_{XY} {(x_2,x_1,x_3)} \ , \\
& & \mathcal{F}^{i,3,2}_{XY}{(x_1,x_2,x_3)} &= -\eta^C_{\Gamma^A,i} \mathcal{F}^{i,3,5}_{XY} {(x_2,x_1,x_3)} \ , \\
& & \mathcal{F}^{i,3,3}_{XY}{(x_1,x_2,x_3)} &= -\eta^C_{\Gamma^A,i} \mathcal{F}^{i,3,6}_{XY} {(x_2,x_1,x_3)} \ .
\end{align}
\end{subequations}
In the chiral-odd sector one now uses these relations to eliminate $\mathcal{F}^{i,j,2}_{XY}$ (if $j=1,2$) and $\mathcal{F}^{i,3,4/5/6}_{XY}$. Additionally we can use that%
\begin{align} \label{chiral_proj_cons_XXY}
(\gamma_Y \Gamma_A \gamma_X C)_{\gamma \beta} (\gamma_X \Gamma_B)_{\alpha \delta} &= 0 \ , \quad \text{if } X\neq Y \text{ and } \Gamma_A \in \{ \mathds{1}, \gamma_5, \sigma_{\mu\nu} \} \ .
\end{align}
Using Fierz transformation this leads to%
\begin{subequations}
\begin{align}
\Gamma^{3,XXY}_{\alpha\beta\gamma\delta} &= \Gamma^{4,XXY}_{\alpha\beta\gamma\delta} + \frac{1}{2} \Gamma^{9,XXY}_{\alpha\beta\gamma\delta} \ , \\
\Gamma^{5,XXY}_{\alpha\beta\gamma\delta} &= 0 \ , \\
\Gamma^{6,XXY}_{\alpha\beta\gamma\delta} &= \Gamma^{4,XXY}_{\alpha\beta\gamma\delta} - \Gamma^{7,XXY}_{\alpha\beta\gamma\delta} \ ,
\end{align}
\end{subequations}
if $X\neq Y$. Therefore we have the freedom to choose%
\begin{align}
\mathcal{F}^{3,j,k}_{\text{odd}}{(x_1,x_2,x_3)}&=
\mathcal{F}^{5,j,k}_{\text{odd}}{(x_1,x_2,x_3)}=
\mathcal{F}^{6,j,k}_{\text{odd}}{(x_1,x_2,x_3)}= 0 \ .
\end{align}
In the chiral-even sector the projection with $\gamma_{L/R}$ leads to similar constraints. The counterpart of eq.~\eqref{chiral_proj_cons_XXY} reads%
\begin{align}
(\gamma_X \Gamma_A \gamma_X C)_{\gamma \beta} (\gamma_X \Gamma_B)_{\alpha \delta} &= 0 \ , \quad \text{if } \Gamma_A \in \{ \gamma_\mu, \gamma_\mu \gamma_5 \} \ .
\end{align}
With a Fierz transformation one obtains%
\begin{subequations}
\begin{align}
\Gamma^{7,XXX}_{\alpha\beta\gamma\delta} &= - \Gamma^{4,XXX}_{\alpha\beta\gamma\delta} + \frac{1}{2} \Gamma^{5,XXX}_{\alpha\beta\gamma\delta} + \Gamma^{6,XXX}_{\alpha\beta\gamma\delta} \ , \\
\Gamma^{8,XXX}_{\alpha\beta\gamma\delta} &= -\frac{1}{4}\Gamma^{5,XXX}_{\alpha\beta\gamma\delta} \ , \\
\Gamma^{9,XXX}_{\alpha\beta\gamma\delta} &= 0 \ .
\end{align}
\end{subequations}
Therefore, we can choose%
\begin{align}
\mathcal{F}^{7,j,k}_{\text{even}}{(x_1,x_2,x_3)}&=
\mathcal{F}^{8,j,k}_{\text{even}}{(x_1,x_2,x_3)}=
\mathcal{F}^{9,j,k}_{\text{even}}{(x_1,x_2,x_3)}= 0 \ .
\end{align}
The operators containing left-/right-handed quarks exclusively also have to be invariant under an exchange of the first and the last quark. Performing a Fierz transformation and using the identities given above we find%
\begin{align}
\Gamma^{i,XXX}_{\gamma \beta \alpha \delta} & = \sum_{i^\prime=1}^6 \Gamma^{i^\prime,XXX}_{\alpha \beta \gamma \delta} c^{i^\prime i} \ .
\end{align}
The matrix $c$ is given by%
\begin{align}
c&=\left(
\begin{array}{*{6}Kc}
\frac{1}{2} & 0 & -\frac{1}{2} & -\frac{3}{2} & 0 & -\frac{1}{2} \\
0 & \frac{1}{2} & 0 & 0 & 6 & -\frac{1}{2} \\
0 & 0 & 1 & 0 & 0 & 0 \\
-\frac{1}{2} & 0 & -\frac{1}{2} & -\frac{1}{2} & 0 & -\frac{1}{2} \\
0 & \frac{1}{8} & 0 & 0 & -\frac{1}{2} & \frac{1}{8} \\
0 & 0 & 0 & 0 & 0 & 1
\end{array}
\right) .
\end{align}
By the use of this relation the symmetry property of the operator under exchange of the first and the last quark translates to the following constraints on the amplitudes:%
\begin{subequations} \label{symmetry_constraints_13}
\begin{align}
&\text{\underline{$j=1,2$:}} & \mathcal{F}^{i,j,1}_{XX}{(x_1,x_2,x_3)} &= -\sum_{i^\prime=1}^6 c^{i i^\prime}\mathcal{F}^{i^\prime,j,3}_{XX} {(x_3,x_2,x_1)} \ , \\
& & \mathcal{F}^{i,j,2}_{XX}{(x_1,x_2,x_3)} &= -\sum_{i^\prime=1}^6 c^{i i^\prime}\mathcal{F}^{i^\prime,j,2}_{XX} {(x_3,x_2,x_1)} \ , \\
& & \mathcal{F}^{i,j,3}_{XX}{(x_1,x_2,x_3)} &= -\sum_{i^\prime=1}^6 c^{i i^\prime}\mathcal{F}^{i^\prime,j,1}_{XX} {(x_3,x_2,x_1)} \ ,\\[0.5cm]
&\text{\underline{$j=3$:}} & \mathcal{F}^{i,3,1}_{XX}{(x_1,x_2,x_3)} &= -\sum_{i^\prime=1}^6 c^{i i^\prime}\mathcal{F}^{i^\prime,3,6}_{XX} {(x_3,x_2,x_1)} \ , \\
& & \mathcal{F}^{i,3,2}_{XX}{(x_1,x_2,x_3)} &= -\sum_{i^\prime=1}^6 c^{i i^\prime}\mathcal{F}^{i^\prime,3,4}_{XX} {(x_3,x_2,x_1)} \ , \\
& & \mathcal{F}^{i,3,3}_{XX}{(x_1,x_2,x_3)} &= -\sum_{i^\prime=1}^6 c^{i i^\prime}\mathcal{F}^{i^\prime,3,5}_{XX} {(x_3,x_2,x_1)} \ , \\
& & \mathcal{F}^{i,3,4}_{XX}{(x_1,x_2,x_3)} &= -\sum_{i^\prime=1}^6 c^{i i^\prime}\mathcal{F}^{i^\prime,3,2}_{XX} {(x_3,x_2,x_1)} \ , \\
& & \mathcal{F}^{i,3,5}_{XX}{(x_1,x_2,x_3)} &= -\sum_{i^\prime=1}^6 c^{i i^\prime}\mathcal{F}^{i^\prime,3,3}_{XX} {(x_3,x_2,x_1)} \ , \\
& & \mathcal{F}^{i,3,6}_{XX}{(x_1,x_2,x_3)} &= -\sum_{i^\prime=1}^6 c^{i i^\prime}\mathcal{F}^{i^\prime,3,1}_{XX} {(x_3,x_2,x_1)} \ .
\end{align}
\end{subequations}
Using these equations one finds for the operator with $j = 3$ that one can eliminate all amplitudes apart from $\mathcal{F}^{i,3,1}_{XX}$, by using the following relations recursively:%
\begin{subequations}
\begin{align}
\mathcal{F}^{i,3,2}_{XX}{(x_1,x_2,x_3)} &= \eta^C_{\Gamma^A,i} \sum_{i^\prime=1}^6 c^{i i^\prime} \mathcal{F}^{i^\prime,3,3}_{XX} {(x_3,x_1,x_2)} \ , \\
\mathcal{F}^{i,3,3}_{XX}{(x_1,x_2,x_3)} &= \eta^C_{\Gamma^A,i} \sum_{i^\prime=1}^6 c^{i i^\prime} \mathcal{F}^{i^\prime,3,1}_{XX} {(x_3,x_1,x_2)} \ , \\
\mathcal{F}^{i,3,4}_{XX}{(x_1,x_2,x_3)} &= -\eta^C_{\Gamma^A,i} \mathcal{F}^{i,3,1}_{XX} {(x_2,x_1,x_3)} \ , \\
\mathcal{F}^{i,3,5}_{XX}{(x_1,x_2,x_3)} &= -\eta^C_{\Gamma^A,i} \mathcal{F}^{i,3,2}_{XX} {(x_2,x_1,x_3)} \ , \\
\mathcal{F}^{i,3,6}_{XX}{(x_1,x_2,x_3)} &= -\eta^C_{\Gamma^A,i} \mathcal{F}^{i,3,3}_{XX} {(x_2,x_1,x_3)} \ ,
\end{align}
\end{subequations}
For the operators with $j =1,2$ we can eliminate%
\begin{subequations}
\begin{align}
\mathcal{F}^{i,j,2}_{XX}{(x_1,x_2,x_3)}&= - \eta^C_{\Gamma^A,i} \mathcal{F}^{i,j,1}_{XX}{(x_2,x_1,x_3)} \ , \\
\mathcal{F}^{i,j,3}_{XX}{(x_1,x_2,x_3)}&= - \sum_{i^\prime=1}^6 c^{i i^\prime} \mathcal{F}^{i^\prime,j,1}_{XX}{(x_3,x_2,x_1)} \ ,
\end{align}
\end{subequations}
and additionally%
\begin{subequations}
\begin{align}
\begin{split}
\mathcal{F}^{1,j,1}_{XX}{(x_1,x_2,x_3)}&= \mathcal{F}^{3,j,1}_{XX}{(x_1,x_2,x_3)} + \mathcal{F}^{4,j,1}_{XX}{(x_1,x_2,x_3)} - 2 \mathcal{F}^{4,j,1}_{XX}{(x_1,x_3,x_2)} \\&\quad + \mathcal{F}^{6,j,1}_{XX}{(x_1,x_2,x_3)} \ ,
\end{split} \\
\mathcal{F}^{2,j,1}_{XX}{(x_1,x_2,x_3)}&= -4 \mathcal{F}^{5,j,1}_{XX}{(x_1,x_2,x_3)} + 8 \mathcal{F}^{5,j,1}_{XX}{(x_1,x_3,x_2)} + \mathcal{F}^{6,j,1}_{XX}{(x_1,x_2,x_3)} \ , \\
\mathcal{F}^{3,j,1}_{XX}{(x_1,x_2,x_3)}&=-\mathcal{F}^{3,j,1}_{XX}{(x_1,x_3,x_2)} \ , \\
\mathcal{F}^{6,j,1}_{XX}{(x_1,x_2,x_3)}&=-\mathcal{F}^{6,j,1}_{XX}{(x_1,x_3,x_2)} \ .
\end{align}
\end{subequations}
From the fact that the local operator at the origin, where $a_1=a_2=a_3=0$, is independent of the light-cone vector $n$ one can deduce constraints for the zeroth moments of the distribution amplitudes%
\begin{align} \label{zeroth_moments}
\int [dx] \mathcal{F}^{i,j,k}_{XY}{(x_1,x_2,x_3)} &= 0 \ , \quad \text{for } i=1,3,4,6,7,9 \ .
\end{align}
\subsection{Elimination of linearly dependent structures} \label{sect_elim}
To avoid overparametrization we will now annihilate linearly dependent structures of those given in table~\ref{symmetry_properties_B}. Considering all possible three-quark operators and all baryons from the octet, one finds (for $j=1,2$) that only two out of the three structures $B^{j,1}_{\delta,a b c}$, $B^{j,2}_{\delta,a b c}$ and $B^{j,3}_{\delta,a b c}$ are linearly independent, since one has%
\begin{align} \label{lin_dep_j12}
0&= B^{j,1}_{\delta,a b c} + B^{j,2}_{\delta,a b c} + B^{j,3}_{\delta,a b c} \ .
\end{align}
In the chiral-odd sector we can use this relation to replace $B^{j,3}_{\delta,a b c}=- B^{j,1}_{\delta,a b c} - B^{j,2}_{\delta,a b c} $, which is equivalent to the replacement%
\begin{subequations}
\begin{align}
\mathcal{F}^{i,j,1}_{\text{odd}}{(x_1,x_2,x_3)} &\longrightarrow \tilde{\mathcal{F}}^{i,j,1}_{\text{odd}}{(x_1,x_2,x_3)} \equiv \mathcal{F}^{i,j,1}_{\text{odd}}{(x_1,x_2,x_3)} - \mathcal{F}^{i,j,3}_{\text{odd}}{(x_1,x_2,x_3)} \ , \\
\mathcal{F}^{i,j,2}_{\text{odd}}{(x_1,x_2,x_3)} &\longrightarrow \tilde{\mathcal{F}}^{i,j,2}_{\text{odd}}{(x_1,x_2,x_3)} \equiv \mathcal{F}^{i,j,2}_{\text{odd}}{(x_1,x_2,x_3)} - \mathcal{F}^{i,j,3}_{\text{odd}}{(x_1,x_2,x_3)} \ , \\
\mathcal{F}^{i,j,3}_{\text{odd}}{(x_1,x_2,x_3)} &\longrightarrow \tilde{\mathcal{F}}^{i,j,3}_{\text{odd}}{(x_1,x_2,x_3)} \equiv 0 \ .
\end{align}
\end{subequations}
Using eqs.~\eqref{symmetry_constraints_12_kj3_1} and~\eqref{symmetry_constraints_12_kj3_2} one finds that the new functions have the same symmetry properties as the old ones. Therefore we can choose%
\begin{align}
\mathcal{F}^{i,j,3}_{\text{odd}}{(x_1,x_2,x_3)}&=0 \ , \quad j=1,2 \ ,
\end{align}
in accordance with symmetry properties and without loss of generality. In the chiral-even sector the situation is different since the amplitudes are already constrained by the symmetry under exchange of the first and the third quark. An elimination of one structure in favor of the two others would therefore not lead to a simplification. Instead one just obtains a reparametrization of the problem for which it would be hard to implement the symmetry properties under exchange of the first and the last quark.%
\section{Calculation at leading one-loop order} \label{sect_calculation}
In this section we describe the leading one-loop calculation. In section~\ref{sect_matching} we explain how we have matched to the standard DAs defined in ref.~\cite{Braun:2000kw}.
\subsection{Meson masses and the \texorpdfstring{$Z$}{Z}-factor}
We work in the limit of exact isospin symmetry, where $m_u=m_d \equiv m_l$. Using the standard leading order meson Lagrangian (see e.g.~\cite{Gasser:1984gg,Scherer:2002tk}) one finds for the meson masses the standard Gell-Mann-Oakes-Renner relations%
\begin{subequations}
\begin{align}
m_\pi^2 &= 2 B_0 m_l = m^2_{i=1,2,3} = 2 B_0 (\bar m_q - \delta m_l) \ , \\
m_K^2 &= B_0 (m_l+m_s) = m^2_{i=4,\dots,7} = B_0 (2 \bar m_q + \delta m_l) \ , \\
m_\eta^2 &= \frac{B_0}{3} (2 m_l+ 4 m_s) = m^2_{i=8} = 2 B_0 ( \bar m_q + \delta m_l) \ ,
\end{align}
\end{subequations}
where%
\begin{subequations}
\begin{align}
\bar m_q & = \frac{1}{3} (2 m_l + m_s) \ , \\
\delta m_l & = \bar m_q - m_l \ ,
\end{align}
\end{subequations}
and $B_0$ is the LEC proportional to the quark condensate in the chiral limit. As additional ingredient we need the first order meson-baryon Lagrangian, which we take from ref.~\cite{Bruns:2011sh} (this version differs from refs.~\cite{Krause:1990xc,Scherer:2002tk} by a minus sign in the terms containing $D$ and $F$ in order to be consistent with the standard sign convention $g_A \approx D + F > 0$):%
\begin{align}
\mathscr{L}_{MB}^{(1)} &= \tr{\bar B \gamma_\mu i D^\mu B} - m_0 \tr{\bar B B} + \frac{D}{2} \tr{\bar B \gamma_\mu \gamma_5 \{u^\mu,B\}} + \frac{F}{2} \tr{\bar B \gamma_\mu \gamma_5 [u^\mu,B]} \ .
\end{align}
For our calculation we need the baryon-meson-baryon vertex for an incoming baryon $B$, an outgoing baryon $B^\prime$ and an incoming meson (the $k$-th one in the Cartesian basis) with momentum $q$, which is given by%
\begin{align} \label{BMB_Vertex}
\frac{-1}{2 F_0} \slashed{q} \gamma_5 \tr{\kappa_{B^\prime}^T (D \{\lambda^k,\kappa_B\}+F [\lambda^k,\kappa_B])} \ .
\end{align}
The self-energy to third chiral order is given by the sum of the irreducible diagrams shown in figure~\ref{FD_selfenergy} (where external legs are to be amputated) multiplied with an $i$. %
\begin{figure}[t]
\centering%
\subfigure[\label{fdself:subfiga}]{
\includegraphics[scale=0.55]{pics/FDSelfenergya.eps}
}
\subfigure[\label{fdself:subfigb}]{
\includegraphics[scale=0.55]{pics/FDSelfenergyb.eps}
}
\caption[]{\label{FD_selfenergy} Feynman diagrams needed for the calculation of the self-energy.}
\end{figure}%
The contribution of diagram~\subref{fdself:subfigb} (in figure~\ref{FD_selfenergy}), which is relevant for the calculation of the $Z$-factor is given by%
\begin{align}
i \times \text{\subref{fdself:subfigb}} &= 3 g_{B,\pi} f(m_\pi,m_0,\slashed{p}) + 4 g_{B,K} f(m_K,m_0,\slashed{p}) + g_{B,\eta} f(m_\eta,m_0,\slashed{p}) \ ,
\end{align}
where%
\begin{align}
f(m,m_0,\slashed{p}) &= \frac{-1}{4 F_0^2} \bigl( (p^2-m_0^2) \slashed{p} I^{(1)}_{11}{(m,m_0,\slashed{p})} + (\slashed{p}+m_0)\bigl( I_{01}{(m_0,\slashed{p})} - m^2 I_{11}{(m,m_0,\slashed{p})} \bigr) \bigr) \ .
\end{align}
The loop functions $I_{kl}$ and $I_{kl}^{(1)}$ are defined as in ref.~\cite{Wein:2011ix} and the coefficients are given by%
\begin{align}
g_{N,\pi} &= (D+F)^2 \ , & g_{N,K} &= \frac{5}{6} D^2 - D F + \frac{3}{2} F^2 \ , & g_{N,\eta} &= \frac{1}{3} (D-3F)^2 \ , \nonumber \\
g_{\Sigma,\pi} &= \frac{4}{9}(D^2 + 6 F^2) \ , & g_{\Sigma,K} &= D^2 + F^2 \ , & g_{\Sigma,\eta} &= \frac{4}{3} D^2 \ , \nonumber \\
g_{\Xi,\pi} &= (D-F)^2 \ , & g_{\Xi,K} &= \frac{5}{6} D^2 + D F + \frac{3}{2} F^2 \ , & g_{\Xi,\eta} &= \frac{1}{3} (D+3F)^2 \ , \nonumber \\
g_{\Lambda,\pi} &= \frac{4}{3} D^2 \ , & g_{\Lambda,K} &= \frac{1}{3} (D^2 + 9 F^2) \ , & g_{\Lambda,\eta} &= \frac{4}{3} D^2 \ . \label{def_g_cefficients}
\end{align}
These constants fulfill the constraints that the sums%
\begin{subequations}
\begin{align}
3 g_{B,\pi} + 4 g_{B,K} + g_{B,\eta} &= \frac{4}{3} (5 D^2+9 F^2) \ , \\
2 g_{N,M} + 3 g_{\Sigma,M} + 2 g_{\Xi,M} + g_{\Lambda,M} &= \frac{4}{3} (5 D^2+9 F^2) \ ,
\end{align}
\end{subequations}
are independent of the baryon/meson species. This yields similar baryon masses and $Z$-factors along the line of equal quark masses and is a consequence of $\operatorname{SU}(3)_f$ symmetry. For a detailed study of baryon masses under symmetry breaking see~\cite{Bruns:2012eh}. The square root of the $Z$-factor needed in our calculation is given by%
\begin{align}
\sqrt{Z_B} &\mathrel{\dot =} 1+\frac{1}{2}\Sigma^\prime_B \ ,
\end{align}
where the prime indicates taking a derivative with respect to $\slashed p$ and substituting $\slashed{p}\rightarrow m_B$, while%
\begin{align}
\begin{split}
\Sigma^\prime_B(\bar m_q, \delta m_l) &= 3 g_{B,\pi} f^\prime(m_\pi,m_0,m_B) + 4 g_{B,K} f^\prime(m_K,m_0,m_B) + g_{B,\eta} f^\prime(m_\eta,m_0,m_B) \\
&\equiv \Sigma^{\prime\star}(\bar m_q) + \Delta \Sigma^\prime_B(\bar m_q, \delta m_l) \ ,
\end{split}
\end{align}
with%
\begin{subequations}
\begin{align}
\begin{split}
\Sigma^{\prime\star}(\bar m_q) &= \Sigma^\prime_B(\bar m_q, 0) = (3 g_{B,\pi}+4 g_{B,K}+g_{B,\eta}) f^\prime(m_m^\star,m_0,m_B) \\
&= \frac{4}{3} (5 D^2+9 F^2) f^\prime(m_m^\star,m_0,m_B) \mathrel{\dot =}\frac{4}{3} (5 D^2+9 F^2) f^\prime(m_m^\star,m_b^\star,m_b^\star) \ ,
\end{split} \raisetag{1cm} \\
\begin{split}
\Delta\Sigma^{\prime}_B(\bar m_q, \delta m_l) &= \Sigma^\prime_B(\bar m_q, \delta m_l) -\Sigma^\prime_B(\bar m_q, 0) \\
&= 3 g_{B,\pi} f^\prime(m_\pi,m_0,m_B) + 4 g_{B,K} f^\prime(m_K,m_0,m_B) + g_{B,\eta} f^\prime(m_\eta,m_0,m_B)\\
&\quad - \frac{4}{3} (5 D^2+9 F^2) f^\prime(m_m^\star,m_0,m_B) \\
&\mathrel{\dot =} 3 g_{B,\pi} f^\prime(m_\pi,m_b^\star,m_b^\star) + 4 g_{B,K} f^\prime(m_K,m_b^\star,m_b^\star) + g_{B,\eta} f^\prime(m_\eta,m_b^\star,m_b^\star)\\
&\quad - \frac{4}{3} (5 D^2+9 F^2) f^\prime(m_m^\star,m_b^\star,m_b^\star) \ ,
\end{split} \raisetag{0.4cm}
\end{align}
\end{subequations}
where $m_{m/b}^\star = m_{m/b}^\star (\bar m_q)$ is the meson/baryon mass along the symmetric line ($\delta m_l = 0$). The dotted equal sign $\mathrel{\dot =}$ means equal up to terms which are of higher order than our level of accuracy (which is second order in chiral power counting). For explicit results see appendix~\ref{sect_gfunctions}.%
\subsection{Baryon-to-vacuum matrix elements of three-quark operators}
In this section we describe the actual loop calculation. From a simple power counting argument one finds that at leading one-loop level the only contributing graphs are the ones shown in figure~\ref{FDDA}. One easily observes that the second order operator insertions only occur without additional mesons. Therefore we only have to compute the vertices where a single baryon couples to the operator. Contributions with additional mesons only occur for the leading order operator insertion ($j=1$). For the BChPT calculation mainly the structure $B^{j,k,XYZ}_{\delta,abc}$ is relevant. Graph \subref{fdda:subfigd} of figure~\ref{FDDA} is an exception because the extra $\gamma_5$ from the baryon-meson-baryon vertex has to be canceled with a $\gamma_5$ from the Dirac structure of the operator. The calculation gets simplified considerably if one uses the fact that (by construction) the $B^{j,k,XYZ}_{\delta,abc}$ with $k\neq 1$ can be obtained from the case $k=1$ by a permutation of indices:%
\begin{figure}[t]
\centering%
\subfigure[\label{fdda:subfiga}]{
\includegraphics[scale=0.55]{pics/FDDAa.eps}
}
\subfigure[\label{fdda:subfigb}]{
\includegraphics[scale=0.55]{pics/FDDAb.eps}
}
\subfigure[\label{fdda:subfigc}]{
\includegraphics[scale=0.55]{pics/FDDAc.eps}
}
\subfigure[\label{fdda:subfigd}]{
\includegraphics[scale=0.55]{pics/FDDAd.eps}
}
\caption[]{\label{FDDA} Feynman diagrams needed for the calculation of the baryon-to-vacuum matrix elements. The squares depict the operator insertions given in eqs.~\eqref{operator_insertions}, the circle stands for the vertex from the meson-baryon Lagrangian given in eq.~\eqref{BMB_Vertex} and the dashed/solid lines represent mesons/baryons. Diagram \subref{fdda:subfiga} has to be multiplied with $\sqrt{Z}$. However, one knows that at higher orders all of the diagrams will receive a $\sqrt{Z}$ contribution, which can be used as an argument in favor of the factorized version of our results (see eq.~\eqref{DAs_chpt_Result_factorized} in section~\ref{sect_result}).}
\end{figure}%
\begin{align}
B^{j,2,XYZ}_{\delta,abc}&=B^{j,1,YZX}_{\delta,bca} \ , &
B^{j,3,XYZ}_{\delta,abc}&=B^{j,1,ZXY}_{\delta,cab} \ , &
B^{j,4,XYZ}_{\delta,abc}&=-B^{j,1,YXZ}_{\delta,bac} \ , \nonumber \\*
B^{j,5,XYZ}_{\delta,abc}&=-B^{j,1,XZY}_{\delta,acb} \ , &
B^{j,6,XYZ}_{\delta,abc}&=-B^{j,1,ZYX}_{\delta,cba} \ ,
\end{align}
which means that we actually only have to calculate the case $k=1$. Defining%
\begin{align}
(-1)_X &\equiv \begin{cases}
+1 \ , & \text{for } X=R \\ -1 \ , & \text{for } X=L
\end{cases} \ ,
\end{align}
we can write down the relevant operator insertions in a quite economic way:%
\begin{subequations} \label{operator_insertions}
\begin{align}
B^{1,1,XYZ}_{\delta,abc}(z) \bigg|_{B_\epsilon (p)}&= (\kappa_B)^{a a^\prime} \varepsilon^{a^\prime b c} e^{-i p \cdot z} \delta_{\delta \epsilon} \ , \\
B^{1,1,XYZ}_{\delta,abc}(z) \bigg|_{B_\epsilon (p-q) \phi^k(q)}&= \frac{i}{2 F_0} \Bigl[ \begin{aligned}[t] & (-1)_X (\lambda^k \kappa_B)^{a a^\prime} \delta^{bb^\prime} \delta^{cc^\prime} + (-1)_Y ( \kappa_B)^{a a^\prime} (\lambda^k)^{bb^\prime} \delta^{cc^\prime} \\ & + (-1)_Z (\kappa_B)^{a a^\prime} \delta^{bb^\prime} (\lambda^k )^{cc^\prime} \Bigr] \varepsilon^{a^\prime b^\prime c^\prime} e^{-i p \cdot z} \delta_{\delta \epsilon} \ , \end{aligned}
\end{align}
\begin{align}
\begin{split}
\MoveEqLeft B^{1,1,XYZ}_{\delta,abc}(z) \bigg|_{B_\epsilon (p-q_1-q_2) \phi^k(q_1) \phi^l(q_2)}=\\
&= \frac{-1}{16 F_0^2} \Bigl[ \begin{aligned}[t] & (\{\lambda^k,\lambda^l\} \kappa_B)^{a a^\prime} \delta^{bb^\prime} \delta^{cc^\prime} + ( \kappa_B)^{a a^\prime} (\{\lambda^k,\lambda^l\})^{bb^\prime} \delta^{cc^\prime} + (\kappa_B)^{a a^\prime} \delta^{bb^\prime} (\{\lambda^k,\lambda^l\})^{cc^\prime} \\
&+ 2 (-1)_X (-1)_Y (\lambda^k \kappa_B)^{a a^\prime} (\lambda^l)^{bb^\prime} \delta^{cc^\prime}
+ 2 (-1)_X (-1)_Z (\lambda^k \kappa_B)^{a a^\prime} \delta^{bb^\prime} (\lambda^l)^{cc^\prime} \\
&+ 2 (-1)_Y (-1)_Z (\kappa_B)^{a a^\prime} (\lambda^k )^{bb^\prime} (\lambda^l)^{cc^\prime}
+ 2 (-1)_X (-1)_Y (\lambda^l \kappa_B)^{a a^\prime} (\lambda^k)^{bb^\prime} \delta^{cc^\prime}\\
&+ 2 (-1)_X (-1)_Z (\lambda^l \kappa_B)^{a a^\prime} \delta^{bb^\prime} (\lambda^k)^{cc^\prime}
+ 2 (-1)_Y (-1)_Z (\kappa_B)^{a a^\prime} (\lambda^l )^{bb^\prime} (\lambda^k)^{cc^\prime}
\Bigr] \\
& \times \varepsilon^{a^\prime b^\prime c^\prime} e^{-i p \cdot z} \delta_{\delta \epsilon} \ . \end{aligned}
\end{split}
\end{align}
The second order tree-level operator insertions read%
\begin{align}
B^{2,1,XYZ}_{\delta,abc}(z) \bigg|_{B_\epsilon (p)}&= 4 B_0 m_0^{-2} \tr{\mathcal{M}}(\kappa_B)^{a a^\prime} \varepsilon^{a^\prime b c} e^{-i p \cdot z} \delta_{\delta \epsilon} \ , \\
B^{3,1,XYZ}_{\delta,abc}(z) \bigg|_{B_\epsilon (p)}&= 4 B_0 m_0^{-2} (\kappa_B)^{a a^\prime} \tilde{\mathcal{M}}^{b b^\prime} \varepsilon^{a^\prime b^\prime c} e^{-i p \cdot z} \delta_{\delta \epsilon} \ ,
\end{align}
\end{subequations}
where $\tilde{\mathcal{M}}= \mathcal{M}-\smash{\tr{\mathcal{M}}}/3$. After performing the loop calculation one finds that the results can be expressed as%
\begin{subequations}
\begin{align}
\begin{split}
\MoveEqLeft \langle 0 | \mathcal O^{abc}_{RR,\alpha \beta \gamma} {(a_1,a_2,a_3)} + \mathcal O^{abc}_{LL,\alpha \beta \gamma}{(a_1,a_2,a_3)}| B(p,s) \rangle= \\
&= \int [dx] e^{- i \; n \cdot p \sum_k x_k a_k} \sum_i \Gamma^{i,\text{even}}_{\alpha\beta\gamma\delta} u^B_\delta (p,s) h_{B,\text{even}}^{i,abc}{(x_1,x_2,x_3)} \ ,
\end{split} \\
\begin{split}
\MoveEqLeft\langle 0 | \mathcal O^{abc}_{RL,\alpha \beta \gamma} {(a_1,a_2,a_3)} + \mathcal O^{abc}_{LR,\alpha \beta \gamma}{(a_1,a_2,a_3)}| B(p,s) \rangle =\\
&= \int [dx] e^{- i \; n \cdot p \sum_k x_k a_k} \sum_i \Gamma^{i,\text{odd}}_{\alpha\beta\gamma\delta} u^B_\delta (p,s) h_{B,\text{odd}}^{i,abc}{(x_1,x_2,x_3)} \ ,
\end{split}
\end{align}
\end{subequations}
where $u^B_\delta (p,s)$ is the baryon spinor,%
\begin{subequations} \label{gamma_structure_odd_even}
\begin{align}
\Gamma^{i,\text{even}}_{\alpha\beta\gamma\delta} &= \Gamma^{i,RRR}_{\alpha\beta\gamma\delta}-\Gamma^{i,LLL}_{\alpha\beta\gamma\delta} \ , \\
\Gamma^{i,\text{odd}}_{\alpha\beta\gamma\delta} &= \Gamma^{i,LLR}_{\alpha\beta\gamma\delta}-\Gamma^{i,RRL}_{\alpha\beta\gamma\delta} \ ,
\end{align}
\end{subequations}
and%
\begin{subequations} \label{def_h_functions}
\begin{align}
h^{i,abc}_{B,\text{even}}{(x_1,x_2,x_3)} &= \sum_{j,k} c^{j,k,abc}_{B,RRR} \mathcal F^{i,j,k}_\text{even} {(x_1,x_2,x_3)} \ , \\
h^{i,abc}_{B,\text{odd}} {(x_1,x_2,x_3)} &= \sum_{j,k} c^{j,k,abc}_{B,LLR} \mathcal F^{i,j,k}_\text{odd} {(x_1,x_2,x_3)} \ .
\end{align}
\end{subequations}
The coefficients $c^{j,k,abc}_{B,XYZ}$ inherit the property that the ones with $k\neq 1$ can be obtained from the case $k=1$ by a permutation of indices:%
\begin{align}
c^{j,2,abc}_{B,XYZ}&= c^{j,1,bca}_{B,YZX} \ , &
c^{j,3,abc}_{B,XYZ}&= c^{j,1,cab}_{B,ZXY} \ , &
c^{j,4,abc}_{B,XYZ}&=- c^{j,1,bac}_{B,YXZ} \ , \nonumber \\*
c^{j,5,abc}_{B,XYZ}&=- c^{j,1,acb}_{B,XZY} \ , &
c^{j,6,abc}_{B,XYZ}&=- c^{j,1,cba}_{B,ZYX} \ .
\end{align}
For those with $k=1$ we find%
\begin{subequations}
\begin{align}
c^{1,1,abc}_{B,XYZ} &= c^{1,1,abc}_{B,XYZ}\bigg|_{(a)} +c^{1,1,abc}_{B,XYZ}\bigg|_{(c)} + c^{1,1,abc}_{B,XYZ}\bigg|_{(d)} \ , \\
c^{1,1,abc}_{B,XYZ}\bigg|_{(a)} &= \sqrt{Z_B} (\kappa_B)^{aa^\prime} \varepsilon^{a^\prime b c} \ , \\
c^{1,1,abc}_{B,XYZ}\bigg|_{(c)} &= \frac{-1}{8 F_0^2} \sum_k
\Bigl[ \begin{aligned}[t] & (\lambda^k \lambda^k \kappa_B)^{a a^\prime} \delta^{bb^\prime} \delta^{cc^\prime} + ( \kappa_B)^{a a^\prime} (\lambda^k \lambda^k)^{bb^\prime} \delta^{cc^\prime} + (\kappa_B)^{a a^\prime} \delta^{bb^\prime} (\lambda^k \lambda^k)^{cc^\prime} \\
&+ 2 (-1)_X (-1)_Y (\lambda^k \kappa_B)^{a a^\prime} (\lambda^k)^{bb^\prime} \delta^{cc^\prime} \\
&+ 2 (-1)_X (-1)_Z (\lambda^k \kappa_B)^{a a^\prime} \delta^{bb^\prime} (\lambda^k)^{cc^\prime} \\
&+ 2 (-1)_Y (-1)_Z (\kappa_B)^{a a^\prime} (\lambda^k )^{bb^\prime} (\lambda^k)^{cc^\prime}
\Bigr] \varepsilon^{a^\prime b^\prime c^\prime} I_{10}{(m_l)} \ , \end{aligned} \\
\begin{split}
c^{1,1,abc}_{B,XYZ}\bigg|_{(d)} &= \frac{-1}{4 F_0^2} \sum_{k,{\tilde{B}}}
\Bigl[ \begin{aligned}[t] & (-1)_X (\lambda^k \kappa_{\tilde{B}})^{a a^\prime} \delta^{bb^\prime} \delta^{cc^\prime} +(-1)_Y ( \kappa_{\tilde{B}})^{a a^\prime} (\lambda^k)^{bb^\prime} \delta^{cc^\prime} \\
& +(-1)_Z ( \kappa_{\tilde{B}})^{a a^\prime} \delta^{bb^\prime} (\lambda^k)^{cc^\prime} \Bigr] \tr{\kappa_{\tilde{B}}^T(D\{\lambda^k,\kappa_B\}+F[\lambda^k,\kappa_B])} \end{aligned} \\
& \ \times\Bigl( I_{10}{(m_k)} + (m_B^2-m_0^2) I_{11}{(m_k,m_0,m_B)} - m_B(m_B+m_0) I^{(1)}_{11}{(m_k,m_0,m_B)} \Bigr) \ .
\end{split}
\end{align}
In the contribution from graph \subref{fdda:subfigd} commuting $\gamma_5$ from the vertex with the Dirac structure in the operator yields $\eta^5_{\Gamma^B,i}(-1)^{d^m_i}=-1$ (compare table~\ref{symmetry_properties_Gamma}). In operators of type $\mathcal O_{XR}$ the $\gamma_5$ has no effect owing to $\gamma_R \gamma_5=\gamma_R$. The relative sign in the vertex in operators of type $\mathcal O_{\bar XL}$ is compensated by $\gamma_L \gamma_5 = - \gamma_L$. Therefore the result only contains structures of the form given in eq.~\eqref{gamma_structure_odd_even}. This is no coincidence but has to happen in order to obtain a result that behaves correctly under parity transformation. For the second order tree-level contributions we find%
\begin{align}
c^{2,1,abc}_{B,XYZ} &= 4 B_0 m_0^{-2} \tr{\mathcal{M}} (\kappa_B)^{a a^\prime} \varepsilon^{a^\prime b c} \ , \\
c^{3,1,abc}_{B,XYZ} &= 4 B_0 m_0^{-2} (\kappa_B)^{a a^\prime} \mathcal{\tilde M}^{bb^\prime} \varepsilon^{a^\prime b^\prime c} \ .
\end{align}
\end{subequations}
Using eq.~\eqref{3q_op_decompRL} the matrix element of the complete three-quark operator reads%
\begin{align} \label{result_not_projected}
\begin{split}
\MoveEqLeft[1]\langle 0 | q^a_\alpha (a_1 n) q^b_\beta (a_2 n) q^c_\gamma (a_3 n) | B(p,s) \rangle= \\
&= \int [dx] e^{- i \; n \cdot p \sum_k x_k a_k} \sum_i
\Bigl( \begin{aligned}[t] & \Gamma^{i,\text{even}}_{\alpha\beta\gamma\delta} h_{B,\text{even}}^{i,abc}{(x_1,x_2,x_3)} + \Gamma^{i,\text{odd}}_{\alpha\beta\gamma\delta} h_{B,\text{odd}}^{i,abc}{(x_1,x_2,x_3)}\\
& + \Gamma^{i,\text{odd}}_{\gamma\alpha\beta\delta} h_{B,\text{odd}}^{i,cab}{(x_3,x_1,x_2)} + \Gamma^{i,\text{odd}}_{\beta\gamma\alpha\delta} h_{B,\text{odd}}^{i,bca}{(x_2,x_3,x_1)} \Bigr) u^B_\delta (p,s) \ . \end{aligned}
\end{split}
\end{align}
\subsection{Projection onto standard DAs} \label{sect_matching}
In this section we relate our parametrization of the baryon-to-vacuum matrix element, which was guided by the behaviour under chiral rotations, to the general decomposition given in ref.~\cite{Braun:2000kw}, which is more convenient for daily use. To do so we have contracted both our result (eq.~\eqref{result_not_projected}) and formula (2.3) of ref.~\cite{Braun:2000kw} with Dirac structures of the form $\Gamma^A_{\alpha\beta}\otimes \Gamma^B_{\gamma^\prime \gamma}$. It is sufficient to use structures where Lorentz indices are either contracted between $\Gamma^A$ and $\Gamma^B$ or with the light-cone vector $n$ or the momentum $p$. Afterwards we have used the identity $\slashed{p} u^B(p)=m_B u^B(p)$ and have matched the prefactors of the remaining Dirac structures ($\gamma_5$ and $\slashed{n} \gamma_5$). Using twist-projection, we obtain the results for the distribution amplitudes $S_i^B$, $P_i^B$, $A_i^B$, $V_i^B$ and $T_i^B$ which are independent of the scalar product $n \cdot p$, due to the scaling property described in section~\ref{sect_sym_3qopp} (For the details of the twist-projection we refer to \cite{Braun:2000kw}). We have collected these lengthy matching relations in appendix~\ref{app_projection_standard_DAs}. The amplitudes have the following symmetry properties under exchange of the first and the second variable%
\begin{align}
S_i^B{(x_1,x_2,x_3)} &= - (-1)_B S_i^B{(x_2,x_1,x_3)} \ , \nonumber \\
P_i^B{(x_1,x_2,x_3)} &= - (-1)_B P_i^B{(x_2,x_1,x_3)} \ , \nonumber \\
A_i^B{(x_1,x_2,x_3)} &= - (-1)_B A_i^B{(x_2,x_1,x_3)} \ , \nonumber \\
V_i^B{(x_1,x_2,x_3)} &= + (-1)_B V_i^B{(x_2,x_1,x_3)} \ , \nonumber \\*
T_i^B{(x_1,x_2,x_3)} &= + (-1)_B T_i^B{(x_2,x_1,x_3)} \ , \label{symmetry_standard_DAs_12}
\end{align}
where we use%
\begin{align}
(-1)_B &\equiv \begin{cases} +1 \ , & \text{for } B \neq \Lambda \\ -1 \ , & \text{for } B = \Lambda \end{cases} \ ,
\end{align}
for brevity. To obtain these nice symmetry properties one has to choose the flavor content in the operator as $p \mathrel{ \hat = } uud$, $n \mathrel{ \hat = } ddu$, $\Sigma^+ \mathrel{ \hat = } uus$, $\Sigma^0 \mathrel{ \hat = } uds$, $\Sigma^- \mathrel{ \hat = } dds$, $\Xi^0 \mathrel{ \hat = } ssu$, $\Xi^- \mathrel{ \hat = } ssd$, $\Lambda \mathrel{ \hat = } uds$, where the order of the flavors is relevant. The different sign for the $\Lambda$ originates from the antisymmetry of the isospin singlet state.%
\section{Results} \label{sect_result}
In this section we present our results and provide a definition for DAs that do not mix under chiral extrapolation. In section~\ref{sect_example_of_application} we work out an explicit parametrization of baryon octet DAs, where we follow the approach presented in refs.~\cite{Anikin:2013aka,Anikin:2013yoa,Braun:2008ia}.
\subsection{General strategy and choice of distribution amplitudes}
We will split up every distribution amplitude in the following way:%
\begin{subequations}
\begin{align}
\begin{split}
\text{DA}(\bar m_q, \delta m_l) &= \text{DA}(\bar m_q, 0) + \bigl(\text{DA}(\bar m_q, \delta m_l) - \text{DA}(\bar m_q, 0)\bigr) \\
&\equiv \text{DA}^\star (\bar m_q) + \Delta \text{DA} (\bar m_q, \delta m_l) \ ,
\end{split} \\
\begin{split}
\text{DA}^\star(\bar m_q) &= \text{DA}^\star(0) + \bigl(\text{DA}^\star(\bar m_q) - \text{DA}^\star(0)\bigr) \\
&\equiv \text{DA}^\circ + \Delta \text{DA}^\star(\bar m_q) \ ,
\end{split}
\end{align}
\end{subequations}
where the main idea is to use the second formula to parametrize everything in terms of the DAs at the symmetric point, which are measurable on the lattice as opposed to the amplitudes in the chiral limit. Lattice simulations where the mean quark mass is fixed at its physical value while $\delta m_l$ is varied are already available for hadron masses and some form factors~\cite{Bietenholz:2011qq,Gockeler:2011ze,Cooke:2013qqa}. Corresponding simulations for the baryon octet DAs treated in this work are in progress. This strategy has the additional advantage that one gets rid of the parameters that describe the behaviour under variation of the mean quark mass. For the presentation of the results it turns out to be convenient to write down the second order tree-level and the loop contribution separately. We define for all baryons%
\begin{align}
\Delta \text{DA} = \Delta \text{DA}^{\text{loop}} + \delta m \ \Delta \text{DA}^{\text{tree}} \ ,
\end{align}
where
\begin{align}
\delta m&=\frac{4 B_0 \delta m_l}{{m_b^\star}^2} \ .
\end{align}
Then we use the fact that we can rewrite $\Delta \text{DA}$ in terms of $m_b^\star$ and $\text{DA}^\star$ using the corresponding expansions in $\bar m_q$. For a specific set of DAs, which do not mix under chiral extrapolation (see below), this allows us to rewrite the loop contribution as the DA along the symmetric line multiplied with a loop function $f$ such that the results have the form%
\begin{align}
\begin{split}
\text{DA}(\bar m_q, \delta m_l) &= \text{DA}^\star (\bar m_q) (1+f) + \delta m \ \Delta \text{DA}^{\text{tree}} \ .
\end{split}
\end{align}
By virtue of $\operatorname{SU}(3)_f$ symmetry we find the following relations between DAs along the line of symmetric quark masses $m_u=m_d=m_s$
\begin{subequations} \label{su3constraints}
\begin{align}
\begin{split}
2 T_{1/6}^{B\star} { (x_1,x_2,x_3)} &= (-1)_B \bigl[V_{1/6}^{B\star}-A_{1/6}^{B\star}\bigr]{ (x_1,x_3,x_2)} \\&\quad+ \bigl[V_{1/6}^{B\star}-A_{1/6}^{B\star}\bigr]{ (x_2,x_3,x_1)} \ ,
\end{split} \\*
\begin{split}
\bigl[T_{3/4}^{B\star} + T_{7/8}^{B\star} + S_{1/2}^{B\star} - P_{1/2}^{B\star}\bigr]{ (x_1,x_2,x_3)} &= \bigl[V_{2/5}^{B\star}-A_{2/5}^{B\star}\bigr]{ (x_2,x_3,x_1)} \\&\quad+ \bigl[V_{3/4}^{B\star}-A_{3/4}^{B\star}\bigr]{ (x_3,x_1,x_2)} \ ,
\end{split} \\*
\begin{split}
2 T_{2/5}^{B\star} { (x_1,x_2,x_3)} &= \bigl[T_{3/4}^{B\star} - T_{7/8}^{B\star} + S_{1/2}^{B\star} + P_{1/2}^{B\star}\bigr]{ (x_3,x_1,x_2)}\\&\quad + \bigl[T_{3/4}^{B\star} - T_{7/8}^{B\star} + S_{1/2}^{B\star} + P_{1/2}^{B\star}\bigr]{ (x_3,x_2,x_1)} \ .
\end{split}
\end{align}
\end{subequations}
Note that we do not impose these relations. They are automatically fulfilled by our calculation (loop contributions included). For the nucleons these relations are fulfilled exactly also for $\delta m_l \neq 0$ owing to isospin symmetry (again this is also true for the loop contributions), which was already shown in ref.~\cite{Braun:2000kw}. If we were only interested in the $\operatorname{SU}(3)_f$ symmetric case (or in nucleons only), it would therefore be enough to define the independent amplitudes as%
\begin{subequations}\label{original_DAs_definition}
\begin{align}
\Phi_{3/6}^B(x_1,x_2,x_3)&=\bigl[ V_{1/6}^B-A_{1/6}^B \bigr](x_1,x_2,x_3) \ , \\
\Phi_{4/5}^B(x_1,x_2,x_3)&=\bigl[ V_{2/5}^B-A_{2/5}^B \bigr](x_1,x_2,x_3) \ , \\
\Psi_{4/5}^B(x_1,x_2,x_3)&=\bigl[ V_{3/4}^B-A_{3/4}^B \bigr](x_1,x_2,x_3) \ , \\
\Xi_{4/5}^B(x_1,x_2,x_3)&=\bigl[ T_{3/4}^B-T_{7/8}^B+S_{1/2}^B+P_{1/2}^B \bigr](x_1,x_2,x_3) \ ,
\end{align}
\end{subequations}
where the $\Phi_i^B$ and $\Psi_i^B$ describe the coupling to chiral-odd operators, while the $\Xi_i^B$ describe the chiral-even sector. The subscript indicates the twist. As it turns out the amplitudes $\Phi_i^B$, $\Psi_i^B$ and $\Xi_i^B$ are not yet the optimal choice for a description of the complete baryon octet, since they mix under chiral extrapolation. Additionally one finds that it is very convenient to use differing definitions for the $\Lambda$, which we choose in such a way that the DAs of the $\Lambda$ coincide with the DAs of the other octet baryons in the limit of equal quark masses. Therefore we define%
\begin{subequations} \label{definition_superior_DAs}
\begin{align}
\Phi_{\pm, 3/6}^B(x_1,x_2,x_3)&=\frac{c^\pm_B}{2} \bigl(\bigl[ V_{1/6}^B-A_{1/6}^B \bigr](x_1,x_2,x_3) \pm \bigl[ V_{1/6}^B-A_{1/6}^B \bigr](x_3,x_2,x_1) \bigr) \ , \\
\Phi_{\pm, 4/5}^B(x_1,x_2,x_3)&=c^\pm_B \bigl(\bigl[ V_{2/5}^B-A_{2/5}^B \bigr](x_1,x_2,x_3) \pm (-1)_B \bigl[ V_{3/4}^B-A_{3/4}^B \bigr](x_2,x_3,x_1)\bigr) \ , \\
\Xi_{\pm, 4/5}^B(x_1,x_2,x_3)&=3(-1)_B c^\pm_B \bigl( \begin{aligned}[t] & \bigl[ T_{3/4}^B-T_{7/8}^B+S_{1/2}^B+P_{1/2}^B \bigr](x_1,x_2,x_3) \\& \pm \bigl[ T_{3/4}^B-T_{7/8}^B+S_{1/2}^B+P_{1/2}^B \bigr](x_1,x_3,x_2) \bigr) \ , \end{aligned}
\end{align}
\end{subequations}%
where%
\begin{align}
c^+_B & = \begin{cases} 1 \ , & \text{for } B\neq\Lambda \\ \sqrt{\frac{2}{3}} \ , & \text{for } B=\Lambda\end{cases} \ , & c^-_B & = \begin{cases} 1 \ , & \text{for } B\neq\Lambda \\ -\sqrt{6} \ , & \text{for } B=\Lambda \end{cases} \ .
\end{align}
Being interested in $\operatorname{SU}(3)_f$ violation one can not use the constraints given in eq.~\eqref{su3constraints} and therefore one needs six additional DAs. Our choice are (up to differing prefactors for the $\Lambda$ and exchange of variables) the left-hand sides in eq.~\eqref{su3constraints} since they coincide with the DAs in eq.~\eqref{definition_superior_DAs} in the $\operatorname{SU(3)}_f$ symmetric limit. We define%
\begin{subequations} \label{definition_superior_DAs_second_part}
\begin{align}
\Pi_{3/6}^B(x_1,x_2,x_3)&=c_B^- (-1)_B \; T_{1/6}^B (x_1,x_3,x_2) \ , \\
\Pi_{4/5}^B(x_1,x_2,x_3)&=c_B^- \bigl[T_{3/4}^B + T_{7/8}^B + S_{1/2}^B - P_{1/2}^B\bigr](x_3,x_1,x_2) \ , \\
\Upsilon_{4/5}^B(x_1,x_2,x_3)&= 6 c_B^- \; T_{2/5}^B(x_3,x_2,x_1) \ ,
\end{align}%
\end{subequations}%
where the $\Pi_i$ describe the chiral-odd sector, while the $\Upsilon_i$ describe the chiral-even part. For each octet baryon the standard DAs can be decomposed into the amplitudes defined in eqs.~\eqref{definition_superior_DAs} and~\eqref{definition_superior_DAs_second_part} (see appendix~\ref{app_handbook_of_DAs}). We find that the DAs for different nucleons, $\Sigma$'s and $\Xi$'s are related to each other exactly by isospin symmetry. Therefore we define%
\begin{subequations} \label{baryon_signs}
\begin{align}
\text{DA}^N &\equiv \text{DA}^p = - \text{DA}^n \ , \\
\text{DA}^\Sigma &\equiv \text{DA}^{\Sigma^-} = - \text{DA}^{\Sigma^+} = \sqrt{2} \text{DA}^{\Sigma^0} \ , \\
\text{DA}^\Xi &\equiv \text{DA}^{\Xi^0} = - \text{DA}^{\Xi^-} \ ,
\end{align}
\end{subequations}
and give the results only for $\text{DA}^N$, $\text{DA}^\Sigma$, $\text{DA}^\Xi$ and $\text{DA}^\Lambda$. In the $\operatorname{SU}(3)_f$ symmetric limit all these DAs can be related to those of the nucleon:%
\begin{subequations}
\begin{align}
\Phi_{+, i}^\star &\equiv \Phi_{+, i}^{N\star} = \Phi_{+, i}^{\Sigma\star} = \Phi_{+, i}^{\Xi\star} = \Phi_{+, i}^{\Lambda\star} = \Pi_{ i}^{N\star} = \Pi_{i}^{\Sigma\star} = \Pi_{i}^{\Xi\star} \ , \\
\Phi_{-, i}^\star &\equiv \Phi_{-, i}^{N\star} = \Phi_{-, i}^{\Sigma\star} = \Phi_{-, i}^{\Xi\star} = \Phi_{-, i}^{\Lambda\star} = \Pi_{ i}^{\Lambda\star}\ , \\
\Xi_{+, i}^\star &\equiv \Xi_{+, i}^{N\star} = \Xi_{+, i}^{\Sigma\star} = \Xi_{+, i}^{\Xi\star} = \Xi_{+, i}^{\Lambda\star}=\Upsilon_{ i}^{N\star} = \Upsilon_{i}^{\Sigma\star} = \Upsilon_{ i}^{\Xi\star} \ , \\
\Xi_{-, i}^\star &\equiv \Xi_{-, i}^{N\star} = \Xi_{-, i}^{\Sigma\star} = \Xi_{-, i}^{\Xi\star} = \Xi_{-, i}^{\Lambda\star}=\Upsilon_{ i}^{\Lambda\star} \ .
\end{align}
\end{subequations}
\subsection{Minimal parametrization of baryon octet distribution amplitudes} \label{sect_minimal_parametrization}
The choice of DAs presented in the previous section allows us to write down our results in a very compact form:%
\begin{subequations} \label{DAs_chpt_Result}
\begin{align}
\Phi_{\pm, i}^B &= \Phi_{\pm, i}^\star \Bigl( 1 + \tfrac{1}{2} \Delta\Sigma^\prime_B + \Delta g^B_{\Phi \pm} \Bigr) + \delta m \ \Delta \Phi_{\pm,i}^B \ , \\*
\Xi_{\pm, i}^B &= \Xi_{\pm, i}^\star \Bigl( 1 + \tfrac{1}{2} \Delta\Sigma^\prime_B + \Delta g^B_{\Xi} \Bigr) + \delta m \ \Delta \Xi_{\pm,i}^B \ , \\
\Pi_{i}^B &= \Phi_{\pm_B, i}^\star \Bigl( 1 + \tfrac{1}{2} \Delta\Sigma^\prime_B + \Delta g^B_{\Pi} \Bigr) + \delta m \ \Delta \Pi_{i}^B \ , \\
\Upsilon_{i}^B &= \Xi_{\pm_B, i}^\star \Bigl( 1 + \tfrac{1}{2} \Delta\Sigma^\prime_B + \Delta g^B_{\Xi} \Bigr) + \delta m \ \Delta \Upsilon_{i}^B \ ,
\end{align}
\end{subequations}
where ``$\pm_B$'' stands for ``$+$'' if $B\neq\Lambda$ and for ``$-$'' if $B=\Lambda$. The second term in the equations above originates from quark mass insertions, while the first term (or, to be more precise, $\Delta\Sigma^\prime_B$ and $\Delta g^B_{\text{DA}}$) is generated by meson loops and contains chiral logarithms. Owing to our choice of DAs the functions $\Delta g^B_{\text{DA}}$, which are listed in appendix~\ref{sect_gfunctions} together with $\Delta\Sigma^\prime_B$, do not depend on the twist of the amplitude. $\Delta g^B_{\text{DA}}$ and $\Delta\Sigma^\prime_B$ vanish for equal quark masses ($\delta m = 0$). The nontrivial dependence on the mean quark mass of the distribution amplitudes $\Phi_{\pm, i}^\star$ and $\Xi_{\pm, i}^\star$ is presented in section~\ref{sect_mean_quark}. The amplitudes describing the tree-level contribution to the $\operatorname{SU}(3)_f$ symmetry breaking are not completely free. It holds for all distribution amplitudes
\begin{align} \label{DeltaXi}
\Delta \text{DA}^{\Xi} = - \Delta \text{DA}^{N}- \Delta \text{DA}^{\Sigma} \ .
\end{align}
Furthermore, the amplitudes $\Delta \Pi_i^B$ and $\Delta \Upsilon_i^B$ can be expressed in terms of $\Delta \Phi_{\pm, i}^B$ and $\Delta \Xi_{\pm, i}^B$:%
\begin{subequations}\label{constraints_on_Delta_DAs}
\begin{align}
\Delta \Pi_i^{N} &= \Delta \Phi_{+,i}^{N} \ , &
\Delta \Upsilon_i^{N} &= \Delta \Xi_{+,i}^{N} \ , \\*
\Delta \Pi_i^{\Sigma} &= -\frac{1}{2} \Delta \Phi_{+,i}^{\Sigma} -\frac{3}{2} \Delta \Phi_{+,i}^{\Lambda} \ , &
\Delta \Upsilon_i^{\Sigma} &= -\frac{1}{2}\Delta \Xi_{+,i}^{\Sigma} - \frac{3}{2} \Delta \Xi_{+,i}^{\Lambda} \ , \\*
\Delta \Pi_i^{\Lambda} &= -\frac{1}{2} \Delta \Phi_{-,i}^{\Lambda} -\frac{3}{2} \Delta \Phi_{-,i}^{\Sigma} \ , &
\Delta \Upsilon_i^{\Lambda} &= -\frac{1}{2}\Delta \Xi_{-,i}^{\Lambda} - \frac{3}{2} \Delta \Xi_{-,i}^{\Sigma} \ ,
\end{align}
\end{subequations}%
which means that the $\Pi^B_i$ and $\Upsilon^B_i$ are completely fixed by the other amplitudes. The divergencies of leading one-loop order contained in $\Delta \Sigma^\prime_B$ and $\Delta g^B_{\text{DA}}$ can be canceled by the introduction of counterterms%
\begin{align}\label{counterterms}
\Delta \Phi^B_{\pm,i} &\longrightarrow \frac{{m_b^\star}^2 c^B_{\Phi \pm}}{24 F_\star^2} \Phi^{\star}_{\pm,i} L + \Delta \Phi^{B,\text{ren.}}_{\pm,i}(\mu) \ , &
\Delta \Xi^B_{\pm,i} &\longrightarrow \frac{{m_b^\star}^2 c^B_{\Xi}}{24 F_\star^2} \Xi^{\star}_{\pm,i} L + \Delta \Xi^{B,\text{ren.}}_{\pm,i}(\mu) \ ,
\end{align}
where $L$ contains the divergence and the typical constants of the modified minimal subtraction scheme (see eq.~\eqref{def_divergence}). $F_\star$ is the meson decay constant in the $\operatorname{SU(3)}_f$ symmetric limit. The coefficients $c^B_{\text{DA}}$ are given by%
\begin{align}
c^{N}_{\Phi \pm} &= -9 (D^2+10 D F - 3 F^2) -23 \mp 24 \ , &
c^{N}_{\Xi} &= -9 (D^2+10 D F - 3 F^2) +9 \ , \nonumber \\*
c^{\Sigma}_{\Phi \pm} &= 18 (D^2 - 3 F^2) +10 \pm 12 \ , &
c^{\Sigma}_{\Xi} &= - c^{\Lambda}_{\Xi}= 18 (D^2 - 3 F^2) -18 \ , \nonumber \\*
c^{\Lambda}_{\Phi \pm} &= -18 (D^2 - 3 F^2) -26 \pm 12 \ .
\end{align}
Note that we give no values for $c^{\Xi}_{\Phi \pm}$ and $c^{\Xi}_{\Xi}$, since the renormalization of the corresponding amplitudes is already fixed via eq.~\eqref{DeltaXi}. The renormalized amplitudes acquire a dependence on the chiral renormalization scale $\mu$, which exactly cancels the scale dependence of the leading chiral logarithms:%
\begin{align}
\mu \frac{\partial}{\partial \mu} \Delta \Phi^{B,\text{ren.}}_{\pm,i}(\mu) &= \frac{-1}{(4\pi)^2} \frac{{m_b^\star}^2 c^B_{\Phi \pm}}{24 F_\star^2} \Phi^\star_{\pm,i} \ , &
\mu \frac{\partial}{\partial \mu} \Delta \Xi^{B,\text{ren.}}_{\pm,i}(\mu) &= \frac{-1}{(4\pi)^2} \frac{{m_b^\star}^2 c^B_{\Xi}}{24 F_\star^2} \Xi^\star_{\pm,i} \ .
\end{align}
The replacements given in eq.~\eqref{counterterms} also have to cancel the divergencies in the distribution amplitudes for the $\Xi$ baryon and the $\Pi^B_i$ and $\Upsilon^B_i$ distribution amplitudes, which is the case and can be seen as a nontrivial check of our calculation. The higher order divergencies, which are contained in our result as a consequence of using $\text{IR}$-regularization~\cite{Becher:1999he}, have to be set to zero by hand. This introduces an unphysical scale dependence in higher order terms, which is usually solved by fixing the scale at a typical hadronic value like $\unit{1}{\giga\electronvolt}$. A variation of this scale within reasonable bounds, say between $\unit{0.8}{\giga\electronvolt}$ and $\unit{1.2}{\giga\electronvolt}$, can be used to estimate higher order effects. \par
If we neglect higher order contributions, we can rewrite eqs.~\eqref{DAs_chpt_Result} in such a way that the complete nonanalytic behaviour is encoded in an overall prefactor:%
\begin{subequations} \label{DAs_chpt_Result_factorized}
\begin{align}
\begin{split}
\Phi_{\pm, i}^B &\mathrel{\dot =} \sqrt{\frac{Z_B}{Z^\star}}\Bigl( 1 + \Delta g^B_{\Phi \pm} \Bigr) \Bigl( \Phi_{\pm, i}^\star + \delta m \ \Delta \Phi_{\pm,i}^B \Bigr) \ ,
\end{split} \\
\begin{split}
\Xi_{\pm, i}^B &\mathrel{\dot =} \sqrt{\frac{Z_B}{Z^\star}} \Bigl( 1 + \Delta g^B_{\Xi} \Bigr) \Bigl(\Xi_{\pm, i}^\star + \delta m \ \Delta \Xi_{\pm,i}^B \Bigr) \ ,
\end{split} \\
\begin{split}
\Pi_{i}^B &\mathrel{\dot =} \sqrt{\frac{Z_B}{Z^\star}} \Bigl( 1 + \Delta g^B_{\Pi} \Bigr) \Bigl( \Phi_{\pm_B, i}^\star + \delta m \ \Delta \Pi_{i}^B \Bigr)\ ,
\end{split} \\
\begin{split}
\Upsilon_{i}^B &\mathrel{\dot =} \sqrt{\frac{Z_B}{Z^\star}} \Bigl( 1 + \Delta g^B_{\Xi} \Bigr) \Bigl( \Xi_{\pm_B, i}^\star + \delta m \ \Delta \Upsilon_{i}^B \Bigr) \ ,
\end{split}
\end{align}
\end{subequations}
where%
\begin{align}
\sqrt{\frac{Z_B}{Z^\star}} &\mathrel{\dot =} 1+\frac{1}{2}\Delta \Sigma^\prime_B \ .
\end{align}
From eq.~\eqref{DAs_chpt_Result_factorized} it follows directly that at leading one-loop order the complete nonanalytic structure is contained in the normalization of the distribution amplitudes, while their shape only exhibits the simple dependence on $\delta m$ shown in eq.~\eqref{shape_of_DAs}. Therefore leading finite volume effects do only affect the normalization. We want to emphasize that this is only true by virtue of our specific choice of DAs. A similar behaviour was found for the meson sector (see refs.~\cite{Chen:2003fp,Chen:2005js}). The zeroth moments of the given DAs are not independent, due to eq.~\eqref{zeroth_moments}. In particular all DAs which correspond to operators of certain symmetry classes are normalized by the same wave function normalization constants independent of the twist of the corresponding amplitude. The zeroth moments define the following normalization constants:%
\begin{subequations} \label{normalization_constants}
\begin{align}
f^B &=\int [dx]\Phi^B_{+,i}(x_1,x_2,x_3) = \sqrt{\frac{Z_B}{Z^\star}}\Bigl( 1 + \Delta g^B_{\Phi +} \Bigr) \Bigl(f^\star + \delta m \ \Delta f^B \Bigr) \ , \\
\lambda_1^B &= \int [dx]\Phi^B_{-,4/5}(x_1,x_2,x_3) = \sqrt{\frac{Z_B}{Z^\star}}\Bigl( 1 + \Delta g^B_{\Phi -} \Bigr) \Bigl(\lambda_1^\star + \delta m \ \Delta \lambda_1^B \Bigr) \ , \\
\lambda_2^B &= \int [dx]\Xi^B_{+,4/5}(x_1,x_2,x_3) = \sqrt{\frac{Z_B}{Z^\star}}\Bigl( 1 + \Delta g^B_{\Xi} \Bigr) \Bigl(\lambda_2^\star + \delta m \ \Delta \lambda_2^B \Bigr) \ ,
\shortintertext{and}
f_T^\Sigma&=\int [dx]\Pi^{\Sigma}_{i}(x_1,x_2,x_3) = \sqrt{\frac{Z_\Sigma}{Z^\star}}\Bigl( 1 + \Delta g^\Sigma_{\Pi} \Bigr) \Bigl(f^\star + \delta m \ \Delta f_T^\Sigma \Bigr) \ , \\
f_T^\Xi&=\int [dx]\Pi^{\Xi}_{i}(x_1,x_2,x_3) = \sqrt{\frac{Z_\Xi}{Z^\star}}\Bigl( 1 + \Delta g^\Xi_{\Pi} \Bigr) \Bigl(f^\star + \delta m \ \Delta f_T^\Xi \Bigr) \ , \\
\lambda_T^\Lambda&=\int [dx]\Pi^{\Lambda}_{4/5}(x_1,x_2,x_3) = \sqrt{\frac{Z_\Lambda}{Z^\star}}\Bigl( 1 + \Delta g^\Lambda_{\Pi} \Bigr) \Bigl(\lambda_1^\star + \delta m \ \Delta \lambda_T^\Lambda \Bigr) \ .
\intertext{For the remaining zeroth moments one finds}
f^{N} &=\int [dx]\Pi^{N}_{i}(x_1,x_2,x_3) \ , \\
0&=\int [dx]\Phi^B_{-,3/6}(x_1,x_2,x_3) = \int [dx]\Xi^B_{-,4/5}(x_1,x_2,x_3) = \int [dx]\Pi^{\Lambda}_{3/6}(x_1,x_2,x_3) \ , \\
& \mathrel{\phantom{=}} \int [dx]\Upsilon^{B}_{4/5}(x_1,x_2,x_3) = \begin{cases} \lambda_2^B \ , &\text{for } B \neq \Lambda \\ 0 \ , &\text{for } B = \Lambda \end{cases} \ .
\end{align}
\end{subequations}
Due to eq.~\eqref{constraints_on_Delta_DAs},%
\begin{align} \label{Delta_fT_LambdaT}
\Delta f_T^\Sigma &= -\frac{3}{2}\Delta f^\Lambda - \frac{1}{2} \Delta f^\Sigma \ , &
\Delta f_T^\Xi &= \frac{3}{2}\Delta f^\Lambda +\frac{1}{2}\Delta f^\Sigma-\Delta f^{N} \ , \nonumber \\
\Delta \lambda_T^\Lambda &= -\frac{1}{2}\Delta \lambda_1^\Lambda -\frac{3}{2} \Delta \lambda_1^\Sigma \ .
\end{align}
In the equations above we have introduced convenient new definitions of $f^\Lambda$, $\lambda_1^\Lambda$, $\lambda_2^\Lambda$, $f_T^\Sigma$, $f_T^\Xi$ and $\lambda_T^\Lambda$ such that, in the limit of exact $SU(3)_f$ symmetry,%
\begin{subequations}
\begin{align}
f^\star &= f^N = f^\Sigma = f^\Xi = f^\Lambda = f^\Sigma_T = f^\Xi_T \ , \\*
\lambda_1^\star &= \lambda_1^N = \lambda_1^\Sigma = \lambda_1^\Xi = \lambda_1^\Lambda = \lambda_T^\Lambda , \\*
\lambda_2^\star &= \lambda_2^N = \lambda_2^\Sigma = \lambda_2^\Xi = \lambda_2^\Lambda \ .
\end{align}
\end{subequations}
If the reader favors a different definition he or she can easily read off the conversion factor from eq.~\eqref{definition_superior_DAs}, noting that additional signs can arise from eq.~\eqref{baryon_signs} if one uses different baryons for the definition of the distribution amplitudes, and that one has to take into account additional factors originating from differing definitions of $S_i$, $P_i$, $V_i$, $A_i$ and $T_i$ (we use the definitions of ref.~\cite{Braun:2000kw}). We have performed this matching procedure for the constants defined in refs.~\cite{Chernyak:1987nu,Braun:2000kw} (see appendix~\ref{app_matching}). Note that the constants $f_T^\Sigma$, $f_T^\Xi$ and $\lambda_T^\Lambda$ given above are (at leading one-loop accuracy) completely fixed by $f^B$ and $\lambda_1^B$. However, without the knowledge of the $\operatorname{SU}(3)_f$ breaking effects one would have to define them as additional free normalization constants. $f^\star$, $\Delta f^B$, $\lambda_i^\star$ and $\Delta \lambda_i^B$ are given by%
\begin{subequations}
\begin{align}
f^\star &= \int [dx]\Phi^\star_{+,i}(x_1,x_2,x_3) \ , & \Delta f^B &= \int [dx]\Delta \Phi^B_{+,i}(x_1,x_2,x_3) \ , \\
\lambda_1^\star &= \int [dx]\Phi^\star_{-,4/5}(x_1,x_2,x_3) \ , & \Delta \lambda_1^B &= \int [dx]\Delta \Phi^B_{-,4/5}(x_1,x_2,x_3) \ , \\
\lambda_2^\star &= \int [dx]\Xi^\star_{+,4/5}(x_1,x_2,x_3) \ , & \Delta \lambda_2^B &= \int [dx]\Delta \Xi^B_{+,4/5}(x_1,x_2,x_3) \ ,
\end{align}
\end{subequations}
where, as a consequence of eq.~\eqref{DeltaXi} (first line) and eq.~\eqref{zeroth_moments} (second line) one has%
\begin{align}
\Delta f^\Xi &= - \Delta f^\Sigma - \Delta f^N \ , &
\Delta \lambda_1^\Xi &= - \Delta \lambda_1^\Sigma - \Delta \lambda_1^N \ , &
\Delta \lambda_2^\Xi &= - \Delta \lambda_2^\Sigma - \Delta \lambda_2^N \ , \nonumber \\*
\Delta \lambda_2^\Lambda &= - \Delta \lambda_2^\Sigma \ . \label{Delta_f_Lambda}
\end{align}
The zeroth moments of $\Phi^B_{-,3/6}$ and $\Pi^\Lambda_{3/6}$ ($\Xi^B_{-,4/5}$ and $\Upsilon^\Lambda_{4/5}$) vanish by construction, since they are antisymmetric under exchange of $x_1$ and $x_3$ ($x_2$ and $x_3$). One possible approach would be to normalize these amplitudes by their first moments. However, our main goal is to divide the DAs by normalization constants in such a way that the nonanalytic prefactor is canceled. This can be achieved without the definition of additional constants, since all prefactors present in eqs.~\eqref{DAs_chpt_Result_factorized} also occur in eqs.~\eqref{normalization_constants}. Explicitly, one can consider the ratios%
\begin{subequations} \label{shape_of_DAs}
\begin{align}
\frac{\Phi_{+, i}^B}{f^B}&= \frac{\Phi_{+, i}^\star + \delta m \ \Delta \Phi_{+,i}^B}{f^\star + \delta m \ \Delta f^B} \ , &
\frac{\Phi_{-, i}^B}{\lambda_1^B}&= \frac{\Phi_{-, i}^\star + \delta m \ \Delta \Phi_{-,i}^B}{\lambda_1^\star + \delta m \ \Delta \lambda_1^B} \ , \\*
\frac{\Pi_i^{N}}{f^{N}}&= \frac{\Phi_{+, i}^\star + \delta m \ \Delta \Phi_{+,i}^{N}}{f^\star + \delta m \ \Delta f^{N}} \ ,
& \frac{\Pi_i^\Lambda}{\lambda^\Lambda_T}&= \frac{\Phi_{-, i}^\star + \delta m \ \Delta \Pi_i^\Lambda}{\lambda_1^\star + \delta m \ \Delta \lambda_T^\Lambda} \ , \\*
\frac{\Pi_i^{\Sigma/\Xi}}{f^{\Sigma/\Xi}_T}&= \frac{\Phi_{+, i}^\star + \delta m \ \Delta \Pi_i^{\Sigma/\Xi}}{f^\star + \delta m \ \Delta f_T^{\Sigma/\Xi}} \ , \\
\frac{\Xi_{\pm, i}^B}{\lambda_2^B}&= \frac{\Xi_{\pm, i}^\star + \delta m \ \Delta \Xi_{\pm, i}^B}{\lambda_2^\star + \delta m \ \Delta \lambda_2^B} \ , & \frac{\Upsilon_i^B}{\lambda_2^B}&= \frac{\Xi_{\pm_B, i}^\star + \delta m \ \Delta \Upsilon_i^B}{\lambda_2^\star + \delta m \ \Delta \lambda_2^B} \ .
\end{align}
\end{subequations}
The idea behind the latter choice is to normalize all DAs with similar behaviour under chiral extrapolation (including the ones with vanishing zeroth moment) with the same normalization constant containing the complete nonanalytic behaviour. In this way one obtains a one-to-one correspondence between a normalization constant and a certain chiral behaviour. Note that, following this argument, some of the moments of the leading twist DA $\Phi_3^B = \Phi_{+,3}^B + \Phi_{-,3}^B$ should be normalized with $\lambda_1^B$ instead of $f^B$. Otherwise the corresponding shape parameters do contain chiral logarithms.%
\subsection{Example of application} \label{sect_example_of_application}
In this section we will work out explicit expressions for the DAs defined in eqs.~\eqref{definition_superior_DAs} and~\eqref{definition_superior_DAs_second_part} in terms of the shape parameters given in refs.~\cite{Anikin:2013aka,Anikin:2013yoa,Braun:2008ia}, where contributions of Wandzura-Wilczek type~\cite{Wandzura:1977qf} are taken into account explicitly. For brevity we apply the approximation advocated in ref.~\cite{Anikin:2013aka}, where contributions that can mix with four-particle operators are systematically neglected. We use the definitions of said references and we define additionally%
\begin{align} \label{parity_polynomials}
\mathcal P_{nk}(x_1,x_2,x_3)&= p_{nk} \mathcal P_{nk}(x_3,x_2,x_1) \ ,
\end{align}
where $p_{nk}=\pm1$, depending on $n$ and $k$. This definition is possible since the polynomials $\mathcal P_{nk}$ have definite parity under exchange of $x_1$ and $x_3$ \cite{Anikin:2013aka}. We will call the polynomials with $p_{nk}=+1$ ($p_{nk}=-1$) even (odd). For the DAs we find%
\begin{subequations} \label{new_def}
\begin{align}
\Phi^B_{+,3} &= f^B \Phi_{+,3}^{B,t=3}\ , \\*
\Phi^B_{-,3}&= \lambda_1^B \Phi_{-,3}^{B,t=3} \ , \\*
\Phi^B_{+,4} &= f^B \Bigl( \Phi_{+,4}^{B,WW_3} + \Phi_{+,4}^{B,t=4} \Bigr) \ , &
\Xi^B_{\pm,4} &= \lambda_2^B \Xi_{\pm,4}^{B,t=4} \ , \\*
\Phi^B_{-,4} &= \lambda_1^B \Bigl(\Phi_{-,4}^{B,WW_3} + \Phi_{-,4}^{B,t=4} \Bigr) \ , \\
\Phi^B_{+,5} &= f^B \Bigl( \Phi_{+,5}^{B,WW_3} + \Phi_{+,5}^{B,WW_4} + \Phi_{+,5}^{B,t=5} \Bigr) \ , &
\Xi^B_{\pm,5} &= \lambda_2^B \Bigl( \Xi_{\pm,5}^{B,WW_4} + \Xi_{\pm,5}^{B,t=5} \Bigr) \ , \\*
\Phi^B_{-,5} &= \lambda_1^B \Bigl( \Phi_{-,5}^{B,WW_3} + \Phi_{-,5}^{B,WW_4} + \Phi_{-,5}^{B,t=5} \Bigr) \ ,
\end{align}
\end{subequations}
where all chiral logarithms are contained in the prefactors. Analogous expressions for the $\Pi$ and $\Upsilon$ DAs will be given below in eq.~\eqref{new_def_2}. Genuine twist $5$ contributions ($\Phi_{\pm,5}^{B,t=5}$, $\Xi_{\pm,5}^{B,t=5}$) will be neglected in this approximation. Also twist $6$ DAs are neglected; one could in principle take into account Wandzura-Wilczek contributions to the twist $6$ DAs, but the corresponding expressions are not known yet. The shape of the DAs is given by the genuine twist $3$ and twist $4$ contributions%
\begin{subequations}
\begin{align}
\Phi_{+,3}^{B,t=3}(x_1,x_2,x_3) &= 120 x_1 x_2 x_3 \sum_{\mathclap{\stackrel{n,k\leq n}{p_{nk}=+1}}} \varphi^B_{nk} \mathcal P_{nk} (x_1,x_2,x_3) \ , \\*
\Phi_{-,3}^{B,t=3}(x_1,x_2,x_3) &= 120 x_1 x_2 x_3 \sum_{\mathclap{\stackrel{n,k\leq n}{p_{nk}=-1}}} \tilde \varphi^B_{nk} \mathcal P_{nk} (x_1,x_2,x_3) \ , \\
\begin{split}
\Phi_{+,4}^{B,t=4}(x_1,x_2,x_3) &= 24 x_1 x_2 \biggl( \frac{10}{3} (2 x_1 - x_2 - 2 x_3) \tilde \eta^B_{11} + \dots \biggr)\ ,
\end{split} \\
\begin{split}
\Phi_{-,4}^{B,t=4}(x_1,x_2,x_3) &= 24 x_1 x_2 \bigl( \eta^B_{00} + 2 (2 - 5 x_2) \eta^B_{10} + \dots \bigr) \ ,
\end{split} \\
\begin{split}
\Xi_{+,4}^{B,t=4}(x_1,x_2,x_3) &= 24 x_2 x_3 \biggl(\xi^B_{00} - \frac{9}{4} (1 - 5 x_1) \xi^B_{10} + \dots \biggr) \ ,
\end{split} \\
\begin{split}
\Xi_{-,4}^{B,t=4}(x_1,x_2,x_3) &= 24 x_2 x_3 \biggl(-\frac{45}{4}(x_2-x_3)\xi^B_{10} + \dots \biggr) \ ,
\end{split}
\end{align}
\end{subequations}
and the Wandzura-Wilczek contributions (see refs.~\cite{Anikin:2013aka,Anikin:2013yoa,Braun:2008ia})%
\begin{subequations} \label{WandzuraWilczek}
\begin{align}
\Phi_{+,4}^{B,WW_3}(x_1,x_2,x_3)&= -\sum_{\mathclap{\stackrel{n,k\leq n}{p_{nk}=+1}}} \frac{240\varphi^B_{nk}}{(n+2)(n+3)} \biggl( n +2 -\frac{\partial}{\partial x_3} \biggr) x_1 x_2 x_3 \mathcal P_{nk} (x_1,x_2,x_3)\ , \\*
\Phi_{-,4}^{B,WW_3}(x_1,x_2,x_3) &= - \sum_{\mathclap{\stackrel{n,k\leq n}{p_{nk}=-1}}} \frac{240 \tilde \varphi^B_{nk}}{(n+2)(n+3)} \biggl( n +2 -\frac{\partial}{\partial x_3} \biggr) x_1 x_2 x_3 \mathcal P_{nk} (x_1,x_2,x_3) \ , \\
\begin{split}
\Phi_{+,5}^{B,WW_3}(x_1,x_2,x_3)&= \sum_{\mathclap{\stackrel{n,k\leq n}{p_{nk}=+1}}} \begin{aligned}[t] & \frac{240\varphi^B_{nk}}{(n+2)(n+3)} \biggl[\biggl( n +2 -\frac{\partial}{\partial x_1} \biggr)\biggl( n +1 -\frac{\partial}{\partial x_2}\biggr) - (n+2)^2 \biggr]\\ &\quad \times x_1 x_2 x_3 \mathcal P_{nk} (x_1,x_2,x_3) \ , \end{aligned}
\end{split} \\
\begin{split}
\Phi_{-,5}^{B,WW_3}(x_1,x_2,x_3)&= \sum_{\mathclap{\stackrel{n,k\leq n}{p_{nk}=-1}}} \begin{aligned}[t] & \frac{240\tilde\varphi^B_{nk}}{(n+2)(n+3)} \biggl[\biggl( n +2 -\frac{\partial}{\partial x_1} \biggr)\biggl( n +1 -\frac{\partial}{\partial x_2}\biggr) - (n+2)^2 \biggr]\\ &\quad \times x_1 x_2 x_3 \mathcal P_{nk} (x_1,x_2,x_3) \ , \end{aligned}
\end{split} \\
\begin{split}
\Phi_{+,5}^{B,WW_4}(x_1,x_2,x_3) &= 4 x_3 \bigl( 5 (x_1^2 + 2 x_2 x_3 - x_3^2) \tilde \eta^B_{11} + \dots \bigr)\ ,
\end{split} \\
\begin{split}
\Phi_{-,5}^{B,WW_4}(x_1,x_2,x_3) &= 4 x_3 (1 - x_2) \bigl( 2 \eta^B_{00} + 3 (1 - 5 x_2) \eta^B_{10} + \dots \bigr) \ ,
\end{split} \\
\begin{split}
\Xi_{+,5}^{B,WW_4}(x_1,x_2,x_3)
&= 4 x_1 (1 + x_1) \xi^B_{00} - \frac{27}{2} (4 - 4 x_1 + x_1^2- 5 x_1^3) \xi^B_{10} + \dots \ ,
\end{split} \\
\begin{split}
\Xi_{-,5}^{B,WW_4}(x_1,x_2,x_3)
&= - 12 x_1 (x_2 - x_3) \xi^B_{00} + \frac{27}{2}(5-x_1 + 5 x_1^2) (x_2 - x_3) \xi^B_{10} + \dots \ ,
\end{split}
\end{align}
\end{subequations}
where the summation over $n$ starts from $0$ and, generally, goes to infinity, but is truncated at $n=2$ in the approximation of ref.~\cite{Anikin:2013aka}.\footnote{We do not take into account possible quark mass corrections to eq.~\eqref{WandzuraWilczek} and eq.~\eqref{WandzuraWilczek2} below (compare e.g.\ refs.~\cite{Ball:1998sk,Braun:2004vf} where such computations have been performed for vector-meson and pseudoscalar-meson DAs), since they can (by definition) be absorbed into the genuine higher twist terms in eq.~\eqref{new_def} and~\eqref{new_def_2}. Let us note in passing that our general result does not rely on the separation of Wandzura-Wilczek and genuine higher twist terms at all, since the calculation within chiral perturbation theory does not distinguish between these contributions.} Note that our separation into ``$+$'' and ``$-$'' amplitudes at leading twist level corresponds to a separation of even and odd polynomials. The normalization constants are still defined such that $\eta^B_{00} = \varphi^B_{00}=\xi^B_{00}=1$, which are only kept for a cleaner notation. Note also that the introduction of $\tilde\varphi^B_{nk}$ and $\tilde\eta^B_{nk}$ only amounts to a redefinition of the shape parameters occurring in $\Phi^{B,t=3}_{-,3}$, $\Phi_{-,4}^{B,WW_3}$ and $\Phi_{-,5}^{B,WW_3}$ by a factor of $f^B/\lambda_1^B$ and the ones occurring in $ \Phi_{+,4}^{B,t=4}$ and $ \Phi_{+,5}^{B,WW_4}$ by a factor of $\lambda_1^B/f^B$ with respect to ref.~\cite{Anikin:2013aka} (the corresponding anomalous dimensions have to be adjusted accordingly):%
\begin{align}
\tilde \varphi^B_{nk} &= \frac{f^B}{\lambda_1^B} \varphi^B_{nk} \ , & \tilde \eta^B_{nk} &= \frac{\lambda_1^B}{f^B} \eta^B_{nk} \ .
\end{align}
The dependence of the shape parameters on the quark mass splitting takes the following form%
\begin{subequations} \label{shape_parameters_1}
\begin{align}
\varphi^B_{nk} &= \frac{\varphi^\star_{nk} + \delta m \ \Delta\varphi^B_{nk}}{f^\star + \delta m \ \Delta f^B} \ , \quad \text{if } p_{nk}=+1 \ , &
\tilde\varphi^B_{nk} &= \frac{\tilde\varphi^\star_{nk} + \delta m \ \Delta\tilde\varphi^B_{nk}}{\lambda_1^\star + \delta m \ \Delta \lambda_1^B} \ , \quad \text{if } p_{nk}=-1 \ , \\
\tilde\eta^B_{11} &= \frac{ \tilde\eta^\star_{11} + \delta m \ \Delta \tilde\eta^B_{11}}{f^\star + \delta m \ \Delta f^B} \ , & \eta^B_{10} &= \frac{ \eta^\star_{10} + \delta m \ \Delta \eta^B_{10}}{\lambda_1^\star + \delta m \ \Delta \lambda_1^B} \ , \\*
\xi^B_{10} &= \frac{ \xi^\star_{10} + \delta m \ \Delta \xi^B_{10}}{\lambda_2^\star + \delta m \ \Delta \lambda_2^B} \ ,
\end{align}
\end{subequations}
which corresponds directly to eq.~\eqref{shape_of_DAs}, while the dependence of the normalization constants is given in eq.~\eqref{normalization_constants}. The parameters describing $\operatorname{SU}(3)_f$ symmetry breaking are restricted by eq.~\eqref{DeltaXi} such that%
\begin{align} \label{su3cons_params_1}
\Delta x^\Xi_{nk} &= - \Delta x^{N}_{nk} - \Delta x^\Sigma_{nk} \ , \quad \text{for } x \in \{ \varphi,\tilde\varphi,\eta,\tilde\eta,\xi \} \ .
\end{align}
For the original twist $3$ and $4$ DAs given in ref.~\cite{Braun:2000kw} (see also eq.~\eqref{original_DAs_definition}) the new choice of normalization yields%
\begin{subequations}
\begin{align}
\begin{split}
\Phi^N_3(x_1,x_2,x_3) &= \Bigl(\Phi^N_{+,3} + \Phi^N_{-,3}\Bigr)(x_1,x_2,x_3) \\*
&= f^N \Phi^{N,t=3}_{+,3}(x_1,x_2,x_3) + \lambda_1^N \Phi^{N,t=3}_{-,3} (x_1,x_2,x_3) \ ,
\end{split} \\
\begin{split}
\Phi^N_4(x_1,x_2,x_3) &= \frac{1}{2} \Bigl( \Phi^N_{+,4} + \Phi^N_{-,4} \Bigr)(x_1,x_2,x_3) \\
&= \frac{f^N}{2} \Bigl( \Phi_{+,4}^{N,WW_3} \! + \Phi_{+,4}^{N,t=4} \Bigr)(x_1,x_2,x_3) + \frac{\lambda_1^N }{2} \Bigl( \Phi_{-,4}^{N,t=4} + \Phi_{-,4}^{N,WW_3} \! \Bigr)(x_1,x_2,x_3) \ , \!
\end{split} \\
\begin{split}
\Psi^N_4(x_1,x_2,x_3) &= \frac{1}{2} \Bigl( \Phi^N_{+,4} - \Phi^N_{-,4} \Bigr)(x_3,x_1,x_2)\\
&= \frac{f^N}{2} \Bigl( \Phi_{+,4}^{N,WW_3} \! + \Phi_{+,4}^{N,t=4} \Bigr)(x_3,x_1,x_2) - \frac{\lambda_1^N }{2} \Bigl( \Phi_{-,4}^{N,t=4} + \Phi_{-,4}^{N,WW_3} \! \Bigr)(x_3,x_1,x_2) \ , \!
\end{split} \\
\begin{split}
\Xi^N_4(x_1,x_2,x_3) &= \frac{1}{6} \Bigl( \Xi^N_{+,4} + \Xi^N_{-,4} \Bigr)(x_1,x_2,x_3)\\
&= \frac{\lambda_2^N}{6} \Bigl(\Xi_{+,4}^{N,t=4} + \Xi_{-,4}^{N,t=4} \Bigr)(x_1,x_2,x_3) \ ,
\end{split}
\end{align}
\end{subequations}
where, as discussed above, the normalization of the odd moments of the leading twist amplitude with $\lambda_1^N$ (instead of $f^N$) appropriately reflects their chiral behaviour. Note that this is consistent with an earlier two-flavor BChPT calculation, where it was found that the odd first and second moments of the leading twist amplitude have the same chiral logarithms as $\lambda_{1}^N$ (see appendix of ref.~\cite{PhysRevD.89.094511}).\par
For a description of the complete baryon octet one also needs the $\Pi$ and $\Upsilon$ DAs defined in eq.~\eqref{definition_superior_DAs_second_part}, which are relevant for the hyperons. These are completely fixed by the $\Phi_\pm$ and $\Xi$ DAs. Consequently, the following equations do not contain any additional parameters:%
\begin{subequations} \label{new_def_2}
\begin{align}
\Pi^{N}_{i} &= \Phi_{+,i}^{N}\ , &
\Upsilon^{N}_{i} &= \Xi_{+,i}^{N}\ , \\*
\Pi^{\Sigma/\Xi}_{3} &= f^{\Sigma/\Xi}_T \Pi_{3}^{\Sigma/\Xi,t=3}\ , \\
\Pi^\Lambda_{3}&= \lambda_T^\Lambda \Pi_{3}^{\Lambda,t=3} \ , \\
\Pi_{4}^{\Sigma/\Xi} &= f^{\Sigma/\Xi}_T \Bigl( \Pi_{4}^{\Sigma/\Xi,WW_3} + \Pi_{4}^{\Sigma/\Xi,t=4} \Bigr) \ , &
\Upsilon^B_{4} &= \lambda_2^B \Upsilon_{4}^{B,t=4} \ , \\
\Pi_{4}^{\Lambda} &= \lambda_T^\Lambda \Bigl( \Pi_{4}^{\Lambda,WW_3} + \Pi_{4}^{\Lambda,t=4} \Bigr)\ , \\
\Pi^{\Sigma/\Xi}_{5} &= f^{\Sigma/\Xi}_T \Bigl( \Pi_{5}^{\Sigma/\Xi,WW_3} + \Pi_{5}^{\Sigma/\Xi,WW_4} + \Pi_{5}^{\Sigma/\Xi,t=5} \Bigr) \ , &
\Upsilon^B_{5} &= \lambda_2^B \Bigl( \Upsilon_{5}^{B,WW_4} + \Upsilon_{5}^{B,t=5} \Bigr) \ , \\*
\Pi^\Lambda_{5} &= \lambda_T^\Lambda \Bigl( \Pi_{5}^{\Lambda,WW_3} + \Pi_{5}^{\Lambda,WW_4} + \Pi_{5}^{\Lambda,t=5} \Bigr) \ ,
\end{align}
\end{subequations}
where the genuine twist $5$ contributions $\Pi_{5}^{B,t=5}$ and $\Upsilon_{5}^{B,t=5}$ will be neglected as above. The genuine twist $3$ and twist $4$ contributions are%
\begin{subequations}
\begin{align}
\Pi_{3}^{\Sigma/\Xi,t=3}(x_1,x_2,x_3) &= 120 x_1 x_2 x_3 \sum_{\mathclap{\stackrel{n,k\leq n}{p_{nk}=+1}}} \pi^{\Sigma/\Xi}_{nk} \mathcal P_{nk} (x_1,x_2,x_3) \ , \\*
\Pi_{3}^{\Lambda,t=3}(x_1,x_2,x_3) &= 120 x_1 x_2 x_3 \sum_{\mathclap{\stackrel{n,k\leq n}{p_{nk}=-1}}} \tilde \pi^{\Lambda}_{nk} \mathcal P_{nk} (x_1,x_2,x_3) \ , \\
\begin{split}
\Pi_{4}^{\Sigma/\Xi,t=4}(x_1,x_2,x_3)
&= 24 x_1 x_2 \biggl( \frac{10}{3} (2 x_1 - x_2 - 2 x_3) \tilde \zeta^{\Sigma/\Xi}_{11} + \dots \biggr)\ ,
\end{split} \\
\begin{split}
\Pi_{4}^{\Lambda,t=4}(x_1,x_2,x_3)
&= 24 x_1 x_2 \bigl( \zeta^\Lambda_{00} + 2 (2 - 5 x_2) \zeta^\Lambda_{10} + \dots \bigr) \ ,
\end{split} \\
\begin{split}
\Upsilon_{4}^{\Sigma/\Xi,t=4}(x_1,x_2,x_3)
&= 24 x_2 x_3 \biggl(\upsilon^{\Sigma/\Xi}_{00} - \frac{9}{4} (1 - 5 x_1) \upsilon^{\Sigma/\Xi}_{10} + \dots \biggr) \ ,
\end{split} \\*
\begin{split}
\Upsilon_{4}^{\Lambda,t=4}(x_1,x_2,x_3)
&= 24 x_2 x_3 \biggl(-\frac{45}{4}(x_2-x_3)\upsilon^\Lambda_{10} + \dots \biggr) \ .
\end{split}
\end{align}
\end{subequations}
The shape parameters are fixed:%
\begin{subequations} \label{shape_parameters_2}
\begin{align}
\pi^{\Sigma/\Xi}_{nk} &= \frac{\varphi^\star_{nk} + \delta m \ \Delta\pi^{\Sigma/\Xi}_{nk}}{f^\star + \delta m \ \Delta f^{\Sigma/\Xi}_T} \ , &
\tilde\pi^\Lambda_{nk} &= \frac{\tilde\varphi^\star_{nk} + \delta m \ \Delta\tilde\pi^\Lambda_{nk}}{\lambda_1^\star + \delta m \ \Delta \lambda_T^\Lambda} \ , \\
\tilde \zeta^{\Sigma/\Xi}_{11} &= \frac{ \tilde \eta^\star_{11} + \delta m \ \Delta \tilde \zeta^{\Sigma/\Xi}_{11}}{f^\star + \delta m \ \Delta f^{\Sigma/\Xi}_T} \ , & \zeta^{\Lambda}_{10} &= \frac{ \eta^\star_{10} + \delta m \ \Delta \zeta^{\Lambda}_{10}}{\lambda_1^\star + \delta m \ \Delta \lambda_T^\Lambda} \ , \\
\upsilon^B_{10} &= \frac{ \xi^\star_{10} + \delta m \ \Delta \upsilon^B_{10}}{\lambda_2^\star + \delta m \ \Delta \lambda_2^B} \ ,
\end{align}
\end{subequations}
where $ \Delta f^{\Sigma/\Xi}_T$ and $\Delta \lambda_T^\Lambda$ are defined in eq.~\eqref{Delta_fT_LambdaT}. The parameters describing $\operatorname{SU}(3)_f$ symmetry breaking can be determined by eqs.~\eqref{DeltaXi} and~\eqref{constraints_on_Delta_DAs}:%
\begin{subequations} \label{su3cons_params_2}
\begin{align}
\Delta \pi_{nk}^\Sigma &= -\frac{1}{2} \Delta \varphi_{nk}^\Sigma -\frac{3}{2} \Delta \varphi_{nk}^\Lambda \ , &
\Delta \tilde\pi_{nk}^\Lambda &= -\frac{1}{2} \Delta \tilde\varphi_{nk}^\Lambda -\frac{3}{2} \Delta \tilde\varphi_{nk}^\Sigma \ , \nonumber \\
\Delta \pi_{nk}^\Xi &= \frac{3}{2} \Delta \varphi_{nk}^\Lambda + \frac{1}{2} \Delta \varphi_{nk}^\Sigma - \Delta \varphi_{nk}^{N} \ , \\
\Delta \tilde\zeta_{11}^\Sigma &= -\frac{1}{2} \Delta \tilde \eta_{11}^\Sigma -\frac{3}{2} \Delta \tilde\eta_{11}^\Lambda \ , &
\Delta \zeta_{10}^\Lambda &= -\frac{1}{2} \Delta \eta_{10}^\Lambda -\frac{3}{2} \Delta \eta_{10}^\Sigma \ , \nonumber \\
\Delta \tilde\zeta_{11}^\Xi &= \frac{3}{2} \Delta \tilde\eta_{11}^\Lambda + \frac{1}{2} \Delta \tilde\eta_{11}^\Sigma - \Delta \tilde\eta_{11}^{N} \ , \\
\Delta \upsilon_{10}^\Sigma &= -\frac{1}{2} \Delta \xi_{10}^\Sigma -\frac{3}{2} \Delta \xi_{10}^\Lambda \ ,&
\Delta \upsilon_{10}^\Lambda &= -\frac{1}{2} \Delta \xi_{10}^\Lambda -\frac{3}{2} \Delta \xi_{10}^\Sigma \ , \nonumber \\
\Delta \upsilon_{10}^\Xi &= \frac{3}{2} \Delta \xi_{10}^\Lambda + \frac{1}{2} \Delta \xi_{10}^\Sigma - \Delta \xi_{10}^{N} \ .
\end{align}
\end{subequations}
The Wandzura-Wilczek contributions take the form%
\begin{subequations}\label{WandzuraWilczek2}
\begin{align}
\Pi_{4}^{\Sigma/\Xi,WW_3}(x_1,x_2,x_3)&= -\sum_{\mathclap{\stackrel{n,k\leq n}{p_{nk}=+1}}} \frac{240\pi^{\Sigma/\Xi}_{nk}}{(n+2)(n+3)} \biggl( n +2 -\frac{\partial}{\partial x_3} \biggr) x_1 x_2 x_3 \mathcal P_{nk} (x_1,x_2,x_3)\ , \\*
\Pi_{4}^{\Lambda,WW_3}(x_1,x_2,x_3) &= - \sum_{\mathclap{\stackrel{n,k\leq n}{p_{nk}=-1}}} \frac{240 \tilde \pi^\Lambda_{nk}}{(n+2)(n+3)} \biggl( n +2 -\frac{\partial}{\partial x_3} \biggr) x_1 x_2 x_3 \mathcal P_{nk} (x_1,x_2,x_3) \ , \\
\begin{split}
\Pi_{5}^{\Sigma/\Xi,WW_3}(x_1,x_2,x_3)&= \sum_{\mathclap{\stackrel{n,k\leq n}{p_{nk}=+1}}} \begin{aligned}[t] & \frac{240\pi^{\Sigma/\Xi}_{nk}}{(n+2)(n+3)} \biggl[\biggl( n +2 -\frac{\partial}{\partial x_1} \biggr)\biggl( n +1 -\frac{\partial}{\partial x_2}\biggr) - (n+2)^2 \biggr]\\ &\quad \times x_1 x_2 x_3 \mathcal P_{nk} (x_1,x_2,x_3) \ , \end{aligned}
\end{split} \\
\begin{split}
\Pi_{5}^{\Lambda,WW_3}(x_1,x_2,x_3)&= \sum_{\mathclap{\stackrel{n,k\leq n}{p_{nk}=-1}}} \begin{aligned}[t] & \frac{240\tilde\pi^\Lambda_{nk}}{(n+2)(n+3)} \biggl[\biggl( n +2 -\frac{\partial}{\partial x_1} \biggr)\biggl( n +1 -\frac{\partial}{\partial x_2}\biggr) - (n+2)^2 \biggr]\\ &\quad \times x_1 x_2 x_3 \mathcal P_{nk} (x_1,x_2,x_3) \ , \end{aligned}
\end{split} \\
\begin{split}
\Pi_{5}^{\Sigma/\Xi,WW_4}(x_1,x_2,x_3)
&= 4 x_3 \bigl( 5 (x_1^2 + 2 x_2 x_3 - x_3^2) \tilde \zeta^{\Sigma/\Xi}_{11} + \dots \bigr)\ ,
\end{split} \\
\begin{split}
\Pi_{5}^{\Lambda,WW_4}(x_1,x_2,x_3)
&= 4 x_3 (1 - x_2) \bigl( 2 \zeta^\Lambda_{00} + 3 (1 - 5 x_2) \zeta^\Lambda_{10} + \dots \bigr) \ ,
\end{split} \\
\begin{split}
\Upsilon_{5}^{\Sigma/\Xi,WW_4}(x_1,x_2,x_3)
&= 4 x_1 (1 + x_1) \upsilon^{\Sigma/\Xi}_{00} - \frac{27}{2} (4 - 4 x_1 + x_1^2- 5 x_1^3) \upsilon^{\Sigma/\Xi}_{10} + \dots \ ,
\end{split} \\
\begin{split}
\Upsilon_{5}^{\Lambda,WW_4}(x_1,x_2,x_3)
&= - 12 x_1 (x_2 - x_3) \upsilon^\Lambda_{00} + \frac{27}{2} (5-x_1 + 5 x_1^2) (x_2 - x_3) \upsilon^\Lambda_{10} + \dots \ .
\end{split}
\end{align}
\end{subequations}
To conclude this section we want to point out the merits of our calculation. First of all, we found that the behaviour under chiral extrapolation of a certain moment correlates to its parity in the sense of eq.~\eqref{parity_polynomials}. Therefore it is advantageous to normalize the odd moments of the leading twist DA with $\lambda_1^B$ instead of $f^B$. Quantitatively more important, however, is the significant reduction of parameters: we find that (within the approximation used above) we only need $43$ parameters to describe the complete set of baryon octet three-quark DAs (including their dependence on the quark mass splitting). In contrast, an ad~hoc linear extrapolation without the knowledge of $\operatorname{SU}(3)_f$ symmetry breaking would require $72$ parameters for the given setup, since one can not make use of eqs.~\eqref{Delta_fT_LambdaT}, \eqref{Delta_f_Lambda}, \eqref{su3cons_params_1} and~\eqref{su3cons_params_2}.%
\subsection{Dependence on the mean quark mass} \label{sect_mean_quark}
The distribution amplitudes $\Phi^\star_{\pm,i}$ and $\Xi^\star_{\pm,i}$ have a nontrivial dependence on the mean quark mass $\bar m_q$. This is not really interesting from a phenomenological point of view, since the number of independent distribution amplitudes can not be further reduced compared to eq.~\eqref{DAs_chpt_Result_factorized}, even if one expands everything around the chiral limit. However, the dependence is of importance for the analysis of lattice data if one wants to include data points from simulations with unphysical mean quark mass. The mass dependence reads%
\begin{subequations} \label{DAs_chpt_Result_symmetric_line}
\begin{align}
\Phi^\star_{\pm,i} &= \Phi^\circ_{\pm,i} \Bigl(1+\tfrac{1}{2} \Sigma^{\prime\star} + g^\star_{\Phi\pm} \Bigr) + \bar m \Delta \Phi^\star_{\pm,i} \ , \\
\Xi^\star_{\pm,i} &= \Xi^\circ_{\pm,i} \Bigl(1+\tfrac{1}{2} \Sigma^{\prime\star} + g^\star_{\Xi} \Bigr) + \bar m \Delta \Xi^\star_{\pm,i} \ ,
\end{align}
\end{subequations}
where%
\begin{align}
\bar m = \frac{12 B_0 \bar m_q}{{m_b^\star}^2} \ .
\end{align}
$ g^\star_{\Phi\pm}$, $g^\star_{\Xi}$ and $\Sigma^{\prime\star}$ are functions of the mean quark mass that can be taken from appendix~\ref{sect_gfunctions}. The divergencies occurring at linear order in the mean quark mass can be canceled via the following introduction of counterterms%
\begin{align}\label{counterterms_symmetric}
\Delta \Phi^\star_{\pm,i} &\longrightarrow \frac{{m_b^\star}^2 c^\star_{\Phi \pm}}{24 F_\star^2} \Phi^{\circ}_{\pm,i} L + \Delta \Phi^{\star,\text{ren.}}_{\pm,i}(\mu) \ , &
\Delta \Xi^\star_{\pm,i} &\longrightarrow \frac{{m_b^\star}^2 c^\star_{\Xi}}{24 F_\star^2} \Xi^{\circ}_{\pm,i} L + \Delta \Xi^{\star,\text{ren.}}_{\pm,i}(\mu) \ ,
\end{align}
where $L$ contains the divergence (see appendix~\ref{sect_gfunctions}) and the coefficients are%
\begin{align}
c^{\star}_{\Phi \pm} &= \frac{4}{3} \bigl( 6 (5 D^2 + 9 F^2) + 13 \pm 6 \bigr) \ , &
c^{\star}_{\Xi} &= \frac{4}{3} \bigl( 6 (5 D^2 + 9 F^2) + 9 \bigr) \ .
\end{align}
This leads to the following scale dependence in the renormalized amplitudes:%
\begin{align}
\mu \frac{\partial}{\partial \mu} \Delta \Phi^{\star,\text{ren.}}_{\pm,i}(\mu) &= \frac{-1}{(4\pi)^2} \frac{{m_b^\star}^2 c^\star_{\Phi \pm}}{24 F_\star^2} \Phi^\circ_{\pm,i} \ , &
\mu \frac{\partial}{\partial \mu} \Delta \Xi^{\star,\text{ren.}}_{\pm,i}(\mu) &= \frac{-1}{(4\pi)^2} \frac{{m_b^\star}^2 c^\star_{\Xi}}{24 F_\star^2} \Xi^\circ_{\pm,i} \ .
\end{align}
The divergencies occurring together with higher orders of the quark masses have to be canceled by hand as discussed in section~\ref{sect_minimal_parametrization}. If one takes eq.~\eqref{DAs_chpt_Result_symmetric_line} and plugs it into eq.~\eqref{DAs_chpt_Result_factorized} one finds (up to terms of higher order)%
\begin{subequations} \label{DAs_chpt_Result_factorized_mean_quark_mass}
\begin{align}
\begin{split}
\Phi_{\pm, i}^B &\mathrel{\dot =} \sqrt{Z_B}\Bigl( 1 + g^\star_{\Phi \pm} + \Delta g^B_{\Phi \pm} \Bigr) \Bigl( \Phi^\circ_{\pm,i} +\bar m \ \Delta\Phi_{\pm, i}^\star + \delta m \ \Delta \Phi_{\pm,i}^B \Bigr) \ ,
\end{split} \\*
\begin{split}
\Xi_{\pm, i}^B &\mathrel{\dot =} \sqrt{Z_B}\Bigl( 1 + g^\star_{\Xi } + \Delta g^B_{\Xi} \Bigr) \Bigl( \Xi^\circ_{\pm,i} +\bar m \ \Delta\Xi_{\pm, i}^\star + \delta m \ \Delta \Xi_{\pm,i}^B \Bigr) \ ,
\end{split} \\
\begin{split}
\Pi_{i}^B &\mathrel{\dot =}\sqrt{Z_B} \Bigl( 1 + g^\star_{\Phi \pm_B} + \Delta g^B_{\Pi} \Bigr) \Bigl( \Phi^\circ_{\pm_B,i} +\bar m \ \Delta\Phi_{\pm_B, i}^\star + \delta m \ \Delta \Pi_{i}^B \Bigr) \ ,
\end{split} \\*
\begin{split}
\Upsilon_{i}^B &\mathrel{\dot =} \sqrt{Z_B} \Bigl( 1 + g^\star_{\Xi} + \Delta g^B_{\Xi} \Bigr) \Bigl( \Xi^\circ_{\pm_B,i} +\bar m \ \Delta\Xi_{\pm_B, i}^\star + \delta m \ \Delta \Upsilon_{i}^B \Bigr) \ ,
\end{split}
\end{align}
\end{subequations}
where%
\begin{align}
\sqrt{Z_B} &\mathrel{\dot =} 1+ \frac{1}{2} \Sigma^{\prime\star}+\frac{1}{2}\Delta \Sigma^\prime_B \ .
\end{align}
Starting from this point everything can be worked out analogously to the case of fixed mean quark mass.%
\section{Summary} \label{sect_summary}
In this work we have presented the first analysis of baryon octet light-cone DAs in the framework of three-flavor BChPT. At next-to-leading order accuracy in the chiral counting scheme, we obtain the leading quark mass dependence and (automatically) the leading $\operatorname{SU}(3)_f$ breaking effects. Describing the baryon octet simultaneously we are able to unify and systemize the efforts made in refs.~\cite{Chernyak:1987nu,Braun:2000kw,Liu:2008yg,Liu:2009uc,Liu:2014uha}.\par
An important insight to be gained from our results is of qualitative nature: in the chiral odd sector the chiral behaviour (i.e.\ the contained chiral logarithms) of a specific moment does not depend on its twist, but on whether it contributes to the $\Phi^B_{+,i}$ or $\Phi^B_{-,i}$ amplitudes (see eq.~\eqref{definition_superior_DAs}). Those contributing to the ``$+$'' (``$-$'') amplitudes have the same chiral logarithms as $f^B$ ($\lambda_1^B$). Therefore the odd moments of the leading twist DA behave like $\lambda_1^B$ instead of (as one might have expected) $f^B$. This result is consistent with an earlier two-flavor calculation, where it was found that the odd first and second moments of the leading twist DA have the same chiral logarithms as $\lambda_{1}^N$ (see appendix of ref.~\cite{PhysRevD.89.094511}).\par
In section~\ref{sect_result} we provide a set of DAs that parametrize the complete baryon octet (including the $\Lambda$ baryon) in a minimal way and do not mix under chiral extrapolation. Eqs.~\eqref{DAs_chpt_Result_factorized} and~\eqref{DAs_chpt_Result_factorized_mean_quark_mass} are our main results. They describe the quark mass dependence of the baryon octet DAs (including all higher twist amplitudes) in a very compact manner. Eq.~\eqref{normalization_constants} contains explicit extrapolation formulas for the wave function normalization constants, while the dependence on the quark mass splitting of the shape parameters, which describe contributions of higher conformal spin, is shown in Eqs.~\eqref{shape_parameters_1} and~\eqref{shape_parameters_2}. The results will be of particular importance for the interpretation and extrapolation of forthcoming lattice QCD data, due to the significant decrease in number of parameters (compare section~\ref{sect_example_of_application}). For the same reason our results are relevant for QCD sum rule analyses and for model building.%
\acknowledgments{We thank P.~C.~Bruns, M.~Gruber, V.~M.~Braun and A.~N.~Manashov for valuable discussions.}
|
\section{Introduction}
The combinatorial game we study in the paper is played by two players, named Mr.\ Paint and Mrs.\ Correct, and is played on a finite, undirected graph in which each vertex has assigned a non-negative number representing the number of erasers at the particular vertex. We assume for simplicity that this number is initially the same for each vertex. In each round, first Mr.\ Paint selects a subset of the vertices and paints them all the same colour; he cannot use this colour in subsequent rounds. Mrs.\ Correct then has to erase the colour from some of the vertices in order to prevent adjacent vertices having the same colour. Whenever the colour at a vertex is erased, the number of erasers at that vertex decreases by $1$, but naturally, Mrs.\ Correct cannot erase the colour if she has no erasers left at that vertex. Vertices whose colours have not been erased can be considered as being permanently coloured and can be removed from the game. The game has two possible endings: (i) all vertices have been permanently coloured, in which case Mrs.\ Correct wins, or (ii) at some point of the game, Mr.\ Paint presents two adjacent vertices $u$ and $v$ and neither $u$ nor $v$ has any eraser left, in which case Mr.\ Paint wins. If, regardless of which sets she gets presented, there is a strategy for Mrs.\ Correct to win the game having initially $k-1$ erasers at each vertex, we say that the graph is \textbf{$k$-paintable}. The smallest $k$ for which the graph is $k$-paintable is called the \textbf{paintability} number of $G$, and denoted by $\chi_P(G)$. Note that this parameter is indeed well defined: for any graph on $n$ vertices, $n-1$ erasers at each vertex always guarantee Mrs.\ Correct to win, as she can always choose one vertex from a set presented to her and erase colours on the remaining ones. This problem is also known as the \textbf{on-line list colouring} and the corresponding graph parameter is also called the \textbf{on-line choice number of $G$}---see, for example,~\cite{Xuding} and below for the relation to the (off-line) list colouring.
\bigskip
Let us recall a classic model of random graphs that we study in this paper. The \textbf{binomial random graph} $\sG(n,p)$ is defined as the probability space $(\Omega, \mathcal{F}, \Pr)$, where $\Omega$ is the set of all graphs with vertex set $\{1,2,\dots,n\}$, $\mathcal{F}$ is the family of all subsets of $\Omega$, and for every $G \in \Omega$,
$$
\Pr (G) = p^{|E(G)|} (1-p)^{{n \choose 2} - |E(G)|} \,.
$$
This space may be viewed as the set of outcomes of ${n \choose 2}$ independent coin flips, one for each pair $(u,v)$ of vertices, where the probability of success (that is, adding edge $uv$) is $p.$ Note that $p=p(n)$ may (and usually does) tend to zero as $n$ tends to infinity.
All asymptotics throughout are as $n \rightarrow \infty $ (we emphasize that the notations $o(\cdot)$ and $O(\cdot)$ refer to functions of $n$, not necessarily positive, whose growth is bounded; whereas $\Theta(\cdot)$ and $\Omega(\cdot)$ always refer to positive functions). We say that an event in a probability space holds \textbf{asymptotically almost surely} (or \textbf{a.a.s.}) if the probability that it holds tends to $1$ as $n$ goes to infinity. We often write $\sG(n,p)$ when we mean a graph drawn from the distribution $\sG(n,p)$. Finally, for simplicity, we will write $f(n) \sim g(n)$ if $f(n)/g(n) \to 1$ as $n \to \infty$ (that is, when $f(n) = (1+o(1)) g(n)$).
\bigskip
Now, we will briefly mention the relation to other known graph parameters. A \textbf{proper colouring} of a graph is a labelling of its vertices with colours such that no two adjacent vertices have the same colour. A colouring using at most $k$ colours is called a (proper) \textbf{$k$-colouring}. The smallest number of colours needed to colour a graph $G$ is called its \textbf{chromatic number}, and it is denoted by $\chi(G)$. Let $L_k$ be an arbitrary function that assigns to each vertex of $G$ a list of $k$ colours. We say that $G$ is \textbf{$L_k$-list-colourable} if there exists a proper colouring of the vertices such that every vertex is coloured with a colour from its own list. A graph is \textbf{$k$-choosable}, if for every such function $L_k$, $G$ is $L_k$-list-colourable. The minimum $k$ for which a graph is $k$-choosable is called the \textbf{list chromatic number}, or the \textbf{choice number}, and denoted by $\chi_L(G)$. Since the choices for $L_k$ contain the special case where each vertex is assigned the list of colours $\{1, 2, \ldots,k\}$, it is clear that a $k$-choosable graph has also a $k$-colouring, and so $\chi(G) \leq \chi_L(G)$. It is also known that if a graph is $k$-paintable, then it is also $k$-choosable~\cite{Schauz}, that is, $\chi_L(G) \leq \chi_P(G)$. Indeed, if there exists a function $L_k$ so that $G$ is not $L_k$-list-colourable, then Mr.\ Paint can easily win by fixing some permutation of all colours present in $L_k$ and presenting at the $i$-th step all vertices containing the $i$-th colour of the permutation on their lists (unless the vertex was already removed before). Finally, it was shown in~\cite{Xuding} that the paintability of a graph $G$ on $n$ vertices is at most $\chi(G) \log n + 1$. (All logarithms in this paper are natural logarithms.) Combining all inequalities we get the following:
\begin{equation}\label{eq:chi}
\chi(G) \leq \chi_L(G) \le \chi_P(G) \le \chi(G) \log n + 1.
\end{equation}
\bigskip
It follows from the well-known results of Bollob\'as~\cite{Bol88}, \L uczak~\cite{Luc91} (see also McDiarmid~\cite{McD}) that the chromatic number of $\sG(n,p)$ a.a.s.\ satisfies
\begin{equation}\label{eq:chi2}
\chi(\sG(n,p)) \sim \frac {\log(1/(1-p))n}{2 \log (np)},
\end{equation}
for $np\to\infty$ and $p$ bounded away from $1$. The study of the choice number of $\sG(n,p)$ was initiated in~\cite{Alon92}, where Alon proved that a.a.s., the choice number of $\sG(n, 1/2)$ is $o(n)$. Kahn then showed (see~\cite{AlonS}) that a.a.s.\ the choice number of $\sG(n,1/2)$ equals $(1+o(1))\chi_{\sG(n,1/2)}$. In~\cite{Krivelevich}, Krivelevich showed that this holds for $p \gg n^{-1/4}$, and Krivelevich, Sudakov, Vu, and Wormald~\cite{KSVW02} improved this to $p \gg n^{-1/3}$. On the other hand, Alon, Krivelevich, Sudakov~\cite{AKS97} and Vu~\cite{Vu} showed that for any value of $p$ satisfying $2 < np \leq n/2$, the choice number is $\Theta(np/\log(np))$. Later, Krivelevich and Vu~\cite{KV02a} generalized this to hypergraphs; they also improved the leading constants and showed that the choice number for $C \leq np \leq 0.9n$ (where $C$ is a sufficiently large constant) is at most a multiplicative factor of $2+o(1)$ away from the chromatic number, the best known factor for $p \leq n^{-1/3}$. Our results below (see Theorem~\ref{thm:main}, Theorem~\ref{thm:main2}, and Theorem~\ref{thm:main3}) show that even for the on-line case, for a wide range of $p$, we can asymptotically match the best known constants of the off-line case. Moreover, if $np \ge \log^{\omega} n$ (for some function $\omega=\omega(n)$ tending to infinity as $n \to\infty$), then we get the same multiplicative factor of $2+o(1)$.
\bigskip
Our main results are the following theorems. The first one deals with dense random graphs.
\begin{thm}\label{thm:main}
Let $\varepsilon > 0$ be any constant, and suppose that
$$
(\log \log n)^{1/3} (\log n)^2 n^{-1/3} \ll p \leq 1-\varepsilon.
$$
Let $G \in \mathcal{G}(n,p)$. Then, a.a.s.,
$$
\chi_P(G) \sim \frac {n}{2 \log_b (np)} \sim \chi(G),
$$
where $b = 1/(1-p)$.
\end{thm}
Note that if $p=o(1)$, then
$$
\frac {n}{2 \log_b (np)} = \frac {n \log (1/(1-p))}{2 \log (np)} \sim \frac {np}{2 \log (np)} = \Theta \left( \frac {np}{\log (np)} \right).
$$
For constant $p$ it is not true that $\log (1/(1-p)) \sim p$ but the order is preserved, provided $p\leq 1-\varepsilon$ for some $\varepsilon>0$.
\bigskip
For sparser graphs we are less successful in determining the asymptotic behaviour of $\chi_P(\mathcal{G}(n,p))$. Nevertheless, we can prove the following two theorems that determine the order of the graph parameter we study.
\begin{thm}\label{thm:main2}
Let $\varepsilon > 0$ be any constant, and suppose that
$$
\frac {(\log n)^{2+\varepsilon}}{n} \leq p = O((\log \log n)^{1/3} (\log n)^2 n^{-1/3}).
$$
Let $G \in \mathcal{G}(n,p)$. Then, a.a.s.,
$$
\chi_P(G)=\Theta \left( \frac {np}{\log(np)} \right) =\Theta(\chi(G)).
$$
Moreover, if $np = (\log n)^{C+o(1)}$, a.a.s.\
$$
\chi(G) \le \chi_P(G) \le (1+o(1))
\begin{cases}
2 \chi(G) & \text{if } C \to \infty \\
\frac {2C}{C-2} \chi(G) & \text{if } C \in [4, \infty) \\
4 \chi(G) & \text{if } C \in (2,4).
\end{cases}
$$
\end{thm}
\bigskip
Finally, for very sparse graphs we have the following.
\begin{thm}\label{thm:main3}
Let $G \in \mathcal{G}(n,p)$ with $p=O(1/n)$. Then, a.a.s., $\chi_P(G)=\Theta(1)=\Theta(\chi(G))$.
\end{thm}
\section{Preliminaries}
Most of the time, we will use the following version of \textbf{Chernoff's bound}. Suppose that $X \in \Bin(n,p)$ is a binomial random variable with expectation $\mu=np$. If $0<\delta<1$, then
$$
\pr [X < (1-\delta)\mu] \le \exp \left( -\frac{\delta^2 \mu}{2} \right),
$$
and if $\delta > 0$,
\[\pr [ X > (1+\delta)\mu] \le \exp\left(-\frac{\delta^2 \mu}{2+\delta}\right).\]
However, at some point we will need the following, stronger, version: for any $t \ge 0$, we have
\begin{equation}\label{eq:strongChernoff}
\pr [X \ge \mu + t] \le \exp \left( - \mu \varphi \left( \frac {t}{\mu} \right) \right),
\end{equation}
where
$$
\varphi(x) = (1+x)\log(1+x)-x.
$$
These inequalities are well known and can be found, for example, in~\cite{JLR}.
\bigskip
Let $G=(V,E)$ be any graph. A set $S \subseteq V$ is called \textbf{independent} if no edge $e \in E$ has both endpoints in $S$. Denote by $\alpha(G)$ the \textbf{independence number} of $G$, that is, the size of a largest independent set of $G$. Let $k_0$ be defined as follows:
$$
k_0=k_0(n,p)=\max \left\{k \in \ensuremath {\mathbb N} : \binom{n}{k}(1-p)^{\binom{k}{2}} \geq n^4 \right\}.
$$
It is well known that $k_0$ is well defined (for all $n$ sufficiently large and provided that $p \le 1 - \varepsilon$ for some $\varepsilon>0$) and that $k_0 \sim 2 \log_b (np)$ with $b=1/(1-p)$. The following result was proved in~\cite{KSVW02}.
\begin{thm}[\cite{KSVW02}]\label{lem:Krivelevich}
Suppose $n^{-2/5} \log^{6/5} n \ll p \leq 1-\varepsilon$ for some constant $\varepsilon > 0$. Let $G \in \mathcal{G}(n,p)$. Then
$$
\pr(\alpha(G) < k_0) = \exp\left(-\Omega\left( \frac{n^2}{k_0^4 \; p}\right)\right).
$$
\end{thm}
We obtain the following immediate corollary that will be useful to deal with dense random graphs. In fact, this lemma is the bottleneck for the argument that prevents us to extend Theorem~\ref{thm:main} for sparser graphs.
\begin{cor}\label{cor:largesets}
Let $\varepsilon>0$ be any constant and let $\omega = \omega(n) = o(\log n)$ be any function tending to infinity as $n \to \infty$. Suppose that
$$
\omega (\log \log n)^{1/3} (\log n)^2 n^{-1/3} \leq p \leq 1-\varepsilon.
$$
Let $G \in \mathcal{G}(n,p)$. Then, a.a.s.\ every set $S \subseteq V(G)$ with $|S| =s \geq s_0 := n/(\omega \log^2 n)$ contains an independent set of size $k_0 = k_0(s,p) \sim k_0(n,p)$.
\end{cor}
\begin{proof}
Fix a set $S \subseteq V(G)$ with $|S| = s \geq n/(\omega \log^2 n)$. First, let us note that
$$
k_0 = k_0(s,p) \sim 2 \log_b (sp) = 2 \log_b (np) \left( 1 - O\left( \frac {\log \log n}{\log n} \right) \right) \sim k_0(n,p).
$$
Moreover, since $n/(\omega \log^2n) \le s \le n$, we can easily verify that $p$ satisfies $s^{-2/5} \log^{6/5} s \ll p \leq 1-\varepsilon$ (with room to spare in the lower bound). It follows immediately from Theorem~\ref{lem:Krivelevich} that the probability of not having an independent set of size $k_0$ in $S$ is at most
$$
\exp\left(- \Omega \left( \frac {s^2}{k_0^4 p} \right) \right) = \exp\left(- \Omega \left( \frac {s^2 p^3}{\log^4 n} \right) \right) = \exp\left(- \Omega \left( \frac {s n p^3}{\omega \log^6 n} \right) \right) = \exp\left(- \Omega \left( s \omega^2 \log \log n \right) \right).
$$
On the other hand, the number of sets of size $s$ can be bounded as follows:
$$
{n \choose s} \le \left( \frac {ne}{s} \right)^s = \exp \left( s \log (ne/s) \right) \le \exp \left( 3 s \log \log n \right),
$$
since $\log (ne/s) \le \log (e \omega \log^2 n) = (2+o(1)) \log \log n$. Hence, the expected number of sets of size $s$ for which the desired property fails is $\exp\left(- \Omega \left( s \omega^2 \log \log n \right) \right)$. Summing over all $s \geq s_0$, the expected number of all such sets is $\exp\left(- \Omega \left( s_0 \omega^2 \log \log n \right) \right) = o(1)$. The claim holds by Markov's inequality.
\end{proof}
\bigskip
Given a graph $G$ and some fixed ordering of its vertices (for example, we may assume that it is the natural ordering $1, 2, \ldots,n$), the greedy colouring algorithm proceeds by scanning vertices of $G$ in the given order and assigning the first available colour for the current vertex. We will use the following lemma due to~\cite{KV02} (see also~\cite{CojaTaraz03}). However, since we use it in a slightly different setting, we point out a few small differences in the argument by providing a sketch of the proof. In fact, this lemma is the bottleneck for the argument that prevents us to extend Theorem~\ref{thm:main2} for sparser graphs.
\begin{lem}\label{lem:largesets2}
Let $\omega = \omega(n) := \log \log n$. Given any constant $0 < \varepsilon < 1$, let $G \in \mathcal{G}(n,p)$ with $\log^{2+\varepsilon} n/n =: p_0 \leq p=o(1/\log n)$. Then, a.a.s.\ every subgraph of $G$ of size $u \geq u_0 := n/(\omega \log^2 n)$ has an independent set of size at least $\varepsilon(1-\varepsilon) \log (np)/(3p)$.
\end{lem}
Note that in the lemma we set $\omega=\log \log n$; other functions tending to infinity slower than $\log \log n$, constant, or even tending to $0$ not too quickly clearly would work as well, but would make the statement of the result weaker.
\begin{proof}[Sketch of the proof] We follow the notation as in~\cite{KV02} and apply the greedy approach given there for each subgraph of size $u$. Let
$$
\alpha_0=\frac{(1-\varepsilon) \log (up)}{p} \hspace{0.5in} \text{ and } \hspace{0.5in} t=\frac{up}{\log (up)}.
$$
Note that
\begin{eqnarray*}
\alpha_0 &\geq& \frac{(1-\varepsilon) \log (np/(\omega \log^2 n))}{p} = \frac{(1-\varepsilon) (\log (np) - (2+o(1)) \log \log n)}{p} \\
&\ge& \frac{(1-\varepsilon) (\varepsilon+o(1)) \log (np)}{(2+\varepsilon) p} \ge \frac{(1-\varepsilon)\varepsilon \log (np)}{3p},
\end{eqnarray*}
since $np \ge \log^{2+\varepsilon} n$. (Note that, if $np \ge \log^C n$, then we get the better estimate $$\alpha_0 \ge (1-\varepsilon)(1-2/C+o(1)) \log(np)/p;$$ see Remark~\ref{rem:densecase} below.) Moreover,
\begin{eqnarray*}
(1-p)^{\alpha_0} &=& \exp \left( -p \alpha_0 (1+O(p)) \right) = \exp \left( - (1-\varepsilon) \log(up) (1+O(p)) \right) \\
&\sim& \exp \left( - (1-\varepsilon) \log(up) \right) = (up)^{-1+\varepsilon},
\end{eqnarray*}
since $p = o(1/\log n)$. For a fixed subgraph of size $u$, it follows from~\cite{KV02} that the probability that the algorithm fails to produce an independent set of size at least $\alpha_0$ is at most
\begin{align*}
\exp\left(-t(1-p)^{\alpha_0}u\varepsilon\right)=&\exp \left( -\frac{(up) (1+o(1)) (up)^{-1+\varepsilon} u \varepsilon}{\log (up) } \right) \\
=& \exp \left( - \frac{u^{1+\varepsilon} p^{\varepsilon} (\varepsilon+o(1))}{\log (up)} \right).
\end{align*}
By taking a union bound over all
$$
\binom{n}{u} \leq \left( \frac {ne}{u} \right)^u = \exp(u \log(ne/u) ) \le \exp(3 u \log \log n)
$$
sets of size $u$, the probability $P_u$ that there exists a subgraph of size $u$ for which the algorithm fails is at most
$$
\exp \left( - u \left( \frac{(up)^{\varepsilon} (\varepsilon+o(1))}{\log (up)} - 3 \log \log n \right) \right) \leq \exp \left( - u \left( \frac{(u_0 p_0)^{\varepsilon} (\varepsilon+o(1))}{\log (u_0p_0)} - 3 \log \log n \right) \right),
$$
as the function $f(x) = x^{\varepsilon}/ \log x$ is increasing for large enough $x$. Since $u_0 p_0 = \log^{\varepsilon} n / \omega = \log^{\varepsilon+o(1)} n$,
\begin{align*}
P_u \le& \exp \left( - u \left( \frac {(\log n)^{(\varepsilon+o(1))\varepsilon} (\varepsilon+o(1))}{(\varepsilon+o(1)) \log \log n} - 3 \log \log n \right) \right) \\
=& \exp \left( - u \left( (\log n)^{\varepsilon^2+o(1)} - 3 \log \log n \right) \right) \le e^{-u}.
\end{align*}
Summing over all $u_0 \leq u \leq n$, we see that the probability that the algorithm fails is at most
$
\sum_{u=u_0}^n e^{-u} = O(e^{-u_0}) = o(1),
$
and the lemma follows.
\end{proof}
\bigskip
\begin{remark}\label{rem:densecase}
Lemma~\ref{lem:largesets2} is sufficient to determine the order of the on-line choice number for sparse random graphs. We comment on improvements in order to obtain the smallest leading constant in the upper bound. From the proof of the lemma it is clear that the bound on the size of the independent set can be improved for denser graphs. More precisely, for $np \geq \log^{\omega} n$ (for $\omega\to\infty$ but still $p = o(1/\log n)$), we can obtain independent sets of size at least $(1-2\varepsilon)\log(np)/p$ for any arbitrarily small $\varepsilon$, and so we are asymptotically just a factor $2+o(1)$ off from the bound of Corollary~\ref{cor:largesets}.
On the other hand, for $C$ being a constant larger than $2$, and $np=\log^{C+o(1)} n$, we can obtain independent sets of size at least $(1-\varepsilon)(1-2/C+o(1)) \log(np)/p$ (again, for any arbitrarily small $\varepsilon$), so by Lemma~\ref{lem:largesets2} we are off by a factor $2/(1-2/C)+o(1) = 2C/(C-2) + o(1)$. We will now show that for $\log^{2+\varepsilon} n/n \leq p=o(n^{-5/6})$, however, the independence number obtained is by at most a factor $4+o(1)$ off from the one obtained by Corollary~\ref{cor:largesets}. First, as shown above, given $u \ge n/(\omega \log^2 n)$, there are at most $\exp(3u \log \log n)$ sets of size $u$, and the expected number of edges induced by each such set is $\binom{u}{2}p$. By Chernoff's bound, the probability that the number of edges induced by one of them deviates by an additive $\binom{u}{2}p / \log\log n$ factor from its expected value is at most $\exp\left(-(1/\log \log n)^2 u^2 p/5\right)$. Hence, with probability at most
$$
\exp\left(u (3 \log \log n- (\log n)^{\varepsilon} (1/\log \log n)^2/(5\omega)) \right) \leq \exp\left(-u \right),
$$
there exists a set of size $u$ that does not satisfy the condition. Summing over all $n/(\omega \log^2 n) \leq u \leq n$, we see that a.a.s.\ for all such sets we have $(1+o(1)) u^2 p/2$ edges. On the other hand, note that for this range of $p$, the expected number of pairs of triangles sharing one vertex in $G$ is $\Theta(n^5 p^6)=o(1)$, and thus by Markov's inequality, a.a.s.\ every vertex is in at most one triangle. It follows that there exists a set $D$ of edges which is a matching such that after removing $D$ the graph is triangle-free. Since in $G$ the average degree of every set of size $u$ is $up(1+o(1))$, the same also holds for $G \setminus D$. Since $G \setminus D$ is triangle-free, by Shearer's lemma~\cite{Shearer}, an independent set of size $(1+o(1))\log (np)/p$ can be found. By eliminating at most half of the set (one vertex for each edge in $D$), we obtain an independent set in the original graph. It follows that for every set of size $u$, there exists an independent set of size at least $\log (np)/((2+o(1))p)$, being therefore at most a factor $4+o(1)$ off from the size obtained by Corollary~\ref{cor:largesets}.
\end{remark}
\bigskip
We will also need the following observation.
\begin{lem}\label{lem:mediumsets}
Let $G \in \mathcal{G}(n,p)$, for $10\log n/n \le p \le 1$, and set $s_0=10 \log n/p$. Then, a.a.s.\ the following holds:
\begin{itemize}
\item [(i)] every set $S \subseteq V(G)$ with $s_0 \leq |S| \le n$ contains an independent set of size at least $1/(9p)$;
\item [(ii)] for every $i\in\ensuremath {\mathbb N} $ and every set $S \subseteq V(G)$ with $2^{-i} s_0 \leq |S| < 2^{-i+1} s_0$, $S$ contains an independent set of size at least $i/(9p 2^i)$.
\end{itemize}
\end{lem}
Before we move to the proof of the lemma, let us note that the lower bound on $p$ is not used in the proof of part~(i), which becomes a trivial statement if $p<10 \log n/n$. For part~(ii), we could easily relax this bound to $p\ge 1/n^k$ (for any constant $k>0$) by changing some of the constants in the statement. Furthermore, part~(i) is trivially true for $p\ge1/9$, and part (ii) is only non-trivial for all $i$ satisfying $9p2^i<i$.
\remove{Before we move to the proof of the lemma, let us note that the lower bound for $p$ is not used in the proof of the result. It is introduced to make sure the statement is non-trivial. Similarly, the condition for $i$ in part (ii) is to make sure that we claim the existence of an independent set of size larger than 1.}
\begin{proof
For part (i), let us fix a set $S$ with $|S|=s$ satisfying $s_0 \le s\le n$. Denote by $E_S$ the set of edges induced by $S$. We have $\ex{|E_S|}=\binom{s}{2}p \sim s^2 p/2$. By Chernoff's bound, with probability at least $1-\exp(-\ex{|E_S|}/4)$, we have $|E_S| \leq s^2 p$. By taking a union bound over all $\binom{n}{s}\leq \exp(s \log n)$ subsets of size $s$, with probability at most $\exp(s (\log n - s p/9))$ there exists a set of size $s$ not satisfying the condition. Since, by assumption, $s \geq s_0 = 10 \log n/p$, this probability is at most $\exp(- s \log n / 9)$, and so summing over all $s \ge s_0$, the claim holds a.a.s.\ for all $s$ in the desired range. We may therefore assume that (deterministically) every subset of size $s \ge s_0$ satisfies $|E_S| \leq s^2 p$. Then, for any set $S$ in the desired range, at least $s/2$ vertices of $S$ have a degree of at most $4sp$
in the graph induced by $S$.
By applying a greedy algorithm to these vertices, we can therefore always find an independent set of size at least
$(s/2)/(4sp + 1) \ge 1/(9 p)$.
Part (ii) is proved in a similar way. Fix $i = i(n) \in \ensuremath {\mathbb N} $ such that
\begin{equation}\label{eq:i1}
9p 2^i < i
\end{equation}
(since otherwise the statement is trivial), and a set $S\subseteq V(G)$ of size $s$ satisfying the following:
\begin{equation}\label{eq:i2}
2^{-i} s_0 \le s < 2^{-i+1} s_0.
\end{equation}
Since $s$ is a natural number, we must have $2^{-i+1} s_0 \ge 1$. Therefore, $2^i \le 2 s_0 \le 2n$ (where we used that $p\ge10\log n/n$), and thus
\begin{equation}\label{eq:i3}
i \le \log(2n)/\log 2 \le 2 \log n,
\end{equation}
for $n$ large enough. Combining~\eqref{eq:i2}, \eqref{eq:i1}, the fact that $ps_0 = 10 \log n$, and~\eqref{eq:i3}, we get
\[
s \ge 2^{-i} s_0 > (9p/i) s_0 = (90/i) \log n \ge 45.
\]
Note that for $s\ge45$,
\[
\ex{|E_S|}=\binom{s}{2}p \ge \frac {s^2 p}{2.1}.
\]
It follows from the stronger version of Chernoff's bound~\eqref{eq:strongChernoff} that
\begin{eqnarray*}
\pr \left( |E_S| \ge \left( 1+\frac {2^i}{i} \right) \ex{|E_S|} \right) &=& \exp \left( - \varphi(2^i/i) \ex{|E_S|} \right) \\
&\le& \exp \left( - \frac {2^{i}}{4} \cdot \frac {s^2 p}{2.1} \right) \\
&\le& \exp \left( - \frac {2^{i} s^2 p}{9} \right).
\end{eqnarray*}
(Note that $\varphi(2^i/i) = (1+2^i/i)\log(1+2^i/i)-2^i/i \sim (2^i/i) \log(2^i/i) \sim 2^i \log 2$ as $i \to \infty$. Moreover, it is straightforward to see that for every $i \in \ensuremath {\mathbb N} $, $\varphi(2^i/i) \ge 2^i/4$.)
As before, by a union bound we get that with probability at most $\exp(s (\log n - 2^i s p/9))$ there exists a set of size $s$ not satisfying the condition. Since, by assumption, $s \geq 2^{-i} s_0 = 2^{-i} 10 \log n/p$, this probability is at most $\exp(- s \log n/9)$,
regardless of which precise interval $[2^{-i}s_0,2^{-i+1}s_0)$ contains $s$.
Summing the $\exp(- s \log n/9)$ bound we obtained over all values $45\le s < s_0$ gives $o(1)$.
Therefore, we may assume that (deterministically), for any $i\in \ensuremath {\mathbb N} $, every subset $S$ of size $s$ in the range~\eqref{eq:i2} satisfies
$$
|E_S| \leq \left(1+\frac {2^i}{i} \right) \binom{s}{2}p \le \left(1+\frac {2^i}{i} \right) \frac {s^2 p}{2} \le 2^i s^2 p / i.
$$
Arguing as before, this guarantees that we can always find an independent set of size at least
\[
(s/2)/(4 \cdot 2^i sp/i + 1) \ge i/(9 \cdot 2^i p).
\]
(At the last step, we used that $4 \cdot 2^i sp/i + 1 \le 4.2 \cdot 2^i sp/i$, which follows easily from~\eqref{eq:i2}, \eqref{eq:i3} and the definition of $s_0$.)
The proof of the lemma is finished.
\end{proof}
\section{Proof of Theorem~\ref{thm:main}}
The lower bound follows immediately from~\eqref{eq:chi}. For the upper bound, we give a winning strategy for Mrs.\ Correct that a.a.s.\ requires only $(1+o(1)) n / (2 \log_b (np))$ erasers on each vertex, where $b = 1/(1-p)$. (Recall from~\eqref{eq:chi2} that a.a.s.\ $\chi(G) \sim n / (2 \log_b (np))$.)
We emphasize that most probabilistic statements hereafter in the proof will not refer to $\sG(n,p)$ but rather to the randomized strategy that Mrs.\ Correct uses to select each independent set, regardless of the strategy of Mr.\ Paint and assuming that $G$ deterministically satisfies the conclusions of Corollary~\ref{cor:largesets} and Lemma~\ref{lem:mediumsets}.
Since $p \gg (\log \log n)^{1/3} (\log n)^2 n^{-1/3}$, we can choose a function $\omega=o(\log n)$ tending to infinity with $n$ and such that $p \ge \omega (\log \log n)^{1/3} (\log n)^2 n^{-1/3}$. Note that our choice of $\omega$ satisfies the requirements of Corollary~\ref{cor:largesets}.
Call a set $S \subseteq V(G)$ \textbf{large} if $|S|=s \geq n/(\omega \log^2 n)$, \textbf{small} if $s \leq np/(\omega \log^2 n)$, and \textbf{medium} otherwise.
\medskip
Whenever Mrs.\ Correct is presented a large set, she can, by Corollary~\ref{cor:largesets} find an independent set of size $k_0=(2+o(1)) \log_b (np)$, and uses erasers for all remaining vertices. Note that, trivially, at most
\begin{equation}\label{eq:large}
\frac {n}{(2+o(1)) \log_b (np)} = (1+o(1)) \frac {n}{2 \log_b (np)}
\end{equation}
large sets can be presented to her before the end of the game, and hence, for all large sets at most that many erasers on each vertex are needed.
\medskip
Suppose now that a small set $S$ is presented to Mrs.\ Correct. Then, she chooses a random vertex $v \in S$ and accepts the colour on that vertex; for all other vertices of the presented set erasers are used. (Note that this is clearly not a optimal strategy; Mrs.\ Correct could extend $\{v\}$ to a maximal independent set but we do not use it in the argument and so we may pretend that a single vertex is accepted.) Let $X_v$ denote the random variable counting the number of erasers used when small sets containing $v$ are presented before eventually $v$ gets a permanent colour (which does not necessarily happen when a small set is presented). We have
\begin{equation}\label{eq:smallset}
\pr \left( X_v \geq \frac {np}{\sqrt{\omega} \log n} \right) \leq \left(1- \frac{\omega \log^2 n}{np}\right)^{np/( \sqrt{\omega}\log n)} \leq \exp(-\sqrt{\omega}\log n) =o(n^{-1}),
\end{equation}
and thus, by a union bound, a.a.s.\ for all vertices the number of erasers used for small sets is at most
\begin{equation}\label{eq:small}
\frac {np}{\sqrt{\omega}\log n} =o \left( \frac{n}{2 \log_b (np)} \right),
\end{equation}
and so is negligible.
\medskip
Now, we are going to deal with medium sets. First, note that the size of a medium set is at least $np/(\omega \log^2 n) \ge 10 \log n / p$ for the range of $p$ considered in this theorem and by our choice of $\omega$. Suppose that some medium set $S$ is presented during the game. Applying Lemma~\ref{lem:mediumsets} repeatedly, Mrs.\ Correct can partition $S$ into independent sets of size $\lceil 1/(9p) \rceil$ and a remaining set $J$ of size at most $10 \log n / p \le np/(\omega \log^2 n)$. The strategy is then the following: with probability $1/2$ she chooses (uniformly at random) one of the independent sets of size $\lceil 1/(9p) \rceil$, and with probability $1/2$ she chooses one vertex chosen uniformly at random from $J$ (as before, this is clearly a suboptimal strategy but convenient to analyze). Selected vertices keep the colour; for the others erasers need to be used. We partition the vertices of $S$ into two groups: group $1$ consists of vertices belonging to independent sets, and group $2$ consists of vertices of $J$. Our goal is to show that each vertex $v$ appears in less than $3np/(\sqrt{\omega}\log n)$ many medium sets, so that the total number of erasers needed to deal with these situations is negligible (see~\eqref{eq:small}).
Suppose that some vertex $v$ appears in at least $3np/(\sqrt{\omega}\log n)$ medium sets. Each time, the corresponding medium set is split into two groups, and we cannot control in which group $v$ ends up. However, by Chernoff's bound, with probability $1-o(n^{-1})$, at least $np/(\sqrt{\omega}\log n)$ times the group to which $v$ belongs is chosen. We condition on this event. Note that if group $2$ is chosen and $v$ belongs to this group, the probability that $v$ is not selected is at most $1- \omega \log^2 n/(np)$ (as in~\eqref{eq:smallset} for small sets). Similarly, if group $1$ is chosen and $v$ belongs to this group, the probability that $v$ is not chosen, is at most $1 - \omega \log^2 n/(9np),$ as each medium set has size at most $n / (\omega \log^2 n)$.
Denote by $Y_v$ the random variable counting the number of erasers used for vertex $v$ corresponding to medium sets, before eventually $v$ gets a permanent colour. As in~\eqref{eq:smallset},
\begin{eqnarray*}
\pr \left(Y_v \geq \frac {3np}{\sqrt{\omega} \log n} \right) &\leq& o(n^{-1}) + \left(1- \frac{\omega \log^2 n}{9np}\right)^{(np)/( \sqrt{\omega}\log n)} \\
&\leq& o(n^{-1}) + \exp \left(- \frac {\sqrt{\omega}\log n}{9} \right)=o(n^{-1}),
\end{eqnarray*}
and thus, as before, by a union bound, a.a.s.\ the desired bound for the number of erasers used for medium sets holds for all vertices.
Combining bounds for the number of erasers used for large, medium, and small sets, we get that, regardless of the strategy used by Mr.\ Paint, Mrs.\ Correct uses at most $(1+o(1)) n/ (2 \log_b (np))$ erasers for each vertex, and the theorem follows.
\section{Proof of Theorem~\ref{thm:main2}}
As in the proof of Theorem~\ref{thm:main}, the lower bound follows immediately from~\eqref{eq:chi}, and so it remains to show that Mrs.\ Correct has a strategy that a.a.s.\ requires only $O(n / \log_b (np))$ erasers on each vertex (recall also~\eqref{eq:chi2}). We will use the same definitions for sets of being small, medium, and large as before, but we set here $\omega=\log \log n$; as pointed out right after Lemma~\ref{lem:largesets2}, other choices of $\omega$ are possible, but our choice gives the strongest result (as we assume the weakest condition). The argument for small sets remains exactly the same; a.a.s.\ it is enough to have $o(n / \log_b (np))$ erasers on each vertex to deal with all small sets that Mr.\ Paint may present. To deal with large sets, by Lemma~\ref{lem:largesets2}, since we aim for a statement that holds a.a.s., we may assume that whenever a large set is presented, Mrs.\ Correct can always find an independent set of size at least $\varepsilon(1-\varepsilon) \log (np)/(3p)$, keeps the colour on these vertices, and uses erasers for all remaining vertices. The number of large sets presented is therefore at most $3 (\varepsilon(1-\varepsilon))^{-1} np / \log(np) = O(n / \log_b (np))$, as needed. This is enough to determine the order of the on-line choice number but, if one aims for a better constant, then Remark~\ref{rem:densecase} implies that for $np = \log^{C+o(1)} n$ we are guaranteed to have an independent set of size $i=i(C)$, where
$$
i = (1+o(1))
\begin{cases}
\log (np)/p & \text{if } C \to \infty \\
(1-\frac {2}{C}) \log (np)/p & \text{if } C \in [4, \infty) \\
\frac {1}{2} \log (np)/p & \text{if } C \in (2,4).
\end{cases}
$$
We will show below that the contribution of medium sets is of negligible order, and hence the upper bound on the number of erasers needed for large sets will also apply (up to lower order terms) to the total number of erasers needed. As a result, we will get the following bounds: $(2+o(1) \chi(G)$ for $C \to \infty$, $(\frac {2C}{C-2} + o(1)) \chi(G)$ for $C \in [4, \infty)$, and $(4+o(1)) \chi(G)$ for $C \in (2, 4)$.
The strategy for medium sets has to be substantially modified. Suppose that a medium set $S$ of size $s$ is presented at some point of the game. Recall that $np/(\omega \log^2 n) < s < n/(\omega \log^2 n)$, where $\omega=\log\log n$. Since we aim for a statement that holds a.a.s., we may assume that Mrs.\ Correct can partition $S$ in the following way: by applying Lemma~\ref{lem:mediumsets}(i) repeatedly, as long as at least $s_0 = 10 \log n/p$ vertices are remaining, she can find independent sets of size $\lceil 1/(9p) \rceil$, and remove them from $S$. (Note that for sparse graphs it might happen that $s < s_0$ and so the lemma cannot be applied even once. In such a situation, we simply move on to the next step.) If the number of remaining vertices, $r$, satisfies $2^{-i} s_0 \leq r < 2^{-i+1} s_0$ for some $i = i(n) \in \mathbb{N}$ and $r > np/(\omega \log^2 n)$, then by Lemma~\ref{lem:mediumsets}(ii), she can find an independent set of size $\lceil i/(9p2^{i}) \rceil$. Then, she removes that independent set, and continues iteratively with the remaining vertices in $S$. Note that there are clearly $M \le \log n / \log 2 =O(\log n)$ different sizes of independent sets corresponding to different values of $i$. Finally, if $r \leq np/(\omega \log^2 n)$, she puts all the remaining vertices into a set $J$ (not necessarily independent), and stops partitioning.
We classify vertices in $S$ into types depending on the size of the set in the partition of $S$ to which they belong:
more precisely, we call vertices from independent sets of size $\lceil 1/(9p) \rceil$ to be of \textbf{type-$0$}; vertices from independent sets of size $\lceil i/(9p2^{i}) \rceil$ for some $1 \leq i \leq M$ are called of \textbf{type-$i$}; and vertices from the last set of vertices $J$ of size at most $np/(\omega \log^2 n)$ are called of \textbf{type-$(M+1)$}. Of course, these definitions depend on the particular time a set $S$ is presented by Mr.\ Paint. In particular,
if a vertex is presented several times by Mr.\ Paint, each time it might be of a different type. Finally, we classify vertices in $S$ into 3 groups: vertices of type-0 form \textbf{group-$0$}, vertices of type-$1$ up to type-$M$ form \textbf{group-$1$}, and vertices of type-$(M+1)$ form \textbf{group-$2$}.
Mrs.\ Correct now chooses with probability $1/3$ one of the three groups. If group-$0$ is selected, she then picks uniformly at random one independent set within the group. In case when group-$2$ is selected, she picks uniformly at random a single vertex inside this group. Finally, if group-$1$ is selected, she first picks a random type. The probability that type-$i$ is selected is equal to
$$
q_i := \frac {1/i}{\sum_{j=1}^{M} 1/j} \ge \frac {1/i}{\sum_{j=1}^{\log n / \log 2} 1/j} = \frac {1+o(1)}{i \log \log n}.
$$
Then, within the selected type, an independent set is selected uniformly at random. If the group or type chosen by Mrs.\ Correct has no vertices in it, she picks no vertex to colour permanently, which is clearly a suboptimal strategy.
Our goal is to show that a.a.s.\ each vertex $v$ appears in less than
$$
\frac {4np}{\sqrt{\omega}\log (np)} = o \left( \frac{n}{\log_b (np)} \right)=o(\chi(G))
$$
many medium sets, before its colour is accepted. If this holds, then the number of erasers needed for each vertex (due to medium sets) is negligible.
Suppose that some vertex $v$ appears in at least $4np/(\sqrt{\omega}\log (np))$ medium sets. Since one of the three groups is selected uniformly at random for each medium set, we expect $v$ to belong to the selected group at least $(4/3)np/(\sqrt{\omega}\log (np)) \ge \log^{2+\varepsilon+o(1)} n$ times. Hence, it follows from Chernoff's bound that, with probability $1-o(n^{-1})$, at least $np/(\sqrt{\omega}\log (np))$ times the group to which $v$ belongs is chosen. Call this event $E$. It remains to show that, conditional on $E$, $v$ will be permanently coloured with probability $1-o(n^{-1})$ within the first $np/(\sqrt{\omega}\log (np))$ times its group is picked for some medium set.
Note that if group-$0$ is selected and $v$ belongs to this group, the probability that $v$ is chosen to be permanently coloured, is at least
\begin{equation}\label{eq:g0}
\frac{1/(9p)}{n/(\omega \log^2 n)} = \frac{\omega \log^2 n}{9np}.
\end{equation}
Suppose then that group-$1$ is chosen, $v$ belongs to this group and is of type-$i$ for some $1 \leq i \leq M$.
The number of independent sets in the partition containing vertices of type-$i$ is at most
\[
\frac{2^{-i+1}s_0/2}{i/(9p2^i)} + 1 = \frac{9s_0p}{i} + 1 = \frac{90\log n}{i} + 1 \le \frac{100\log n}{i},
\]
where for the last step we used that $i\le M\le \log n/\log 2$. This time, the probability that $v$ is permanently coloured is at least
\begin{equation}\label{eq:g1}
q_i \frac{i}{100\log n} \ge \frac{1+o(1)}{100(\log n)(\log\log n)} \ge \frac{\omega \log(np) \log n}{np},
\end{equation}
where we used that $\omega=\log\log n$ and $np/(\log(np)) \ge \log^{2+\varepsilon/2}n$. Finally, if group-$2$ is chosen and $v$ belongs to this group, the probability that $v$ is permanently coloured is at least
\begin{equation}\label{eq:g2}
\frac{1}{np / (\omega \log^2 n)} = \frac{\omega \log^2 n}{np}.
\end{equation}
Summarizing~\eqref{eq:g0}, \eqref{eq:g1} and~\eqref{eq:g2}, whenever the group of $v$ is selected (regardless of which group it is), the probability that $v$ is chosen to be permanently coloured is at least
$$
\frac{\omega \log(np) \log n}{9np}.
$$
The rest of the argument works as before. Given a vertex $v$ which is presented in at least $4np/(\sqrt{\omega}\log (np))$ medium sets, denote by $Y_v$ the random variable counting the number of erasers used by $v$ (due to medium sets) before $v$ gets a permanent colour. Then,
\begin{eqnarray*
\pr \left( Y_v \geq \frac { 4 np}{\sqrt{\omega} \log (np)} \; \Big| \; E \right) &\leq& \left(1- \frac{\omega \log(np) \log n}{9np} \right)^{np/( \sqrt{\omega}\log (np))} \\
&\le& \exp \left(- \Omega \left(
\sqrt\omega \log n \right) \right)
= o(n^{-1}).
\end{eqnarray*}
Since $\pr(E)=1-o(n^{-1})$,
the unconditional probability that $Y_v \ge 4 np/(\sqrt{\omega} \log (np))$ is also $o(n^{-1})$. Hence, regardless of the strategy followed by Mr.\ Paint, we can take a union bound over all vertices that were presented in at least $4 np/(\sqrt{\omega} \log (np))$ medium sets, and deduce that a.a.s.\ for every vertex the number of erasers used (due to medium sets) is less than $4 np/(\sqrt{\omega} \log (np))$ and thus negligible.
The proof of the theorem is finished.
\section{Proof of Theorem~\ref{thm:main3}}
Before we move to the proof of Theorem~\ref{thm:main3}, let us state the following simple observation.
\begin{lem}\label{lem:trees}
Let $G$ be a graph whose components are all trees or unicyclic graphs. Then Mrs.\ Correct has a winning strategy using $1$ eraser at each vertex for tree components and using $2$ erasers at each vertex for unicyclic components.
\end{lem}
\begin{proof}
Since we may consider different components separately, we may assume that $G$ is connected. First assume that $G$ is a tree. For a contradiction, suppose that $1$ eraser at each vertex is not enough for Mrs.\ Correct to win on some tree, and consider a smallest such tree $T$. Clearly, $|T| \geq 2$, and let us consider a leaf $\ell$ of $T$. By minimality of $T$, Mrs.\ Correct has a winning strategy on $T \setminus \ell$. Then, she extends the strategy to $T$ as follows. The first time she is presented vertex $\ell$, she considers the optimal strategy when playing on the restriction of the set to $T \setminus \ell$. If the set yielded by that strategy does not contain the neighbour of $\ell$ in $T$, she follows this strategy and simply adds $\ell$ to the set, and plays for the rest of the game on $T \setminus \ell$. On the other hand, if the set does contain the neighbour of $\ell$, then she also follows the strategy on $T \setminus \ell$, but uses the eraser for $\ell$. Since the only neighbour of $\ell$ cannot appear later on, she can continue with her optimal strategy on $T \setminus \ell$ and simply adds $\ell$ the second time she is presented this vertex. We get a contradiction, and the proof of the first part is finished.
Similarly, suppose now that $G$ is unicyclic. As before, for a contradiction, suppose that $2$ erasers at each vertex are not enough for $G$, and consider a smallest unicyclic graph $U$ with this property. Clearly, $|U| \geq 3$. If $U$ contains a leaf $\ell$, then, as before, she plays optimally using two erasers on $U \setminus \ell$ and adapts her strategy exactly as before. If $U$ does not contain a leaf, then $U$ is a cycle. The first time she is presented a set, she picks one vertex and uses erasers for all other vertices. The rest of the game is played on a tree (in fact, a path), and she has still $1$ eraser at each vertex at her disposal. By the first part of the lemma, she has a winning strategy and the proof is finished.
\end{proof}
We come now back to the proof of Theorem~\ref{thm:main3}. It is well known that if $p < 0.99/n$, then a.a.s.\ $G$ contains only trees and unicyclic components, and Mrs.\ Correct can win using at most $2$ erasers by Lemma~\ref{lem:trees}. Assume then that $p\le c/n$ for some (perhaps large) constant $c \ge 0.99$ (assume w.l.o.g.\ that $c\in\ensuremath {\mathbb N} $). If $G$ contains a component with at least two cycles, the component must contain a subgraph which either consists of two cycles connected by a path (or sharing a vertex), or is a cycle with a ``diagonal'' path. Let us call such structures \textbf{complex}. Clearly, on $k$ vertices one can construct at most $k^2 k!$ complex structures. Note that the degree of any vertex given by the existence of the complex structure is at most $4$. We will show that a.a.s.\ there is no complex structure with the property that each vertex of this structure has a degree in the whole graph of at least $100c^2$. Note that, given a complex structure, in order for each vertex of this structure to have degree in the whole graph at least $100c^2$, it must have either at least $47c^2$ incident edges towards vertices outside the subgraph on which the complex structure is built, or it has at least $47c^2$ additional incident edges inside the subgraph on which the complex structure is built. Therefore, either half of the vertices of the complex structure have at least $47c^2$ incident edges outside, or there are at least $11c^2k$ additional edges inside the complex structure, and no information about these edges has been revealed so far. Thus, the expected number of complex structures in which each vertex has degree in the whole graph at least $100c^2$ in the whole graph is at most
\begin{align}
\sum_{k=4}^n {n \choose k} k^2 k! p^{k+1} &\left( \binom{k}{\lfloor k/2 \rfloor} \left( {n \choose 47c^2} p^{47c^2}\right)^{\lfloor k/2 \rfloor} + \binom{\binom{k}{2}} {11c^2k}p^{11c^2k}\right) \nonumber \\
& \leq \sum_{k=4}^n \frac{k^2}{n}c^{k+1} 2\left(\frac{e}{22 c}\right)^{11c^2k} = O\left( \sum_{k=4}^n \frac {k^2}{n} c^{k+1} c^{-2k} \right) \nonumber \\
&= O\left( \sum_{k=4}^n \frac {k^2}{n} c^{-k} \right) = O\left( \int_{x=0}^{\infty} \frac {x^2}{n}c^{-x} dx \right) = O(1/n),
\label{eq:densegraph}
\end{align}
and therefore, by Markov's inequality, the subgraph induced by vertices of degree at least $100c^2$ a.a.s.\ consists of components that are either trees or unicyclic components. Since we aim for a statement that holds a.a.s., we may assume that this is the case.
We may assume that each time Mrs.\ Correct selects a maximal independent set. Hence, the number of erasers needed to be placed at vertex $v$ is at most $\deg(v)$ (if $v$ uses one of its erasers at some point of the game, it must be the case that at least one of its neighbours belongs to the maximal independent set). As a result, vertices of degree at most $100c^2$ require only a constant number of erasers. Call the set of such vertices $L$, and let $H=V(G)\setminus L$. By~\eqref{eq:densegraph}, the graph induced by the vertices in $H$ a.a.s.\ consists of components that are either trees or unicyclic components. Mrs.\ Correct plays as follows: whenever she is presented a set $U$, she plays optimally on the restriction of $U$ to $H$ (on which, by Lemma~\ref{lem:trees} she uses at most 2 erasers), and then she extends an independent set found there to any maximal independent set in $U$. This strategy uses in total at most $100c^2=O(1)$ erasers for each vertex in $L$. In this way, clearly $O(1)=\Theta(\chi(G))$ colours are used.
|
\section{Introduction}
After Pauli and Weisskopf published their anti-Dirac paper\cite{Pauli1934}, the Klein-Gordon-Fock (K.G.F.) field theory became a well-respected concept
for describing the behavior of massive spinless scalar particles, like pions.
Homogeneous solutions of the K.G.F. equation as well as Green's functions for the inhomogeneous K.G.F. equation were worked out in detail.
It might not be well known that Pauli and Majorana never thought very highly of Dirac's hole theory.
More recently a paper was published in which Majorana, several years before Pauli and Weisskopf, studied the quantization of the relativistic K.G.F. equation\cite{Esposito2007}.
But the final blow to Dirac's hole theory came from Pauli and Weisskopf. Admittedly, it took a long time until Dirac's idea was relegated to the corner of "mere historical interest".
With the discovery of pions and other (pseudo-)scalar particles it became clear that one could do without Dirac holes for antiparticles.
Pauli and Weisskopf had always rejected the Dirac equation as an equation for a relativistic probability amplitude.
They regarded the Dirac equation as a relativistic matter field equation and not an equation of the probability amplitude in the $(x,y,z)$ space.
Although the idea of a second quantized field operator $\psi(\vec{x},t)$ for many-particle systems was clear to Jordan in the three-man paper\cite{Born1925} of 1925,
it was Pauli and Weisskopf who insisted that a concept like the probability of a particle to be found in $\vec{x}$ space does not make much sense for relativistic particles,
and this holds true for electrons, photons and K.G.F. particles alike.
Consequently, the Dirac equation as well as the K.G.F. equation should be treated as matter-field equations for many particles rather than as an equation for single-particle probability
amplitudes\cite{Tomonaga1997}.
After having constructed the relativistic scalar field theory, Pauli wondered "why nature had made no use of the possibility of the theory that there exist spin-zero bosons [...]".
Needless to say, Pauli's question was answered by nature in the affirmative. Not long after Pauli's and Weisskopf's paper of 1934, the relativistic scalar wave field theory would be finally
established in the appearance of $\pi$ mesons.
Admittedly, most of the facts in the present paper have been known for quite some time. But what is probably new is the attempt to construct the probability amplitude
$\langle\vec{x},t \vert \vec{x}_0,t_0=0\rangle$ in relativistic single-particle quantum mechanics(QM). We will discover that the probability amplitude of finding the particle
outside the light cone is not zero - it drops exponentially. So it can be found in an area which is forbidden by special relativity. To get out of this trouble we will use instead the
the language of second quantization. Here we will study the Pauli-Jordan commutator function $\left[ \psi(\vec{x},t),\psi^\dagger (\vec{x}_0)\right]$
which is related to the probability amplitude of detecting the particle at $(\vec{x},t)$ while it was created at $(\vec{x}_0,t_0=0)$.
Our first attempt ends up in a disaster. Not only have we violated causality - meaning no signal can travel faster than light. In addition, we find a breakdown of simultaneity, i.e., we cannot
give $\left[\psi,\psi^\dagger\right]$ an invariant meaning outside the light cone.
After we have remedied these failings, we begin a thorough study of the so-called invariant functions and propagation functions. Here we rely on the totally neglected but
wonderful article by J. Schwinger\cite{Schwinger1949}. Finally, we present a list of invariant functions of the K.G.F. theory in $(\vec{x},t)$ space.
\section{The Free Klein-Gordon-Fock Theory - Particle Description}
Our goal is to investigate the consequences of particles with relativistic energy spectrum ($\hbar=c=1$):
\begin{equation}
H:= \sum_{\vec{p}} a^\dagger(\vec{p})\sqrt{\vec{p}^2+m^2} a(\vec{p})\quad .\label{eq:DefHamiltonian}
\end{equation}
Now it is useful to recall that in the single-particle formalism we would start with the transition amplitude ($t_0=0$):
\begin{equation}
\langle \vec{x},t\vert \vec{p}\rangle = \frac{1}{\sqrt{V}} e^{i\left( \vec{p}\cdot\vec{x}-t \sqrt{\vec{p}^2+m^2}\right)} \quad ,\label{eq:TransAmpl}
\end{equation}
which satisfies the Schr\"{o}dinger equation
\begin{align}
i\frac{\partial}{\partial t}\langle \vec{x},t\vert \vec{p}\rangle &= \langle \vec{x},t\vert H \vert \vec{p}\rangle\notag \\
&= \sqrt{\vec{p}^2+m^2}\langle \vec{x},t\vert \vec{p}\rangle\quad .\label{eq:SchroedingerEqnBraKetNotation}
\end{align}
One might wonder as to whether it is reasonable to define a probability amplitude $\langle \vec{x},t\vert \vec{x}_0\rangle$ also in relativistic quantum mechanics.
To find out let us begin with
\begin{align*}
\langle \vec{x},t\vert \vec{x}_0\rangle &= \sum_{\vec{p}} \langle \vec{x},t\vert \vec{p}\rangle\langle\vec{p}\vert\vec{x}_0\rangle,\qquad \left(\langle\vec{p}\vert\vec{x}_0\rangle = \frac{1}{\sqrt{V}}e^{-i\vec{x}_0\cdot \vec{p}} \right)\\
&= \sum_{\vec{p}} \frac{1}{V}e^{i\left(\vec{p}\cdot (\vec{x}-\vec{x}_0) - t\sqrt{\vec{p}^2+m^2}\right)},\quad \vec{r}:=\vec{x}-\vec{x}_0\\
&\overset{V\rightarrow\infty}{=} \int \frac{1}{(2\pi)^3} e^{i\left(\vec{p}\cdot\vec{r}-t\sqrt{\vec{p}^2+m^2} \right)} d^3\vec{p}\\
&= \frac{1}{(2\pi)^3}\int_0^\infty p^2 dp\ 2\pi\int_{-1}^{1} e^{iprz}e^{-it\sqrt{\vec{p}^2+m^2}} dz\\
&\overset{z=\cos\Theta}{=} -\frac{4\pi}{(2\pi)^3}\frac{1}{r}\frac{\partial}{\partial r}\frac{1}{2}\int_{-\infty}^{\infty} e^{ipr}e^{-it\sqrt{\vec{p}^2+m^2}}dp\quad .
\end{align*}
Here we change variables: $p=m\sinh\phi,\ \cosh\phi=\sqrt{1+\sinh^2\phi}$ to obtain:
\begin{align}
\langle\vec{x},t\vert \vec{x}_0\rangle &= -\frac{1}{(2\pi)^2}\frac{1}{r}\frac{\partial}{\partial r}\int m\cosh\phi\ e^{i\left(mr\sinh\phi - mt\cosh\phi\right)}d\phi\notag\\
&= -\frac{1}{(2\pi)^2}\frac{1}{r}\frac{\partial}{\partial r}i\frac{\partial}{\partial t}\int e^{i\left(mr\sinh\phi - mt\cosh\phi\right)}d\phi\quad .\label{eq:TransAmplParametric}
\end{align}
If we believe in causality we want $\langle \vec{x},t\vert\vec{x}_0\rangle$ to be zero for two points in a space-like relation:
\begin{equation*}
\left( \vec{x}-\vec{x}_0\right)^2 -t^2 = r^2-t^2 >0 ,\quad \text{space-like} (r>t)\quad .
\end{equation*}
So let us compute the integral \eqref{eq:TransAmplParametric} for $r>t$, and since a particle of mass $m> 0$ cannot travel with the speed $v \geq c$, we expect the integral to vanish.
For $r>t$:
\begin{equation*}
\left.\begin{aligned}
mr &= \lambda \cosh\phi_0 \\
mt &= \lambda \sinh\phi_0
\end{aligned}\right\rbrace\quad \lambda:=m\sqrt{r^2-t^2}\quad .
\end{equation*}
In \eqref{eq:TransAmplParametric} we then obtain:
\begin{align*}
mr\sinh\phi-mt\cosh\phi &= m\sqrt{r^2-t^2}\left(\sinh\phi\cosh\phi_0-\cosh\phi\sinh\phi_0\right)\\
&= m\sqrt{r^2-t^2}\sinh\left(\phi-\phi_0\right)\quad .\\
\intertext{Changing $\phi - \phi_0\rightarrow\phi$ we obtain for \eqref{eq:TransAmplParametric}: }
\int e^{im\sqrt{r^2-t^2}\sinh\phi}d\phi &= 2\int_0^\infty \cos\left( m\sqrt{r^2-t^2}\sinh\phi\right)d\phi\\
&\overset{\psi:=\sinh\phi}{=} 2\int_0^\infty \frac{\cos\left(m\sqrt{r^2-t^2}\psi\right)}{\sqrt{1+\psi^2}}d\psi\\
&= 2 K_0 (m\sqrt{r^2-t^2})\quad .
\end{align*}
This yields
\begin{equation}
\langle \vec{x},t\vert\vec{x}_0\rangle = -\frac{2i}{(2\pi)^2} \frac{1}{r}\frac{\partial}{\partial r}\frac{\partial}{\partial t}K_0(m\sqrt{r^2-t^2})\quad .\label{eq:TransAmplParametricSimplified}
\end{equation}
But $K_0(m\sqrt{r^2-t^2})$ and derivatives thereof are unequal zero for $r^2>t^2$, e.g. if $r\gg t$, then:
\begin{equation}
K_0(m\sqrt{r^2-t^2})\cong \sqrt{\frac{\pi}{2m\sqrt{r^2-t^2}}} e^{-m\sqrt{r^2-t^2}}\quad .\label{eq:Asymptotic_K0}
\end{equation}
Hence, the probability amplitude for finding the particle outside the light cone is non-zero - it drops exponentially. Hence it can be found in an area which is
forbidden by special relativity. We are in great trouble.
Now, let us use instead the language of second quantization.
We create a particle at $(\vec{x}_0,t_0=0)$ and detect it at $(\vec{x},t)$:
\begin{align*}
\psi^\dagger(\vec{x}_0) &= \sum_{\vec{p}'} \frac{1}{\sqrt{V}}a^\dagger(\vec{p}')e^{-i\vec{p}'\cdot \vec{x}_0}\\
\psi(\vec{x},t) &= \sum_{\vec{p}} \frac{1}{\sqrt{V}}a(\vec{p})e^{i(\vec{p}\cdot \vec{x}-t\sqrt{\vec{p}^2+m^2})}\quad .
\end{align*}
We wish to calculate the commutator, which is related to the probability amplitude of finding the particle at $(\vec{x},t)$ when it came into existence
at $\vec{x}_0$ at time $t_0=0$:
\begin{equation}
\left[ \psi(\vec{x},t),\psi^\dagger (\vec{x}_0)\right] = \sum_{\vec{p}} \frac{1}{V}e^{i\left(\vec{p}\cdot (\vec{x}-\vec{x}_0) - t\sqrt{\vec{p}^2+t^2}\right)}\quad .\label{eq:PsiFldCommutatorDiffTime}
\end{equation}
This looks exactly like the expression we had for $\langle \vec{x},t\vert\vec{x}_0\rangle$ in the single-particle description. Therefore the commutator will not vanish if the two points
$\vec{x},\vec{x}_0$ are in a space-like relation (c.f. Eq. \eqref{eq:TransAmplParametricSimplified}). Even worse: for equal times we obtain:
\begin{equation*}
\left[ \psi(\vec{x},t),\psi^\dagger (\vec{x}_0)\right]_{t=0} = \delta^3(\vec{x}-\vec{x}_0)\quad .
\end{equation*}
Not only have we violated causality - meaning no signal can travel faster than light. In addition, we find a breakdown of simultaneity, i.e., we cannot give the commutator an invariant
meaning outside the light cone. In order to remedy the whole situation, let us start anew by finding a relativistic invariant expression for the commutator.
The first step is to write
\begin{align*}
\left[ \psi(\vec{x},t),\psi^\dagger (\vec{x}_0)\right] &= \int \frac{1}{(2\pi)^3}e^{i\left(\vec{p}\cdot(\vec{x}-\vec{x}_0)-t\sqrt{\vec{p}^2+m^2}\right)}d^3\vec{p}\\
&= \int\frac{d^3\vec{p}}{(2\pi)^3}\int dp^0 e^{i\left(\vec{p}\cdot(\vec{x}-\vec{x}_0) - p^0t\right)}\Theta(p^0)\delta(p^0-\sqrt{\vec{p}^2+m^2})\\
= \int \frac{d^3\vec{p}}{(2\pi)^2}\int dp^0 e^{i\left(\vec{p}\cdot(\vec{x}-\vec{x}_0) - p^0t\right)} &\Theta(p^0) \delta\left[ (p^0-\sqrt{\vec{p}^2+m^2})(p^0+\sqrt{\vec{p}^2+m^2})\right]
2\sqrt{\vec{p}^2+m^2}
\end{align*}
or
\begin{align}
\left[ \psi(\vec{x},t),\psi^\dagger (\vec{x}_0)\right] &= \int\frac{d^3\vec{p}}{(2\pi)^3}\int dp^0 \Theta(p^0)2\sqrt{\vec{p}^2+m^2}\delta\left({p^0}^2-\vec{p}^2-m^2\right)
e^{i\left(\vec{p}\cdot(\vec{x}-\vec{x}_0) - p^0t\right)}\notag \\
&= \int\frac{1}{(2\pi)^3}\Theta(p^0) 2\sqrt{\vec{p}^2+m^2} \delta(p^2+m^2)e^{ip\cdot(x-x_0)} d^4p\quad .\label{eq:PsiFldCommutatorRelInv}
\end{align}
In our metric, $p^\mu =(\vec{p},p^0),p^2=\vec{p}^2-{p^0}^2,(x-x_0)=(\vec{x}-\vec{x}_0,t)$.
Without the factor $2\sqrt{\vec{p}^2+m^2}$ the integral in \eqref{eq:PsiFldCommutatorRelInv} is a Lorentz scalar, i.e., an invariant function under proper ortochronous Lorentz transformation:
\begin{equation*}
\int \frac{1}{(2\pi)^3}\Theta(p^0)\delta(p^2+m^2)e^{ip\cdot y} d^4p = F(y^2)\quad .
\end{equation*}
From our decomposition of the $\delta$-function,
\begin{equation*}
\Theta(p^0)\frac{1}{2\sqrt{\vec{p}^2+m^2}}\left[ \delta(p^0-\sqrt{\vec{p}^2+m^2}) + \delta(p^0+\sqrt{\vec{p}^2+m^2})\right] = \Theta(p^0)\delta(p^2+m^2)\quad ,
\end{equation*}
we see that the integral in \eqref{eq:PsiFldCommutatorRelInv} picks up a contribution from the top sheet of the mass shell hyperboloids.
Finally, to get rid of the factor $2\sqrt{\vec{p}^2+m^2}$ in \eqref{eq:PsiFldCommutatorRelInv}, we define a new operator:
\begin{equation}
\phi(\vec{x},t) := \sum_{\vec{p}} \frac{1}{\sqrt{V}}e^{i\left(\vec{p}\cdot\vec{x}-\omega_pt\right)}\frac{1}{\sqrt{2 \left(\vec{p}^2+m^2 \right)^{\scriptstyle \frac{1}{2} } } }a(\vec{p}),\quad \omega_p:=\sqrt{\vec{p}^2+m^2}\quad .\label{eq:PhiOpDef}
\end{equation}
The new commutator function is then given by
\begin{align}
\left[ \phi(\vec{x},t), \phi^\dagger (\vec{x}_0)\right] &= \int \frac{d^3\vec{p}}{(2\pi)^3}\frac{1}{2\sqrt{\vec{p}^2+m^2}} e^{i\left(\vec{p}\cdot(\vec{x}-\vec{x}_0)-t\sqrt{\vec{p}^2+m^2}\right)} \notag\\
&= \int\frac{d^4p}{(2\pi)^3}\Theta(p^0)\delta(p^2+m^2)e^{ip\cdot(x-x_0)}\quad ,\label{eq:PhiCommutatorR3,1}
\end{align}
which is an invariant function. Therefore $\phi$ and not $\psi$ is the appropriate operator.
For later purposes it is useful to calculate the $\int dp^0$ term in \eqref{eq:PhiCommutatorR3,1}, which reveals the invariant measure in momentum space:
\begin{equation}
d\omega_p = \frac{1}{2\omega_p}\frac{d^3\vec{p}}{(2\pi)^3},\quad \omega_p=\sqrt{\vec{p}^2+m^2}\quad .\label{eq:InvMeasureMomentumSpace}
\end{equation}
Having given $[\phi,\phi^\dagger]$ an invariant meaning we look at it in the frame where $t=0$:
\begin{align}
\left[\phi(\vec{x},t),\phi^\dagger(\vec{x}_0)\right]_{t=0} &= \int \frac{d^3\vec{p}}{(2\pi)^3} \frac{1}{2\sqrt{\vec{p}^2+m^2}}e^{i\vec{p}\cdot(\vec{x}-\vec{x}_0)}\notag\\
&= \int e^{i\vec{p}\cdot (\vec{x}-\vec{x}_0)}d\omega_p\quad ,\label{eq:PhiCommutatorSimultaneous}
\end{align}
which is not zero!
So our new commutator, although relativistic invariant, still violates causality (compare to the discussion of the transformation amplitude $\langle\vec{x},t\vert\vec{x}_0\rangle$).
Our next goal is therefore to restore causality.
By the way, we can use the operator \eqref{eq:PhiOpDef},
\begin{equation*}
\phi(\vec{x},t)=\frac{1}{\sqrt{V}}\sum_{\vec{p} } e^{i\left(\vec{p}\cdot\vec{x}-\omega_pt\right)} \frac{1}{2\sqrt{\omega_p}}a(\vec{p})
\end{equation*}
to build a wave packet,
\begin{equation}
\langle 0\vert \phi(\vec{x},t) = \sum_{\vec{p}} \frac{1}{\sqrt{V}}e^{i\left(\vec{p}\cdot\vec{x}-\omega_pt\right)} \frac{1}{2\sqrt{\omega_p}}\langle \vec{p}\vert\quad
\bigl(\langle 0\vert a(\vec{p})=\langle\vec{p}\ \vert\bigr)\quad .\label{eq:PhiWavePacket}
\end{equation}
The inner product of two wave packets is
\begin{align}
\langle 0\vert \phi(\vec{x},t)\phi^\dagger(\vec{x}_0)\vert 0\rangle &= \langle 0 \vert \left[ \phi(\vec{x},t),\phi^\dagger(\vec{x}_0)\right]\vert0\rangle\notag\\
&= \int \frac{d^4p}{(2\pi)^3}\Theta(p^0)\delta(p^2+m^2)e^{ip\cdot(x-x_0)}\quad ,\label{eq:PhiWavePacketsIP}
\end{align}
which means that the localization of a particle has an invariant meaning, i.e., again, looks the same for all observers.
However, the description is still complicated from the point of causality.
Although $[\phi,\phi^\dagger]$ is Lorentz invariant, it does not vanish at $t=0$:
\begin{equation}
\left[\phi(\vec{x},t),\phi^\dagger(\vec{x}_0)\right] = \int \frac{d^3\vec{p}}{(2\pi)^3} \frac{1}{2\omega_p}e^{i\vec{p}\cdot (\vec{x}-\vec{x}_0)}e^{-i\omega_p t}\quad .\label{eq:PhiCommutatorR3}
\end{equation}
Recall that this expression was obtained by using $\phi,\phi^\dagger$ given above and the commutation relation $[a,a^\dagger]=1$.
In order to construct a commutator that vanishes for $t=0$ we have to subtract something from our former expression \eqref{eq:PhiCommutatorR3}, namely the second term in
\begin{equation}
\int \frac{d^2\vec{p}}{(2\pi)^3}\frac{1}{2\omega_p}\left( e^{i\vec{p}\cdot(\vec{x}-\vec{x}_0)} e^{-i\omega_pt} - e^{-i\vec{p}\cdot(\vec{x}-\vec{x}_0)} e^{i\omega_pt}\right)\quad .\label{eq:PhiCancellationExpression}
\end{equation}
For the second term we use the lower sheet of the hyperboloid, i.e., take the solution $p^0 = -\sqrt{\vec{p}^2+m^2}$. Then $\Theta(-p^0)\delta(p^2+m^2)$ selects the bottom sheet as
$\Theta(p^0)\delta(p^2+m^2)$ picks out the top sheet - our situation so far.
Now comes the point: the minus sign in \eqref{eq:PhiCancellationExpression} plays a significant role. It is the same minus sign that occurs if we look at $[a,a^\dagger]=1$ and write instead
$[a^\dagger,a]=-1$.
Therefore, if we interchange the role of creation and destruction operators, we can convert the minus sign into a plus sign:
\begin{equation}
\tilde{\phi}(\vec{x},t)=\frac{1}{\sqrt{V}}\sum_{\vec{p}} \frac{1}{\sqrt{2\omega_p}} \left( e^{i(\vec{p}\cdot\vec{x}-\omega_pt)}a(\vec{p})+e^{-i(\vec{p}\cdot\vec{x}-\omega_pt)}b^\dagger(\vec{p})\right)\quad ,\label{eq:PhiOpDefMultiOp}
\end{equation}
with $[a,a^\dagger]=1,[a,b]=0,[b,b^\dagger]=1,\omega_p=+\sqrt{\vec{p}^2+m^2}$.
Then, with this new object $\tilde{\phi}\rightarrow\phi$ we obtain:
\begin{equation*}
\left[\phi(\vec{x},t),\phi^\dagger(\vec{x}_0)\right] = \int \frac{d^2\vec{p}}{(2\pi)^3}\left( e^{i\vec{p}\cdot(\vec{x}-\vec{x}_0)} e^{-i\omega_pt} -
e^{-i\vec{p}\cdot(\vec{x}-\vec{x}_0)} e^{i\omega_pt}\right)\quad ,
\end{equation*}
which yields at last $[\phi(\vec{x},t),\phi^\dagger(\vec{x}_0)]_{t=0} = 0$ for $\vec{x}\neq \vec{x}_0$ and $\vec{x} = \vec{x}_0$.
The invariant form of $[\phi,\phi^\dagger]$ becomes obvious when we write ($y:=(\vec{x}-\vec{x}_0,t)$):
\begin{align}
\left[\phi(\vec{x},t),\phi^\dagger(\vec{x}_0)\right] &= \int \frac{d^3pdp^0}{(2\pi)^3}\left(\Theta(p^0)\delta(p^2+m^2)e^{ip\cdot y}-\Theta(-p^0)\delta(p^2+m^2)e^{ip\cdot y}\right)\notag\\
&=: i\Delta(y)\quad .\label{eq:PJCFDef}
\end{align}
This so-called Pauli-Jordan invariant commutator function is zero for $t=0$ and invariant for space-like distances $y^2>0$ with $y^2=(\vec{x}-\vec{x}_0)^2-t^2$.
Using
\begin{equation}
\epsilon(p^0) = \left\lbrace \begin{array}{lr}
+1 &,\quad p^0>0\\
-1 &,\quad p^0<0
\end{array}\right\rbrace =\Theta(p^0)- \Theta(-p^0)\quad ,
\end{equation}
we obtain the final form
\begin{equation}
\left[\phi(\vec{x},t),\phi^\dagger(\vec{x}_0,0)\right] = i\Delta(y) = \int \frac{d^4p}{(2\pi)^3}\epsilon(p^0)\delta(p^2+m^2)e^{ip\cdot y}\quad .\label{eq:PhiCommutatorFinalForm}
\end{equation}
If we calculate $\langle 0\vert \phi(\vec{x},t)\phi^\dagger(\vec{x}_0,0)\vert 0\rangle$ and insert the expressions for our old $\tilde{\phi},\tilde{\phi}^\dagger$ from \eqref{eq:PhiOpDefMultiOp},
the result becomes identical to our old calculation with the $\phi$ from \eqref{eq:PhiWavePacketsIP}. Therefore the vacuum expectation value is the same - since we assume that the
particles associated with the operators $a$ and $b$ have the same mass.
If we take $b= a$, then
\begin{equation}
\phi(\vec{x},t)=\frac{1}{\sqrt{V}}\sum_{\vec{p}} \frac{1}{2\sqrt{\omega_p}} \left( e^{i(\vec{p}\cdot\vec{x}-\omega_pt)}a(\vec{p}) + e^{-i(\vec{p}\cdot\vec{x}-\omega_pt)}a^\dagger(\vec{p})\right)
\quad ,\label{eq:PhiOpDefMultiOpSameType}
\end{equation}
meaning $\phi^\dagger=\phi$, i.e., $\phi$ is Hermitean.
If we consider \eqref{eq:PhiOpDefMultiOp} again and calculate the derivative $i\tfrac{\partial\phi}{\partial t}$, we find that it does not satisfy a first-order differential equation but
a second-order one:
\begin{equation}
\frac{\partial^2}{\partial t^2}\phi(\vec{x},t) = \left(\vec{\nabla}^2-m^2\right) \phi(\vec{x},t)\quad ,\label{eq:KGDiffEq2ndOrder}
\end{equation}
which is local in space-time; $a$ and $b^\dagger$ correspond to the two constants of integration.
The equation
\begin{equation}
\left(\frac{\partial}{\partial t}\right)^2\phi = \left(\vec{\nabla}^2-m^2\right)\phi\ ,\label{eq:KGFSEqn}
\end{equation}
is called the Klein-Gordon-Fock-Schr\"{o}dinger (K.G.F.Sch.) equation. It is a local relativistic equation.
\textbf{Remark} When we constructed the local operator $\phi$ we used $[a,a^\dagger]=1$. At one point it was necessary - in order to restore causality - to use the minus sign in
$[b^\dagger, b]=-1$.
If we now would go back and use instead an anti-commutation relation $aa^\dagger+ a^\dagger a=\lbrace a,a^\dagger\rbrace=1$, our procedure to construct a local field operator would fail.
The particles have to have spin.
Conventional textbooks start with the K.G.F.Sch. equation
\begin{equation}
\left( -\partial^2+m^2\right)\phi(\vec{x},t) = \left[ \left(\frac{\partial}{\partial t}\right)^2 - \vec{\nabla}^2 + m^2\right] \phi(\vec{x},t) = 0\label{eq:KGFSEqnR31}
\end{equation}
and look for solutions with the two constants of integration specified by $\phi(\vec{x},0)$ and $\dot{\phi}(\vec{x},0)$ and find that \eqref{eq:PhiOpDefMultiOp}
is the local solution to the K.G.F.Sch. equation. This is the path that mathematicians would take. Our procedure is more suited to physics-minded students and teachers.
We found in \eqref{eq:PhiCommutatorFinalForm} a most important invariant function $\Delta$ (Pauli-Jordan). Let us write it as
\begin{equation*}
\Delta(x,\kappa_0^2) = -i (2\pi)^{-3} \int (dk) e^{ikx} \epsilon(k^0)\delta(k^2+\kappa_0^2),
\end{equation*}
where $(dk)=dk_0dk_1dk_2dk_3$.
But there is another invariant function:
\begin{equation*}
\Delta^{(1)}(x,\kappa_0^2) = (2\pi)^{-3} \int (dk) e^{ikx}\delta(k^2+\kappa_0^2)\quad .
\end{equation*}
Evidently, both are solutions to the K.G. equation:
\begin{equation*}
(\kappa_0^2-\partial^2 ) \lbrace \Delta, \Delta^{(1)}\rbrace = 0\quad .
\end{equation*}
The two functions fulfill different boundary conditions and different symmetry properties. In fact,
\begin{align*}
\Delta(-x) &= -\Delta(x),\quad\text{since }\epsilon\text{ is odd}\\
\Delta^{(1)} (-x) &=\Delta^{(1)} (x)\quad .
\end{align*}
They both share invariance under proper ortochronous Lorentz transformation:
\begin{equation*}
\left( \Delta, \Delta^{(1)}\right)(\Lambda x) = \left( \Delta, \Delta^{(1)}\right) (x)\quad .
\end{equation*}
The basic fact about $\Delta$ is that it is zero outside the light cone while $\Delta^{(1)}$ reaches into the space-like sector where it dies out exponentially.
In other words, microcausality is realized by $\Delta$ and not determined by $\Delta^{(1)}$! It is due to the behavior of $\Delta$, i.e., the disappearance of the commutator of the
fields, that measuring the field at $x_0$ can have no consequence on measuring the field at $x$ since the points are not causally connected.
However both $\Delta$ and $\Delta^{(1)}$ are the basic functions for constructing the remaining invariant functions. More about their explicit expression in coordinate space will be
given and thoroughly discussed in the remaining chapters.
\section{Selection of invariant commutation and propagation functions}
In the last chapter, the Pauli-Jordan commutation function was constructed, starting from a scalar field:
\begin{equation}
\Delta(x) = -i\frac{1}{(2\pi)^3}\int e^{ik\cdot x}\epsilon(k_0)\delta(k^2+\mu^2)(dk)\quad .\label{eq:PJCFDef2.1}
\end{equation}
We also mentioned a second invariant function:
\begin{equation}
\Delta^{(1)}(x) = \frac{1}{(2\pi)^3}\int e^{ik\cdot x}\delta(k^2+\mu^2)(dk)\quad .
\end{equation}
Although they both satisfy the K.G.F. equation, they play a totally different role in scalar quantum field theory (Q.F.T.). While $\Delta(x)$ vanishes if
$x^2>0$ (space-like argument), $\Delta^{(1)}(x)$ does not vanish for space-like distances(c.f. Appendix \ref{sec:Appendix} ). Instead, $\Delta^{(1)}$ extends into the space-like region,
dropping off on the scale of the Compton wavelength $\tfrac{1}{\mu}$. However, it is also the famous Feynman propagator function that reaches into the space-like region.
In our convention, $\Delta_F (x) = \Delta^{(1)}(x)+i\epsilon(x)\Delta(x),\ \Delta_F =:2i\Delta_c$.
Note that the inhomogeneous $\Delta_F$ is constructed from the two homogeneous invariant functions $\Delta$ and $\Delta^{(1)}$. For our purposes we will use
the momentum representation of $\Delta_c(x)$:
\begin{equation}
\Delta_c(x) = \frac{1}{(2\pi)^4}\int \frac{e^{ik\cdot x}}{k^2+\mu^2-i\epsilon} (dk)\quad .\label{eq:DeltaCMomentumRepresentation}
\end{equation}
Employing
\begin{equation*}
\frac{1}{k^2+\mu^2-i\epsilon} = i\int_0^\infty e^{-is(k^2+\mu^2-i\epsilon}ds\quad ,
\end{equation*}
we obtain:
\begin{align}
\Delta_c(x) &= \int\frac{(dk)}{(2\pi)^4} i\int_0^\infty e^{-is(k^2+\mu^2)} e^{ik\cdot x}ds\notag\\
&= \frac{i}{(2\pi)^4}\int_0^\infty ds\ e^{-is\mu^2} \int e^{-isk^2+ik\cdot x}(dk)\quad .\label{eq:DeltaCMomentumHilbertTransform}
\end{align}
The $k$-integral in the previous equation is given by
\begin{equation}
\int e^{-isk^2+ik\cdot x}(dk) = -i\frac{\pi^2}{s^2} e^{i\frac{x^2}{4s}}\quad .
\end{equation}
Therefore
\begin{align*}
\Delta_c(x) &= \frac{1}{(2\pi)^4}\int_0^\infty e^{-is\mu^2}\frac{\pi^2}{s^2}e^{i\frac{x^2}{4s}} ds = \frac{1}{16\pi^2}\int_0^\infty \frac{1}{s^2} e^{-is\mu^2}e^{i\frac{x^2}{4s}}ds\\
\Delta_c(x) &: \begin{cases}
\sim \text{Hankel function}\sim\frac{1}{\sqrt{-x^2}},\text{propagation},x^2<0 & \text{inside light cone}\\
\sim K_1\text{ function}\sim e^{-\text{const.}\sqrt{x^2}},\text{not propagation}, x^2>0 & \text{outside light cone .}
\end{cases}
\end{align*}
Here is the result for the causal Green's function $\Delta_c(x)$:
\begin{align*}
\Delta_c(x) &= \frac{1}{4\pi}\delta(x^2)+\Theta(x^2)\frac{i\mu}{4\pi^2\sqrt{x^2}} K_1(\mu\sqrt{x^2}) - \Theta(-x^2)\frac{\mu}{8\pi\sqrt{-x^2}} H_1^{(2)}(\mu\sqrt{-x^2})\\
&= \frac{1}{4\pi}\delta(x^2) - \Theta(-x^2)\frac{\mu}{8\pi\sqrt{-x^2}}\left[ J_1(\mu\sqrt{-x^2}) -i N_1(\mu\sqrt{-x^2})\right]\\
&+\Theta(x^2)\frac{i\mu}{4\pi\sqrt{x^2}}K_1(\mu\sqrt{x^2})\quad .
\end{align*}
For the Pauli-Jordan function \eqref{eq:PJCFDef2.1} we use, from the detailed calculations in the appendix ($\Delta(x) := -2\bar{\Delta}(x)\epsilon(x)$), the result
\eqref{eq:AppDeltaBarCompact}:
\begin{align*}
\bar{\Delta}(x) &= \frac{1}{4\pi}\delta(x^2) -\frac{\mu^2}{8\pi}\mathfrak{Re}\left[\frac{H_1^{(1)}(\mu\sqrt{-x^2}) }{\mu\sqrt{-x^2}}\right]\\
\mathfrak{Re}\left[\frac{H_1^{(1)}(\mu \sqrt{-x^2} ) }{\mu\sqrt{-x^2}}\right] &= \begin{cases}
\frac{J_1(\mu \sqrt{-x^2} )}{\mu \sqrt{-x^2}} & x^2 <0\\
0 & x^2 >0\quad .
\end{cases}
\end{align*}
$\bar{\Delta}(x)$ and therefore $\Delta(x)$ vanishes if $x^2>0$ (space-like distance), which is how we constructed the Pauli-Jordan commutator function. Finally,
\begin{equation*}
\Delta(x) = -\frac{1}{2\pi}\epsilon(x_0)\delta(x^2)+\frac{\mu}{4\pi\sqrt{-x^2}}\epsilon(x_0)\Theta(-x^2)J_1(\mu\sqrt{-x^2})\quad .
\end{equation*}
Unlike the situation for $\bar{\Delta}(x)$ there is no discontinuity at $x^2=0$ for $\Delta^{(1)}(x)$ and therefore $\Delta^{(1)}$ does not vanish for space-like distances($x^2>0$).
To show this we refer to the appendix:
\begin{align*}
\Delta^{(1)}(x) &= \frac{1}{\pi}\mathcal{P}\hspace*{-0.25ex}\int\Delta(x-\epsilon\tau)\frac{1}{\tau}d\tau\quad .
\intertext{Here we insert}
\Delta(x) &= -\frac{i}{(2\pi)^3}\int e^{ik\cdot x} \delta(k^2+\mu^2)\epsilon(k_0)(dk)\quad ,
\intertext{such that}
\Delta^{(1)}(x) &= -\frac{i}{\pi}\frac{1}{(2\pi)^3}\int(dk)\epsilon(k)
\underset{{i\pi\epsilon(k)}}{\underbrace{\left[\mathcal{P}\hspace*{-0.25ex}\int\frac{1}{\tau}e^{-ik\cdot\epsilon\tau}d\tau\right] }} e^{ik\cdot x}\delta(k^2+\mu^2)\quad ,
\intertext{which yields:}
\Delta^{(1)}(x) &= \frac{1}{(2\pi)^3}\int e^{ik\cdot x}\delta(k^2+\mu^2)(dk)\quad ,
\end{align*}
which is nothing but the momentum representation of $\Delta^{(1)}(x)$.
In \eqref{eq:AppDelta1FinalRepresentaion} we present the final result for $\Delta^{(1)}(x)$:
\begin{equation}
\Delta^{(1)}(x) = \frac{\mu}{4\pi}\frac{1}{\sqrt{-x^2}}\Theta(-x^2)N_1(\mu\sqrt{-x^2})+\frac{\mu}{2\pi^2}\frac{1}{\sqrt{x^2}}\Theta(x^2)K_1(\mu\sqrt{x^2})\quad .\label{eq:Delta1FinalResultShortForm}
\end{equation}
For a space-like distance and equal time we have $(x-x')\rightarrow (\vec{x}-\vec{x}')=\vec{z},|\vec{z}|=:r$. Then the last term in \eqref{eq:Delta1FinalResultShortForm} allows us to study
its behavior for large $r$:
\begin{align*}
\Delta^{(1)}(\vec{z},0) &= \frac{\mu}{2\pi^2 r}K_1(\mu r) = 2\frac{\mu}{4\pi^2r}K_1(\mu r)\\
&\overset{r\rightarrow\infty}{\sim} \frac{2\sqrt{\mu}}{(2\pi r)^{\frac{3}{2}} }e^{-\mu r}\\
&\overset{x'=0}{=} \frac{2\sqrt{\mu} }{ \left(2\pi\sqrt{x^2}\right)^{ \frac{3}{2} } } e^{-\mu\sqrt{x^2} },\quad x^2>0\quad ,
\end{align*}
which means that $\Delta^{(1)}$ drops as $e^{-\mu r}$ for $\mu\sqrt{x^2}\gg 1$, where $\tfrac{1}{\mu}\sim$ Compton wavelength.
\section{Pair Production of Charged Scalar Particles}
In the last chapter of their paper\cite{Pauli1934} Pauli and Weisskopf calculate the probability for a scalar charged pair production with photons in the presence of a Coulomb
field. The theory of particle production ($e^+,e^-$) by $\gamma$ rays in the Coulomb field of a nucleus had already been calculated by Bethe and Heitler with the aid of
Dirac's hole theory\cite{Bethe1934}. According to Pauli: "The most interesting part of our theory is that the energy (of the produced particles) is always positive automatically (namely,
without using a superfluous hypothesis such as hole theory)."
We will refrain from calculating any processes of scalar QED. The textbook literature offers abundant examples, including scattering as well as bound-state problems. We will instead finish
this article by working out the problem of pair production in scalar QED in presence of an external, constant, electromagnetic field. Research in producing spinor as well as
scalar charged particles in external fields is, at the moment, of worldwide interest.
Let us define our problem more explicitly. It is known that the Lagrangian of a free electromagnetic field is given by $\mathcal{L}_0 = -\tfrac{1}{4} F_{\mu\nu}^2(x)$.
We want to find out how this Lagrangian is modified if the quantum vacuum is taken into account. More specifically, we want to compute a charged scalar particle loop to all orders in a constant,
prescribed, external field and determine how it affects the free electromagnetic Lagrangian.
In other words, we want to calculate the effective Lagrangian for scalar QED. I will present as many details as possible. But for a thorough understanding of the whole problem,
scalar as well as spinor particles in the loop, I would like to invite the reader to consult the Springer Lecture Notes \cite{Dittrich1985} or the monography \cite{Dittrich2000}.
Here are some basic facts:
\begin{itemize}
\item Hamiltonian: $H=\Pi^2$, $E$ field and $H$ field in $x_3$ direction
\item Gauge, background field $A_\mu(x) = \left( -\frac{1}{2}Hx_2,\frac{1}{2}Hx_1, -Ex^0;0\right)$
\item Non-zero field components $F_{12} = -F_{21} = H, F_3^0 = E$
\end{itemize}
\begin{equation}
\Pi_\mu = p_\mu - eA_\mu : \begin{cases}
\left. \begin{array}{c}
\Pi_1 = p_1+\frac{e}{2}Hx_2\\
\Pi_2 = p_2 - \frac{e}{2}Hx_1
\end{array} \right\rbrace & \left[\Pi_1,\Pi_2\right] = ieH\\
\Pi_3 = p_3 +eEx^0 & \\
\Pi^0 = p^0 &
\end{cases}\quad .
\end{equation}
In general $[\Pi_\mu,\Pi_\nu] = ieF_{\mu\nu}$.
\begin{align*}
\Pi^2 = (p-eA)^2 &= \left(\Pi_1^2+\Pi_2^2\right)+\left(\Pi_3^2-\Pi_0^2\right) = \Pi_\perp^2 + \Pi_\parallel^2,\quad [\Pi_\perp,\Pi_\parallel] = 0\\
&= \left(p_1+\frac{e}{2} Hx_2\right)^2+\left(p_2-\frac{e}{2} Hx_1\right)^2 + \left(p_3+eEx^0\right)^2-{p^0}^2\ .
\end{align*}
Now, without derivation - it can be found in our Lecture Notes\cite{Dittrich1985} - the effective Lagrangian for a complex scalar K.G.F. field is given by
\begin{equation}
\mathcal{L}_{eff} = \mathcal{L}_0 + \mathcal{L}_0^{(1)}\quad ,
\end{equation}
where $\mathcal{L}_0^{(1)}$ denotes the one-loop (effective) Lagrangian
\begin{equation}
\mathcal{L}_0^{(1)} = -i\int_0^\infty \frac{1}{s}e^{-im^2s}\langle x \vert e^{-is\Pi^2}\vert x\rangle ds\quad .
\end{equation}
Hence we need to work out the following diagonal element($ x^0=t$):
\begin{equation*}
\langle x \vert e^{-is\Pi^2}\vert x\rangle = \langle x \vert e^{-is\Pi^2_\perp} e^{-is\Pi_\parallel^2}\vert x\rangle =
\langle x_1 x_2 \vert e^{-is\Pi^2_\perp}\vert x_1x_2\rangle\langle x_3 x^0\vert e^{-is\Pi_\parallel^2}\vert x_3x^0 \rangle\quad .
\end{equation*}
The two transition amplitudes we need to calculate are
\begin{align*}
\int \langle x_1x_2\vert p_1p_2\rangle\langle p_1p_2\vert &e^{-is\left(\Pi_1^2+\Pi_2^2\right)} \vert p_1'p_2'\rangle\langle p_1'p_2'\vert x_1x_2\rangle dp_1dp_2dp_1'dp_2'\\
\int \langle x_3t\vert p_3\omega\rangle\langle p_3\omega\vert &e^{-is\left(\Pi_3^2-\Pi_0^2\right)} \vert p_3'\omega'\rangle\langle p_3'\omega'\vert x_3t\rangle dp_3d\omega dp_3'd\omega'\quad .
\end{align*}
Here we have to insert
\begin{align*}
\langle x_1x_2\vert p_1p_2\rangle &= \frac{1}{2\pi}e^{i(p_1x_1+p_2x_2)} & \langle p_1'p_2'\vert x_1x_2\rangle &= \frac{1}{2\pi}e^{-i(p_1'x_1+p_2'x_2)}\\
\langle x_3t\vert p_3\omega\rangle &= \frac{1}{2\pi}e^{i(p_3x_3-\omega t)} & \langle p_3'\omega'\vert x_3 t\rangle &= \frac{1}{2\pi}e^{-i(p_3'x_3-\omega't)}\quad .
\end{align*}
With a few steps in between, the result for the $E$ part is
\begin{align*}
\frac{1}{(2\pi)^2}\int e^{i(p_3-p_3')x_3}e^{-i(\omega-\omega')t} \langle p_3\omega\vert &e^{-is\left[ (eE)^2(x^0-\frac{p_3}{eE})^2-{p^0}^2\right]}\vert p_3'\omega'\rangle dp_3dp_3'd\omega d\omega'\\
&= \frac{1}{(2\pi)^2}\frac{\pi}{s} \frac{esE}{\sinh(esE)} = \frac{1}{4\pi s} \frac{esE}{\sinh(esE)}\\
&= \langle x_3x^0\vert e^{-is\Pi_\parallel^2}\vert x_3x^0\rangle\quad .
\end{align*}
For the $H$ term we take from elementary quantum mechanics the spectrum for $H=\left(\vec{p}-e\vec{A}\right)_\perp^2 : E_n =(2n+1)eH$ so that
(the factor $\tfrac{eH}{2\pi}$ takes into account the degeneracy per unit area of the Landau levels):
\begin{align*}
\langle x_1x_2\vert e^{-is\Pi_\perp^2} \vert x_1x_2\rangle &= \frac{eH}{2\pi} \sum_n e^{-(2n+1)(ieHs)} = \frac{1}{4\pi}\frac{eH}{\sinh(ieHs)}\\
&= \frac{1}{4\pi s}\frac{eHs}{i\sin(eHs)}\quad .
\end{align*}
Hence we obtain the diagonal element:
\begin{equation*}
\langle x \vert e^{-is\Pi^2}\vert x\rangle = \left( \frac{1}{4\pi s i}\frac{eHs}{\sin(eHs)}\right)\left( \frac{1}{4\pi s}\frac{eEs}{\sinh(eEs)}\right) =
\frac{1}{16\pi^2}\frac{1}{is^2}\left[ \frac{eHs}{\sin(eHs)}\frac{eEs}{\sinh(eEs)}\right]
\end{equation*}
(c.f. also (5.15) in \cite{Dittrich1985}).
With this result we can write:
\begin{equation}
i\mathcal{L}_0^{(1)} = \frac{1}{16\pi^2}\frac{1}{i}\int_0^\infty \frac{1}{s^3}e^{-im^2s} \frac{eHs}{\sin(eHs)}\frac{eEs}{\sinh(eEs)}ds\quad .
\end{equation}
Our final result, with the necessary subtraction terms to produce a finite answer, is given by the Heisenberg-Euler(H.E.) Lagrangian for scalar electrodynamics $\mathcal{L}_{HE} :=
\mathcal{L}_0 + \mathcal{L}_0^{(1)}$ with
\begin{equation}
\mathcal{L}_0^{(1)}(H,E) = \frac{1}{16\pi^2}\int_0^\infty \frac{ds}{s^3}e^{-m^2s}\left[ \frac{eHs}{\sinh(eHs)}\frac{eEs}{\sin(eEs)} - 1 - \frac{e^2s^2}{6}(E^2-H^2)\right]\ .
\end{equation}
Let us limit ourselves to the case of a pure applied constant $H$ field. Then there are a number of ways to explicitly compute the $s$-integral, e.g., with dimensional regularization, $\zeta$-
function regularization, etc. They all agree with the answer
\begin{equation}
\mathcal{L}_0^{(1)}(H) = \frac{1}{64\pi^2}\left\lbrace \left[ 2m^4-\frac{2}{3}\left(eH\right)^2\right]\left(1+\ln\frac{m^2}{2eH}\right) -3m^4-(4eH)^2\zeta'\left(-1,\frac{m^2+eH}{2eH}\right)\right\rbrace .
\end{equation}
The conversion from a pure magnetic field to a pure electric field takes place by substitution $B\rightarrow -iE$. Like in spinor electrodynamics it is possible to compute
via $2\mathfrak{Im}\left[\mathcal{L}_0^{(1)}(E)\right]$ the pair creation probability for spin 0 charged scalar particles in an external constant $E$ field\cite{Dittrich2014}.
Here is the final formula:
\begin{equation}
2 \mathfrak{Im}\left[\mathcal{L}^{(1)}(E)\right] = (2s+1)\frac{(eE)^2}{(2\pi)^3}\sum_{n=1}^\infty \frac{(\pm 1)^{n+1}}{n^2}e^{-n\frac{\pi m^2}{eE}}\quad ,
\end{equation}
where the upper and lower signs correspond to $s=\tfrac{1}{2}$ and $s=0$, respectively.
It should be emphasized that the H.E. Lagrangian yields a totally non-perturbative theory for low-energy photons. It encodes a tremendous amount of physics,
like photon-photon scattering, vacuum polarization, determination of the $\beta$ function in scalar as well as spinor QED, etc. All together the Pauli-Weisskopf theory
is as far-reaching for spinless particles as the Dirac theory is for spin-$1/2$ particles - without ever mentioning Dirac's hole theory.
|
\section{Introduction}
The mathematical modeling and simulation of pedestrian dynamics,
such as large human crowds in public space or buildings, has
become a topic of high practical relevance. The complex behavior
of these large crowds poses significant challenges on the
modeling, analytic and simulation level. These aspects initiated a
lot of research in the mathematical community within the last
years, which we briefly outline below. Mathematical modeling
approaches for pedestrian dynamics can be roughly grouped into the
following categories:
\begin{enumerate}
\item \textit{Microscopic models} such as the social force model\cite{helbing1995,helbing2000,treuille2006} or
cellular automata approaches\cite{burstedde2001}.
\item \textit{Fluid dynamic approaches}\cite{colombo2012,moussaid2012,appert2011} and related \textit{macroscopic models}, see for example the popular Hughes model\cite{hughes2002,difrancesco2011,goatin2013,burger2014,VenutiBruno}.
\item \textit{Kinetic models} \cite{bellomo2013,DARPT2013} which uses ideas from gas kinetics to models interactions between individuals via so-called collisions.
\item In \textit{optimal control}\cite{hoogendoorn2004} and \textit{mean field game approaches}\cite{lachapelle2011,dogbe2010} pedestrians act as rational individuals, which adjust
their velocity optimal to a specific cost.
\item \textit{Multiscale models} coupling between different scales to describe for example crowd leader dynamics \cite{CPT2011,BPT2012}.
\end{enumerate}
A detailed survey on crowd modelling can e.g. be found in Bellomo
and Dobge\cite{BellomoDogbe2011}. Several aspects are considered
to be important in the mathematical modeling to capture the
complex behavior in a correct way. For example repulsive forces
when getting too close to other individuals or obstacles play an
important role in the dynamics. Another popular assumption is the
fact that individuals act rationally and try to make the optimal
decision based on their actual knowledge level. Partial knowledge
of the overall pedestrian density or the domain is another
important factor which should be taken into account in the
modelling. While these nonlocal effects can be implemented quite
intuitively on the microscopic level, their translation for
macroscopic models is not straightforward. Most macroscopic
nonlocal models are based on the continuity equation for the
pedestrian density, where the nonlocal effects correspond to the
deviation of the crowd from its preferred
direction\cite{colombo2012,colombo2012-2,crippa2013}. This
deviation is determined by the average density felt by the
pedestrians and modelled via a convolution operator acting on the
velocity. The development of numerical schemes for conservation
laws with nonlocal effects gained substantial interest in the last
years. This was, among other factors, also initiated by the
development of nonlocal models in traffic
flow \cite{ACT2014,BG2014}. \\
\noindent The original model of Hughes\cite{hughes2002} describes
fast exit and evacuation scenarios, where a group of people wants
to leave a domain $\Omega \subset \mathbb{R}^2$ with one or
several exits/doors and/or obstacles as fast as possible. The
driving force towards the exit is the gradient of a potential
$\phi = \phi(x,t)$, $x \in \Omega,~t > 0$. This potential
corresponds to the expected travel time to manoeuvrer through the
present pedestrian density towards an exit. Hughes assumed that
the global distribution of pedestrians is known to every
individual, an assumption not generally satisfied in real world
applications.
\\
In this paper we present a generalization of the classical Hughes
model, which includes local vision via partial knowledge of the
pedestrian density. We discuss the proper modeling setup, the
implementation of suitable numerical schemes as well as their
computational complexity. Furthermore we compare how the reduced
perception of each pedestrian effects the overall "performance" of
the crowd in evacuation scenarios. Inevitably, one expects the
crowd to behave less efficient as less information is available.
Quantifying how localised vision influences performance and
decision making is a very interesting question in terms of
collective behaviour. Surprisingly, it will turn out that
evacuation times can even improve. The question we investigate is
therefore complementary to mean-field game approaches, where
pedestrians anticipate future crowds states and hence are
\emph{more} capable than in the original Hughes' model
\cite{lachapelle2011,dogbe2010,burger2014}.
\\
\noindent This paper is structured as follows: We start with a
review on the modeling and analytic results of the classical
Hughes model for pedestrian flow and its microscopic
interpretation in section \ref{s:hughes}. In section
\ref{s:mathmod} we present the local version of the Hughes model
on the micro- and macroscopic level. Section \ref{s:computmeth}
presents the numerical strategies for the microscopic and
macroscopic model. We compare the behavior and performance of the
models in Section \ref{s:numexp} and conclude with a discussion of
the proposed model in Section \ref{s:conclusions}.
\section{Hughes' model for pedestrian flow}\label{s:hughes}
\subsection{Original formulation and analytic results}
Let us start by presenting the original modeling assumptions and
the corresponding partial differential equation system of the
Hughes model for pedestrian flow. Hughes considered an exit
scenario, in which a crowd modelled by a macroscopic density $\rho
= \rho(x,t)$ wants to leave a domain as fast as possible. The
nonlinear PDE system for $\rho$ and the potential $\phi =
\phi(x,t)$ on the domain $\Omega \subset \mathbb{R}^2$ reads as:
\begin{subequations}\label{e:hughes}
\begin{align}
\frac{\partial \rho}{\partial t}-\mathrm{div}(\rho f(\rho)^2 \nabla \phi) &= 0 \label{e:hughescont}\\
\norm{\nabla \phi} &= \frac{1}{f(\rho)}. \label{e:hugheseikonal}
\end{align}
\end{subequations}
The first equation describes the evolution of $\rho$ in time,
driven by the gradient of $\phi$ and weighted by a nonlinear
mobility $f = f(\rho)$. This mobility includes saturation effects,
i.e. degenerate behaviour when approaching a given maximum density
$\rho_{\max} \in \mathbb{R}^+$. Possible choices are $f(\rho) =
\rho-\rho_{\max}$ or $f(\rho) = (\rho-\rho_{\max})^2$ amongst
others. The former is inherited from the
Lighthill-Whitham-Richards model for one-dimensional traffic flow
\cite{LW1955,R1956}.
\\
The potential $\phi$ corresponds to the weighted shortest distance
to an exit in the following sense: Solving the eikonal equation
\eqref{e:hugheseikonal} determines the optimal path $\nabla \phi$
minimising the expected travel time throughout the crowd towards
an exit. This cost is measured as the inverse of $f(\rho)$, hence
the cost of walking through dense regions is high. Equation
\eqref{e:hugheseikonal} is also a stationary
Hamilton-Jacobi-Bellman equation, and the optimal path property of
$\nabla\phi$ can be rigorously derived \cite{BC,Holm}. The fact
that the potential $\phi$ solely determines the direction of the
flow can be easily seen as $f^2(\rho)\nabla\phi =
f(\rho)\nabla\phi / \norm{\nabla\phi}$ using
\eqref{e:hugheseikonal}.
\\
\noindent Hughes model \eqref{e:hughes} is supplemented with
different boundary conditions for the walls and the exits. We
assume that the boundary is divided into two parts: either
impenetrable walls $\partial\Omega_{\text{wall}} \subset \partial \Omega$ or exits/doors
$\partial\Omega_{\text{exit}} \subset \partial \Omega$, with $\partial\Omega_{\text{wall}} \cap \partial\Omega_{\text{exit}} =
\emptyset$. Typical conditions for the density $\rho$ in
\eqref{e:hughes} are zero flux boundary conditions at $\partial\Omega_{\text{wall}}$,
which are either automatically satisfied as $\nabla\phi \cdot \vec
n =0$ or artificially enforced. The flux at $\partial\Omega_{\text{exit}}$ has to be
defined according to the arriving density and our choices are
discussed in Section \ref{s:mathmod}. The boundary conditions of
\eqref{e:hugheseikonal} are set as $\phi(x,t) = 0$ for all $ x \in
\partial\Omega_{\text{exit}}$.
\noindent There has been a lot of recent mathematical research on
the classical Hughes model
\cite{difrancesco2011,goatin2013,amadori2012,elkhatib2013}. Up to
the authors knowledge all analytic results are restricted to 1D
only, which is caused by the low regularity of the potential
$\phi$. This low regularity, i.e. $\phi \in C^{0,1}$, results in
the formation of shocks and rarefaction waves in the conservation
law. It is caused particularly by the existence of \emph{sonic
points}, which are hypersurfaces in space, where costs towards two
or more exists coincide, and therefore $\nabla\phi$ does not exist
and the orientational field is discontinuous. In spatial dimension
one the system can be reduced to the conservation law with a
discontinuous flux function. In this case it is possible to solve
the corresponding Riemann problem \cite{difrancesco2011}, which
also serves as a basis for different numerical schemes
\cite{goatin2013,BG2014}.
\subsection{Microscopic interpretation}
Hughes motivated system \eqref{e:hughes} on the macroscopic level
only. Recently Burger et al. \cite{burger2014} were able to give a
microscopic interpretation of \eqref{e:hughes}, which will serve
as a basis for our local particle model. Microscopic models based
on Hughes' modelling assumptions are also used in the field of
computer vision \cite{treuille2006}.
\noindent Let us consider $N$ particles with position $X^j =
X^j(t)$ and velocity $V^j = V^j(t)$, $j=1,\ldots N$. Then the
empirical density $\rho^N = \rho^N(t)$ is given by
\begin{align}\label{e:rhon}
\rho^N(t) = \frac{1}{N} \sum_{j=1}^N \delta(x-X^j(t)).
\end{align}
Furthermore we introduce its smoothed approximation $\rho^N_g = \rho^N_g(t)$, given by
\begin{align*
\rho^N_g(x,t) = (\rho^N * g)(x,t) = \frac{1}{N} \sum_{j=1}^N g(x-X^j(t)),
\end{align*}
where the function $g = g(x)$ corresponds to a sufficiently smooth
positive kernel. The walking cost is given by the sum of a
weighted kinetic energy and the exit time, defined as $T_{exit} =
\sup\lbrace t>0 \mid x \in \Omega \rbrace$. Then the problem reads
as:
\begin{align}\label{e:minC}
\begin{split}
&C(X;\rho(t)) = \min_{(X,V)} \frac{1}2\int_t^{T+t} \frac{\norm{V(s)}^2}{f^2(\rho^N_g(\xi(s;t),t))} ds + \frac{1}2 T_{exit}(X,V),\\
&\text{ subject to } ~~\frac{d \xi}{ds}=V(s) \text{ and }
\xi(0)=X(t).
\end{split}
\end{align}
Hence the optimal trajectory is determined by 'freezing' the
empirical density $\rho^N = \rho^N(t)$, in other words it
corresponds to extrapolating the empirical density $\rho^N$
in time when looking for the optimal trajectory.
\\
Burger et al. were able to show that Hughes' model can be
formally derived from the optimality conditions of \eqref{e:minC}
and letting $T \rightarrow 0$ (corresponding to the long-time
behavior of the corresponding adjoint Hamilton-Jacobi equation).
\\
We will use this microscopic interpretation to propose a numerical
approximation by a particle method in Section 4 of Hughes type
models. In fact, \eqref{e:hughescont} is seen as a continuity
equation with velocity field $v(x,t)=-f(\rho)^2 \nabla \phi$
driven by \eqref{e:hugheseikonal}, and thus particles in
\eqref{e:rhon} are advected by the velocity field $v$, e.g.
$$
\frac{d X^j}{dt}=v(X^j(t),t)\,,\qquad j=1,\dots N\,.
$$
\section{A localized smooth Hughes-type model for pedestrian flow}\label{s:mathmod}
\noindent The Hughes model \eqref{e:hughes} assumes that at any
time $t>0$ the global distribution of all other individuals $\rho(x,t)$ is known to every
pedestrian. Therefore she chooses her optimal walking direction $\nabla
\phi$ in order to minimise its expected travel time/costs. Here,
all walking costs are based on the current density, which means
that pedestrians do not anticipate future dynamics of the crowd.
Instead they are capable to react to changes in the global density
ad-hoc as the path optimisation is repeated continuously in time.
In a mean-field game type model, the capabilities of pedestrians
would increase, as the planning decision of all agents can be
correctly predicted into the future. We follow an opposite
approach and reduce the capabilities of pedestrians, to obtain a
more realistic model.
\\
\noindent The assumption of continuous and complete perception of
global density information at current time is highly questionable
in practical situations. Limited vision cones and restricted
perception of global information comes through obstacles (walls,
buildings), physical distance, visual orientation or the inability
to see through a very dense crowd. Some effects of local vision
on the behaviour of crowds are obvious: in an evacuation scenario
with two exit corridors, which cannot be seen from each other,
pedestrians caught in a jam in front of one exit will not be able
to see whether the other exit is free or also jammed.
\noindent These considerations motivated a new version of
Hughes-type pedestrian dynamics based on localised perception of
information, which we introduce in this section. The decision of
each pedestrian is based on the perceived local density available
in a limited domain, which can be e.g. interpreted as a vision
cone. Furthermore
a local interaction mechanism between individuals as well as a smoothening kernel
on the velocity field (to prevent unrealistic high frequency
oscillations in the direction of motion) are incorporated. We
begin with the detailed introduction of the macroscopic model and
discuss the microscopic analogue thereafter.
\subsection{Macroscopic equations}
The starting point of our model is the assumption that pedestrians
still perform the same path-optimisation selection as in the
Hughes' model, while the crowd state they act upon subjectively
depends on their position and the amount of information they are
able to perceive. Let $y\in\mathbb{R}^2$ be an auxiliary variable and
$\phi(x,y):\mathbb{R}^4\rightarrow\mathbb{R}$ a parameterised potential, such
that
y\mapsto \phi(x_0,y)
$
denotes the cost potential calculated by pedestrians located at
$x_0\in\Omega$. To model space-dependent perception of
information, suppose that for every $x$ the domain $\Omega$
decomposes into a visible subdomain $V_x\ni x$ and a hidden or
invisible part $H_x = \Omega \backslash V_x$. We propose the
following mechanism of restricted vision: If an area is visible
its density is known and priced accordingly in the path
optimisation. If however an area is not visible, its density is
thought to be a constant $\rho_H \in \mathbb{R}^+_0$, which we assume to
be uniform among all pedestrians. Exemplarily, setting $\rho_H=0$
implies that pedestrians assume that not visible areas to them are
empty, hence they will have a strong incentive to explore unseen
parts of the domain. On the contrary, pedestrians will avoid
invisible areas when $\rho_H\approx \rho_{\max}$, as they assume
high costs. An eikonal equation in $\Omega$ is hence solved in the
auxiliary variable $y$ for every point $x$ as
\begin{equation}
\norm{\nabla_y \phi(x,y )} = \displaystyle{\begin{cases}
\frac{1}{f(\rho(y,t))} & y\in V_x \\[1mm]
\frac{1}{f(\rho_H)} & y\in H_x
\end{cases}},
\label{model-local-eikonal}
\end{equation}
which gives the potential $\phi$ as function of two space
variables. Note, that this notion of local perception differs from
other recent work\cite{BG2014}, where a local average of the
density is used. Each pedestrian uses the cost potential at her
own position for the decision process. Computing $\nabla_y
\phi(x,x)$ hence would, after normalisation, give a new
orientational vector field to be used in the unchanged transport
equation. We however argue that it makes sense to include a notion
of conviction to the model, which has previously not been
considered. In order to do so, \eqref{model-local-eikonal} is
solved for every single exit. This results in the computation of
$M$ potentials $\phi_k=\phi_k(x,t)$, $k=1,\dots, M$, which allows
for cost comparison between exits, see Remark 3.2.
In regions of high density, decisions on the walking direction
towards any of $k=1,\dots,M$ exits $\partial\Omega_{E_k}$ cannot
arbitrarily deviate between neighbours. If a pedestrians prefers
to walk against the direction of a predominant local flow,
collision or friction losses in the movement will occur.
Especially on the macroscopic level, in which we take an aggregate
perspective, an incentive to change the flow cannot arise from one
point in space alone. As we do not model the granular level of
individual collision or friction, we propose the following
mechanism:
\begin{enumerate}
\item Each pedestrian carries an individual conviction strength $u(x)$
measuring its preference of its chosen exit over all others.
\item There exists a local consensus process within the crowd, which results
in the adjustment of the individual walking direction according to the
predominant direction around them.
\end{enumerate}
Hence, pedestrians adjust their own direction in order to prevail
the flow rather than obstructing it. This can be seen as either a
cognitive decision rule or a forced physical restriction. For a
compactly supported interaction kernel
$\mathcal{K}:\mathbb{R}^2\rightarrow \mathbb{R}$, we define the final walking
direction $\varphi(x)$ at any point $x\in\Omega$ as
\begin{equation}
\varphi = \frac{\rho u \star \mathcal{K}}{\rho\star\mathcal{K}},
\end{equation}
where the conviction $u(x)$ is given as
\begin{equation}
u = \frac{\nabla_y \phi_{k^\text{opt}}}{\norm{\nabla_y
\phi_{k^\text{opt}}}}(\phi_{k^{\text{opt}+1}}-\phi_{k^\text{opt}}),
\end{equation}
obtained by comparing the cost potentials $\phi_k$, $k=1,\dots,
M$, associated to each of the exits:
\begin{gather}
k^{\text{opt}}(x) = \operatorname{argmin}\limits_k \phi_k(x,x), \\
k^{\text{opt+1}}(x) = \operatorname{argmin}\limits_{k\neq k^{\text{opt}}} \phi_k(x,x).
\end{gather}
Discontinuities in the velocity field due to the heterogeneity of
decision making amongst pedestrians are hence partially
compensated. To further smooth the model, we relax the strict
restriction of $\norm{\nabla\varphi}=1$ of Hughes' model and
replace the normalisation operator with a smooth approximation
$\mathcal{P}:\mathbb{R}^2\rightarrow \mathbb{R}^2$ defined as:
\begin{equation}
\mathcal{P}[x] := \begin{cases}
\frac{x}{\norm{x}} & \norm{x}>\ell, \\
\sin\left(\frac{\pi}{2 \arctan(k\ell)}\arctan(k\norm{x})\right)\frac{x}{\norm{x}} & 0<\norm{x}\leq\ell, \\
{0} & x=0 ,
\end{cases},
\label{e-relaxation}
\end{equation}
for some parameters $k,\ell>0.$ We stress that this is not a
technicality, as we here allow pedestrians to \emph{stop when
being undecided}. This is highly desirable from the modelling
point of view, though on the other hand the modulus of the flux
now is not a function of density alone, as one can see below.
\\
Next we discuss the boundary conditions for the eikonal equations.
Since we treat each exit separately, we set
$\left.\phi_k\right\vert_{\partial \Omega_{E_k}} = 0$ in the
computation of $\phi_k$. No boundary conditions are imposed on the
rest of the boundary $\partial\Omega_{\text{wall}} $.
\noindent Near-wall and near-obstacle effects have a strong
influence on the dynamics on constrained macroscopic evolutions.
We propose that pedestrians take into account walls and obstacles
in their computation of optimal paths. Hence it is natural to
include these effects as an additional fixed cost $W(x)$ on the
right-hand side of the eikonal equation
\eqref{model-local-eikonal}. We introduce a smooth layer profile
$\chi_w(x)\in[0,1]$, which identifies areas close to walls but
smoothly vanishes elsewhere and around exits to allow outflow. A
typical choice of $\chi_w$ is illustrated in Figure
\ref{fig-layerprofile}. For the sake of simplicity, we set
\begin{equation}
W(x) = \frac{\chi_w(x)}{f(\rho_{\max}-\epsilon)},
\label{e:wall}
\end{equation}
hence areas close to walls are penalised similar to high density areas.
\begin{figure}
\centering \psxinput{\input{draw_layerprofile}}{.75}
\caption{Illustration of layer profile $\chi_w$: wall and obstacle
repulsion are embedded to the model with a fixed cost $W$ defined
in terms of the layer profile $\chi_w(x)\in[0,1]$, see
\eqref{e:wall}, which indicates proximity to walls less than a
width $w$, but vanishes away from obstacles and near exits to
allow a proper outflow of pedestrians. $\chi_w$ equals one at walls and is the linear function given by vertex values within triangles $T_1$, $T_2$ near exits.}
\label{fig-layerprofile}
\end{figure}
Finally all terms are coupled to the continuity equation with
velocity field $v(x,t)= -f(\rho) \mathcal{P}[\nabla \varphi]$ as
in the original model. At exits, we prescribe a maximum outflow,
given by $v(\xi,t)=-f(\rho) \vec n $ for all $ \xi\in\partial\Omega_{\text{exit}}$.
Taking all these considerations into account the full macroscopic
model reads as:
\begin{subequations}
\begin{gather}
\partial_t \rho(x,t) + \nabla_x \cdot \left(- f(\rho(x,t)) \mathcal{P}[\nabla \varphi(x,t)] \rho(x,t) \right) =0 \\
\varphi(x,t) = \frac{(\rho u \star \mathcal{K})(x,t)}{(\rho\star\mathcal{K})(x,t)} \label{e:localinteraction}\\
u(x,t) = \frac{\nabla_y \phi_{k^\text{opt}}(x,x,t)}{\norm{\nabla_y \phi_{k^\text{opt}}(x,x,t)}} (\phi_{k^{\text{opt}+1}}(x,x,t)-\phi_{k^\text{opt}}(x,x,t)) \label{e:conviction}\\
k^{\text{opt}}(x,t) = \operatorname{argmin}\limits_k \phi_k(x,x,t) \\
k^{\text{opt+1}}(x,t) = \operatorname{argmin}\limits_{k\neq k^{\text{opt}}} \phi_k(x,x,t)\label{e:secondbest} \\
\norm{\nabla_y \phi_k(x,y,t )} = \displaystyle{\begin{cases}
\frac{1}{f(\rho(y,t)} + W(y) & y\in V_x \\[1mm]
\frac{1}{f(\rho_H)} & y\in H_x \\
\end{cases}}\label{e:localeikonal}\\
\text{s.t.:} \,\left.\phi_k\right\vert_{\partial \Omega_{E_k}} = 0 , k=1,\dots,M \, , \forall t \\
\text{s.t.:} \,\mathcal{P}[\varphi(\xi,t)] = \vec n \,,\,\, \forall \xi\in\partial\Omega_{\text{exit}}\,,\,\,\, \label{e:maxoutflow}
\text{s.t.:} \,\rho(x,0)=\rho_0(x).
\end{gather}\label{e:localmodel}
\end{subequations}
We conclude the section with remarks on specific modeling assumptions.
\begin{remark}[vision cones]
We have set aside formal statements regarding assumptions on the visible set $V_x$, but clearly we think of at least regular, connected and closed sets.
A necessary condition is
\begin{equation}\label{e:Vinfo}
x \in V^\circ_x = \operatorname{int} V_x,
\end{equation}
which implies that every pedestrian perceives some information
from all directions. This restriction rules out e.g. angular
vision cones (see Degond et.al.\cite{DRMPT2013}) where pedestrians
do not see what is happening behind them. In our model,
\eqref{e:Vinfo} is necessary to exclude unrealistic situations
where the chosen walking direction points outside the visible
area. The inclusion of angular-dependent vision cones is certainly
possible, but would imply a velocity-dependency and lead towards a
second-order macroscopic model.
\end{remark}
\begin{remark}[conviction term]
The introduction of the conviction term $u(x)$ requires the
computation of exit costs $\phi_k$ via the eikonal equation for
individual exits, which appears to be a significant complication
of the model. However, it is worth noting that the mechanism is
almost identical to the original model. In equation
\eqref{e:hugheseikonal} the costs of walking towards any of the
$K$ exits are compared, but only the minimal costs are used. Here,
we simply store more information. This connection is also
illustrated by looking at the numerical schemes for solving the
eikonal equation (see also Section ?): If a Fast Sweeping Method
is used in e.g. a corridor with two exists, this essentially
corresponds to solving for each exit separately if the
minimization step is left out. If a Fast Marching Method is used,
the conviction is directly related to the sequence in which
vertices are promoted, with the least convinced vertex being
assigned a cost the latest.
\end{remark}
\begin{remark}[waiting behavior]
The relaxation $\norm{\mathcal{P}[x]}\leq 1$ implies that the
modulus of the flux can be less than $f(\rho)\rho$ when
pedestrians are undecided. This makes a rigorous analysis of the
model equations a difficult task, which is not tackled in this
work. The benefit of our formulation is that the problem of
discontinuous velocity fields at sonic points has disappeared.
Pedestrians at those hyper surfaces will not move unless the sonic
points move.
\end{remark}
\subsection{The one-dimensional case}
Consider a one-dimensional corridor $\Omega=[0,1]$ with two exits
and the uniform radial vision cone $V_x=[x-L/2,x+L/2]\cap \Omega$
of length $L$. Exit costs towards the left and right exit are
computed at $y\in V_x$ as
\begin{gather*}
\phi_L(x,y,t) = \int\limits_{z<y, z\in H_x} \frac{1}{f(\rho_H)} \mathrm{d} z + \int\limits_{z<y, z\in V_x} \frac{1}{f(\rho(z,t))} \mathrm{d} z, \\
\phi_R(x,y,t) = \int\limits_{z>y, z\in H_x} \frac{1}{f(\rho_H)} \mathrm{d} z + \int\limits_{z>y, z\in V_x} \frac{1}{f(\rho(z,t))} \mathrm{d} z , \\
u(x,t)=\phi_L(x,x,t)-\phi_R(x,x,t),
\end{gather*}
as illustrated in Fig. \ref{fig-1dillustration}.
\begin{figure}
\centering \psxinput{\input{1dpic1}}{.75} \caption{Illustration of
path optimisation mechanism in 1D: A pedestrian located at $x_0$
computes and compares the cost potential $\phi_L,\phi_R$ of left
vs.\@ right exit in a corridor $[0,1]$. Next to its own negligible
density, the present crowd consists of three blocks. Outside the
vision cone $V_x$, the evacuation costs grow linearly at constant
rate, as the local density is unknown. Within $V_x$, the slope of
the cost potential increases with the pedestrian density.
Preference is then given towards the exit with lower estimated
cost. The conviction towards this decision is given as the cost
benefit $|u(x_0)|=|\phi_L-\phi_R|$.} \label{fig-1dillustration}
\end{figure}
The cost potential $\phi$ is two-dimensional and $\partial_y
\phi(x,y)$ gives the preferred walking direction that a pedestrian
located at $x$ seeing $V_x$ assigns to $y\in[0,1]$. The walking
directions chosen prior to the consensus process are hence given
as $\partial_y \phi(x,x)$ along the diagonal of $[0,1]^2$. For
every fixed $x\in[0,1]$, there is a unique sonic point $z(x)$,
where $\phi_L(x,z(x))=\phi_R(x,z(x))$ and $\partial_y
\phi(x,z(x))$ does not exist. As illustrated in Fig.
\ref{f:1Ddirectionswitching}, the individually preferred walking
directions can switch multiple times between both exits, depending
on the current density and the vision cones. At switching points,
the preferred directions can point outwards (separation) or
inwards (collision) and only the weighted interaction process
\eqref{e:localinteraction}-\eqref{e:conviction} generates a smooth
velocity profile. In the Hughes' model, all vision cones are
identical and there is a single separation point.
\begin{figure}
\centering
\begin{tabular}{c c}
\psxinput{\input{draw_1d_phidiagram1}}{.9}&
\psxinput{\input{draw_1d_phidiagram2_hughes}}{.9}\\
(a) Localized model &
(b) Hughes' model
\end{tabular}
\caption{Illustration of turning decisions of a 1D population for
a given $\rho(x,t)$: For every point $x,$ we show the individual
sonic point $z(x)$, where the costs $
\phi_L(x,z(x))=\phi_R(x,z(x))$ coincide. The preferred walking
direction for a pedestrian at $x$ is found along the diagonal
$(x,x)$. If the sonic point is to the left, the pedestrians aim to
walk towards the right and vice-versa. No direction is preferred
and the conviction is zero if the curve of sonic points intersects
the diagonal. (a) Local vision cones: The preferred direction
alters and creates multiple points of separation and collision.
The resulting velocity field is obtained by the smoothening
interaction process
\eqref{e:localinteraction}-\eqref{e:conviction} . (b) Hughes'
model: All vision cones coincide, hence there is one identical
sonic point common to all pedestrians.}
\label{f:1Ddirectionswitching}
\end{figure}
\subsection{Microscopic interpretation}
We conclude this section by briefly commenting on the modelling of
local vision at the microscopic level. The microscopic
modification is straightforward and uses the same ideas as at the
macroscopic level. It corresponds to updating the position $X =
X(t)$ according to a potential which depends on local information
only. Its calculation is based on the same equations as in the
macroscopic model \eqref{e:localeikonal} but using the smoothed
empirical density $\rho^N_g$ instead of $\rho$. The position
update is based on equations
\eqref{e:localinteraction}-\eqref{e:secondbest}. Hence individuals
choose the path towards the exit with the lowest cost, but weigh
their decision according to the predominant direction chosen
around them. For further details on the implementation we refer to
Algorithm \ref{a:microsim} presented in Section
\ref{s:computmeth}.
\subsection{Analysis of the domains of dependence}
In this subsection, we will discuss some mathematical properties
of the solutions of the eikonal equations \eqref{e:localeikonal}.
From the construction of the model, the potential $\phi(x,y,t)$
has to be computed for every $x\in \Omega$ on the entire domain
$\Omega$, which counterbalances the idea of locality and increases
the computational cost considerably. We show here, that the
computation of the potential can actually be reduced to a subset
of $\Omega$, called the effective domain of dependence, for every
$x$. Only this subset, which contains $V_x$, is considered in the
individual local planning problem and corresponds to the reduction
of the computational cost.
The following proofs rely crucially on the optimal path property of
the characteristics associated to the eikonal equation
\eqref{e:localeikonal}. We recall\cite{Holm,BC} that by Fermat's
principle the characteristic paths associated to $\phi(x,y,t)$,
given by the solution of:
\begin{equation}\label{e:chareikonal}
\gamma_{x,t}^z(s) \subset \Omega: \gamma(0)=z, \dot{\gamma}(s) =-
\nabla \phi(x,\gamma(s),t) \,\, \mbox{ for all} \, s\geq 0\,,
\end{equation}
are the optimal paths for the cost defined as
$$
c(y,t)=\displaystyle{\begin{cases}
\frac{1}{f(\rho(y,t))} + W(y) & y\in V_x \\[1mm]
\frac{1}{f(\rho_H)} & y\in H_x \\
\end{cases}}\,.
$$
Moreover, the potential is the value function for that cost. Hence
it is decreasing along these paths and satisfies the optimality
condition
\begin{equation}\label{e:optimality}
\phi(x,\gamma_{x,t}^z(a),t)-\phi(x,\gamma_{x,t}^z(b),t)= \int_a^b
c(\gamma_{x,t}^z(s),t) \,ds\,, \mbox{ for all } 0\leq a<b\,,
\end{equation}
being zero at its corresponding exit $\partial
\Omega_{\text{exit}}$. Furthermore, the curves $\gamma_{x,t}^z$
are the optimal paths to achieve the exit, i.e., they verify the
following global optimality condition
\begin{equation}\label{e:optimality2}
+\phi(x,z,t)= \int_0^{T_z} c(\gamma_{x,t}^z(s),t) \,ds \leq
\int_0^{\tilde T_z} c(\tilde\gamma(s),t) \,ds \,,
\end{equation}
for all $\tilde\gamma$ curves joining $z$ to any point in the exit
$\partial \Omega_{\text{exit}}$, where $T_z$ is the optimal time
to achieve the exit for the point $z\in\Omega$ and $\tilde T_z$ is
the time to achieve the exit for the path $\tilde\gamma$.
\begin{lemma}
\label{lemma-Mreduction} Consider any fixed $V_x\subset\Omega$ and
that $f(\rho)>0$, $0\leq\rho<\rho_{\max}$. Let $\phi_H$ be the
global solution of the eikonal equation $\norm{\nabla \phi_H } =
1/f(\rho_H), \phi_H(x)=0 \text{ on } \partial
\Omega_{\text{exit}}$. Define the minimum of $\phi_H$ in $V_x$ as
\begin{equation*}
m_H:= \min_{z\in V_x} \phi_H(z)
\end{equation*}
and the corresponding superlevel set of $\phi_H$ as
\begin{equation*}
M_H := \{x\in \Omega: \phi_H(x) \geq m_\phi \}.
\end{equation*}
Then the problem of computing the local potential $\phi(x,y,t)=:
\tilde{\phi}(y)$ out of \eqref{e:localeikonal} on $\Omega$ reduces
to the following problem on $M_\phi$:
\begin{equation*}
\begin{cases} \norm{\nabla_y \tilde{\phi}(y) }= \frac{1}{f(\rho(y,t))}+W(y) &\text{ in } V_x \\
\norm{\nabla_y \tilde{\phi}(y) }= \frac{1}{f(\rho_H)} &\text{ in } M_H \backslash V_x \\
\tilde{\phi}(y) = m_H &\text{ on } \partial M_H \backslash
\partial\Omega_{\text{wall}}\,\, (\text{B.C. } )
\end{cases} \,.
\end{equation*}
\end{lemma}
\begin{proof}
If an exit is visible then $m_H=0, M_H=\Omega$ and the assertion
is trivial. If no exit is visible then by construction $V_x
\subset M_H$ and $\phi_H= m_H>0$ on $\partial M_H$. As the walking
costs are always positive, $c(y,t)>0$, we get $\phi(x,y,t)>m_H$
for all $y \in \operatorname{int} M_H$. On the other hand, any
point $z\in\Omega\backslash M_H$ satisfies $\phi(x,z,t)<m_H$ and
hence $\gamma_{x,t}^z(s)$ does not intersect $V_x$, otherwise the
cost should be larger at a middle point than initially
contradicting the optimality of the path $\gamma_{x,t}^z(s)$ in
\eqref{e:optimality}. Hence $\partial M_H$ is the maximal level
set consisting of points whose optimal paths do not cross $V_x$,
and therefore, $\phi(x,z,t)$ can be computed from
\eqref{e:localeikonal} with constant right-hand side outside
$M_H$.
\end{proof}
\begin{definition}
Consider a fixed visibility area $V_x$. For a $z\in\Omega$, denote the
\emph{default optimal path} $\gamma_H^z$ as the parameterised
curve associated to a gradient walk along $\phi_H$ starting in
$z$, that is
\begin{equation*}
\gamma_H^z(s) \subset \Omega: \gamma(0)=z, \dot{\gamma}(s) =-
\nabla \phi_H(\gamma(s)) \,\, \forall \, s\geq 0.
\end{equation*}
Next, define the \emph{characteristics' shadow} $V^\#$ as the set of all points, whose default optimal paths crosses the visibility area, hence
\begin{equation*}
V^\# := \{ z\in\Omega: \gamma_H^z \cap \operatorname{int} V_x \neq
\emptyset \}.
\end{equation*}
\end{definition}
Note, that $V^\# \subset M_H$ since any default optimal path
outside of $M_H$ cannot intersect with $V_x$ as proven in the
previous lemma.
\begin{lemma}\label{lemma-Vreduction}
Consider any fixed $V_x\subset\Omega$ and assume that $f(\rho)>0$ is
increasing in $0\leq\rho<\rho_{\max}$, with $\rho_H=0$, then the
problem of computing the local potential $\tilde{\phi}(y)$ out of
\eqref{model-local-eikonal} further reduces to the following
problem on $V^\#$
\begin{equation*}
\begin{cases} \norm{\nabla_y \tilde{\phi}(y) }= \frac{1}{f(\rho(y,t))}+W(y) &\text{ in } V_x \\
\norm{\nabla_y \tilde{\phi}(y) }= \frac{1}{f(0)} &\text{ in } V^\# \backslash V_x\\
\tilde{\phi} \equiv \phi_H &\text{ on } \partial V^\#
\end{cases}
\,.
\end{equation*}
\end{lemma}
\begin{proof} For any point $z$ whose default optimal path
$\gamma_H^z$ that does not intersect with $V$, the claim is that
$\tilde{\phi}(z)=\phi_H(z)$ due the monotonicity of the cost
function, i.e.,
$$
\frac{1}{f(\rho(y,t))}+W(y) \geq \frac{1}{f(0)}\,.
$$
To prove this, let us denote by $T_z^H$ the optimal time to get to
the exit for the default optimal path $\gamma_H^z$.
We first take $\gamma_{x,t}^z(s)$ as a candidate path in the
global optimality condition \eqref{e:optimality2} for the eikonal
equation with right hand side $c_H=\tfrac{1}{f(0)}$. Being
$\gamma_{x,t}^z(s)$ a path joining $z$ to a point in the exit and
$\gamma_H^z(s)$ the optimal one, we conclude
$$
\phi_H(z) = T_z^H c_H \leq T_z c_H \leq \int_0^{T_z}
c(\gamma_{x,t}^z(s),t) \,ds = \tilde\phi(z) \,.
$$
Now, we take $\gamma_H(z)$ as a candidate path in the global
optimality condition \eqref{e:optimality2} for the eikonal
equation with right hand side $c(y,t)$. It is an admissible path
as it connects $z$ to a point at the exit and the cost along its path
coincides with $c_H=c(\gamma_H^z(s),t)$ for all $s\in [0,T_Z^H]$
since the path does not cross $V$. Then, we get
$$
\tilde\phi(z) \leq \int_0^{T_z^H} c(\gamma_H^z(s),t) \,ds = T_z^H
c_H = \phi_H(z) \,,
$$
leading to the stated result.
\end{proof}
\begin{figure}
\begin{center}
\psxinput{\input{draw_2d_cones}}{1} \caption{Illustration of the
domains $V,M_H$ and $V^{\#}$ for computation of the visibility
area potential $\tilde{\phi}$ for the case of a corridor with two
opposing exits: The problem on $\Omega$ generally reduces to a
HJ-equation on $M_H$, as by construction $\tilde{\phi}$ coincides
with $\phi_H$ outside of $M_H$ (Lemma \eqref{lemma-Mreduction},
$-\nabla \phi_H$ solid arrows). If $\rho_H=0$, any default optimal
path of $\phi_H$ that does not intersect $V$ remains optimal, as
indicated by dotted arrows, and the problem reduces to $V^\#$
(Lemma \ref{lemma-Vreduction}).} \label{fig-MVreduction}
\end{center}
\end{figure}
We illustrate Lemmata \ref{lemma-Mreduction} and \ref{lemma-Vreduction} in Figure \ref{fig-MVreduction}.
It can be seen, that the reduction of the computational domain
from $M_H$ to $V^{\#}$ can be significant, as the size of $M_H$
depends on the closeness of $V$ to the nearest exit, not on the
size of $V$. For the exemplary geometry of Figure \ref{fig-MVreduction}, the boundary of $V^{\#}$ coincides with
the sonic points of $\phi_H$, but this is not true in general.
Furthermore, it is easy to see why the computational domain cannot be reduced further. Suppose that $\rho(\cdot,t)$ is spatially homogeneous, then
$-\nabla\phi$ in
$V^\#\backslash V$
points to the left exit as in the eikonal case. On the other side, one can choose
a situation with a large density at the left boundary of
$V$ that leads to right-pointing $-\nabla\phi$ in
$V^\#\backslash V$.
\section{Computational methods}\label{s:computmeth}
In this section we present a microscopic and a macroscopic numerical solver to simulate the classic and the local
version of the Hughes model. The proposed methods have been implemented on regular and triangular meshes in 2D to allow for flexible
discretizations of polygonal domains with one or several obstacles.
For the macroscopic system \eqref{e:localmodel} we use the
following explicit iterative algorithm:
\begin{algorithm}
\caption{Macroscopic version of the localised model}
\label{a:macrosim}
Initialisation: \begin{itemize}
\item A discretisation $\hat\Omega=( \mathcal{V}, \mathcal{E}, \mathcal{T})$ of $\Omega$ consisting of vertices, edges and cells.
\item An initial density $\hat\rho_0$ given on $\mathcal{T}$, such that $\int_{T} \rho(x,0)\mathrm{d} x = \hat\rho_0(T) \,\, \forall \, T \in \mathcal{T}$.
\item A list of exits, a list of boundary edges per exit and $|\mathcal{V}|$ subsets of $\mathcal{V}$ containing the vision cones defined in terms of vertices.
\end{itemize}
\begin{enumerate}
\item Compute the cost potential $\hat\phi_k$ for all exits out of the current density $\hat\rho$ by solving \eqref{e:localeikonal} along the vertices for every $v\in \mathcal{V}$.
\item Determine the cell values of $\hat\phi_k$ and $\nabla\hat\phi_k$ by an averaging / finite difference approximation of the values at neighbouring vertices, e.g. $\hat\phi_k(T)=\frac{1}{|\{v\in\partial T\}|}\sum_{v\in\partial T} \hat\phi_k(u)$, and obtain $\hat u(T)$ here from.
\item Compute a numerical convolution of $\hat u$ with $\mathcal{K}$, which gives $\hat\varphi$ on the cells.
\item Update the density with a cell-based Finite Volume Method using the velocity field $-f(\hat\rho)\mathcal{P}[\hat\varphi]$ and a suitably chosen time step.
\end{enumerate}
\label{algo:macro}
\end{algorithm}
The discretisation is either a regular grid or an unstructured
regular triangular mesh to allow more complex geometries. For
solving the eikonal equations, one can chose between Fast Sweeping
Methods\cite{zhao05,QZZ2007} and Fast Marching
Methods\cite{kimmel1998computing,SethianRev1999}. The former is
based on a Gauss-Seidel iteration, which updates the solution by
passing through the computational domain in alternate pre-defined
sweeping directions. A rectangular grid provides a natural
ordering of all grid points. This ordering does not exist on an
unstructured grid and is replaced by a general ordering strategy
by introducing reference points, which is done once. Then the
solution at each node is consecutively updated by running through
the ordered lists. Marching methods update vertices in a monotone
increasing order, where in every iteration a list of candidate
values is available by finite difference approximation from
previously approved values. The smallest value all of candidate
values is then promoted and assigned to its vertex.
As a Finite Volume Method we use the first-order monotone FORCE
scheme\cite{toro2009force,Toro2010}. Some postprocessing between
the steps of Algorithm \ref{algo:macro} is required:
outward-pointing components of $\nabla\hat\phi_k$ are removed
along the boundary, suitable values of $\nabla\hat\phi_k$ are
ensured at corners of $\Omega$, and the max outflow condition
\eqref{e:maxoutflow} is enforced at cells neighbouring exit edges.
The analogous algorithm used for the numerical simulation of the
microscopic model is:
\begin{algorithm}
\caption{Microscopic version of the localised model}
\label{a:microsim}
Let us consider a system of $N$ particles, which are initially located at positions $X^j(0) = X^j_0$. In every time step
$t^i = i \Delta t$ we update the particle position as follows:
\begin{enumerate}
\item Determine the empirical density at time $t^i$:
\begin{align*}
\rho^N_g(x,t^i) = \frac{1}{N} \sum_{j=1}^N g(x-X^j(t^i)),
\end{align*}
where $g$ denotes a Gaussian.
\item Solve the eikonal equation to determine the weighted distance to each exit $\phi^k = \phi^k(x,t^i)$, $k=1, \ldots M$:
\begin{subequations}\label{e:microeikonal}
\begin{align}
\norm{\nabla \phi^k(x,y,t^i)} &=
\begin{cases}
\frac{1}{ f(\rho^N_g(y,t^i))} +W(y) &\text{ if } y \in V_x\\
\frac{1}{f(\rho_H)} &\text{ otherwise.}
\end{cases}\\
\phi^k(x,t^i) &= 0
\end{align}
\end{subequations}
\item Update the position of each particle $X^j$ via:
\begin{align}
\dot{X^j}(t^i) = -f^2(\rho^N_g(x,t^i) \cdot \nabla \varphi(x,t^i))),\label{e:microupdate}
\end{align}
where $\varphi(x,t^i)$ is determined by \eqref{e:localinteraction}.
\end{enumerate}
\end{algorithm}
\section{Results}\label{s:numexp}
In this section we illustrate the dynamics of the localised model
for crowd dynamics with examples in one and two dimensions. In all
simulations we consider an evacuation scenario of a corridor,
where a given initial distribution of people tries to leave the
rectangular domain through either one of the two exits as fast as
possible. We compare the evacuation time, i.e. the time at which
all individuals have left the domain, with respect to different
parameters, e.g. vision cones. In the case of a global vision cone
we obtain Hughes' type dynamics. As a flux law, we chose the LWR
function \begin{equation}
f(\rho)=\rho(1-\rho)
\label{fluxlaw},
\end{equation}
setting $\rho_{\max}=1$ throughout this section.
\subsection{1D corridor - macroscopic model} \label{s:1dcorrmac}
In our first example the domain $\Omega$ corresponds to the unit
interval $\Omega=[0,1]$ with two exits located at either end, i.e.
at $x=0$ and $x=1$. We consider an evacuation scenario in which
two groups, one of them being densly packed, want to leave through
either one of the exits:
\begin{equation}
\rho_0(x) = \begin{cases}
0.85 & 0\geq x \geq 0.3 \\
0 & 0.3< x < 0.6 \\
0.25 & 0.6\geq x \geq 1
\end{cases},
\label{e:initsymcorridor}
\end{equation}
and we set the width of the vision cone to $L=0.75$. The resulting
dynamics are illustrated at 4 time steps in Fig.
\ref{fig-1Dturnaround}. Within the left block, some pedestrians
decide to walk towards the right exit, as they are aware of the
high density on their left and account for a higher walking cost
compared to the relatively empty right hand side. After the
separation the right-moving part evolves as a rarefaction wave, as
known from the LWR model. As the distance between the wave and the
left-moving shock grows, the effects of the local vision cone
become apparent. At some point pedestrians moving to the right
do not see the high density at the left exit anymore and start to
turn around. Therefore the rarefaction wave splits again - one
part continues while the other one turns around and moves back to
the left exit. The turn-around occurs in several stages:
\begin{enumerate}
\item A new sonic point arises, where pedestrians are undecided between both exits. The walking direction is unchanged as the local consensus process \eqref{e:localinteraction} prevents an immediate switching.
\item When a critical mass of density and conviction opting for walking to the left, the velocity after consensus switches continuously and passing through zero. This creates a temporary collision point, as there a still pedestrians to the left of the sonic point which walk towards the right.
\item The density at the collision point increases, which causes pedestrians to the left of the collision point to turn around too, as a higher density is in their way, as it can bee seen in Fig. \ref{fig-1Dturnaround}(c).
\item Finally, all pedestrians to the left of the initial sonic point have turned and walk towards the left ( Fig. \ref{fig-1Dturnaround}(d)).
\end{enumerate}
\begin{figure}
\centering
\begin{tabular}{c c}
\psxinput{\input{draw_num_1d_ex1_a}}{.9} &
\psxinput{\input{draw_num_1d_ex1_b}}{.9} \\
(a) $t_0=0 $& (b) $t_1=0.31 $\\
\psxinput{\input{draw_num_1d_ex1_c}}{.9} &
\psxinput{\input{draw_num_1d_ex1_d}}{.9} \\
(c) $t_2=0.71 $& (d) $t_3=1.29 $
\end{tabular}
\caption{Exemplary evolution of the 1D model showcasing a
turnaround behavior due to localised perception of information
[density $\rho$ solid ($-$), speed $v=-f(\rho)\mathcal{P}[\nabla
\varphi]$ dashed ($- -$) and directional conviction
$\phi_R-\phi_L$ dotted ($\cdot \cdot$)]. (a) Piece-wise constant
initial density. Part of the left crowd initially decides to move
right in order to avoid the high-density jam. (b) The separated
block moves to the right in a rarefaction-wave manner. (c) The
wave is again separated as the high-density jam gets out of sight
for centrally located pedestrians, who hence prefer the left exit
and turn. (d) The turnaround is complete and remaining pedestrians
will exit on the left.} \label{fig-1Dturnaround}
\end{figure}
This new behavioural pattern is entirely consistent with the idea
of constant re-evaluation of the optimal path based on restricted
information and cannot be observed in the original Hughes' model.
We note that without the smoothening properties of the model
around points of equal costs one obtains strong oscillations in
the turning behavior, which causes severe numerical problems. The
exact parameters of the simulation can be found in Appendix \ref{a:1d}
\subsection{2D corridor - microscopic model}
\label{s:num2dmicro}
\begin{figure}
\begin{center}
\subfigure[Time $t=0$]{\includegraphics[width=0.325 \textwidth]{particles_at_time0.0.eps}}
\subfigure[Time $t=0.2$]{\includegraphics[width=0.325 \textwidth]{particles_at_time0.2.eps}}
\subfigure[Time $t=0.4$]{\includegraphics[width=0.325 \textwidth]{particles_at_time0.4.eps}}\\
\subfigure[Time $t=0.6$]{\includegraphics[width=0.325 \textwidth]{particles_at_time0.6.eps}}
\subfigure[Time $t=0.8$]{\includegraphics[width=0.325 \textwidth]{particles_at_time0.8.eps}}
\subfigure[Time $t=1$]{\includegraphics[width=0.325 \textwidth]{particles_at_time1.eps}}\\
\subfigure[Exit percentage]{\includegraphics[width=.45 \textwidth]{maryplot}}
\end{center}
\caption{Dynamics of the microscopic model in a two-dimensional
corridor with two exits at the right and left. We observe a
similar behavior as in the 1D simulation in \ref{s:1dcorrmac} -
individuals (visualized by red triangles) initially decide to move
to the more distant exit, but after the congestion at the resolves
in time, they turn around and take the closer
exit. (a) - (f): particle solution for different times and $L=0.25$, (g): exit percentage over time for different values of L. }
\label{f:corridor_micro}
\end{figure}
We illustrate the dynamics of the microscopic model in a
two-dimensional symmetric corridor $\Omega = [0,1] \times [0,
\frac{1}{2}]$ with exits at the left and right side, i.e. $x = 0$
and $x = 1$. The 1D case of section \ref{s:1dcorrmac} can be
interpreted as a projection of this two-dimensional geometry. We
consider the same initial distribution of individuals, i.e. the
positions of all $500$ particles are distributed according to the
initial pedestrian density \eqref{e:initsymcorridor}. For
$L=0.25$, Figure \ref{f:corridor_micro}(a)-(f) nicely illustrates
a similar turn-around behavior as in the 1D macroscopic
simulations. At the beginning the group close to the left exit
splits, one part exits through the left exit the other one moves
towards the more distant right exit. As the density close to the
left exit decreases in time, the group moving towards the more
distant exit splits again, i.e. parts of the group turn around and
move back again. We marked all individuals, which initially moved
towards the right but then turn around, with red triangles.
Furthermore, Figure \ref{f:corridor_micro}(g) shows the change of
the evacuation performance for different sizes of the local vision
cones $L$. Here we plot the percentage of the total initial mass
outside the domain versus time. Decreasing $L$ and hence the
perceived information, we observe that the overall evacuation
performance first is merely diminished, and only begins to drop
significantly after a certain threshold. The evacuation time will
approach the uninformed eikonal case $L=0$, which is not shown.
All parameters can be found in Appendix \ref{a:2dmicro}.
\subsection{2D non-symmetric corridor - macroscopic model}
\label{s:num2dmacro}
Now we turn to the macroscopic model in two dimensions. Again we
consider the corridor $\Omega=[0,1]\times[0,\frac{1}{2}]$, the
exits however form only a part of the left and right edges, hence
we obtain a fully two-dimensional dynamics where boundary
conditions matter. The left exit is located between $(0,0)$ and $(0,0.1)$ and the right exit is the segment connecting $(1,\tfrac{1}{2})$ and $(1,0.4)$.
The initial density Figure \ref{fig-2dmacroglobal}(a) is given as a low density group of pedestrians on the left and a high density group on the right
\begin{equation}
\rho_0(x,y) = \begin{cases}
0.1 & 0.05\leq x \leq 0.3 \,,\, 0\leq y \leq 0.25, \\
0.95 & 0.6\leq x \leq 0.95,\\
0 & \text{otherwise}.
\end{cases}
\end{equation}
\begin{figure}
\begin{center}
\subfigure[Time $t=0$]{ \includegraphics[height=.205\textwidth, width=.4325\textwidth]{pic_2dmacro_local_t000} }\hspace{1mm}
\subfigure[Time $t=0.8$]{\includegraphics[keepaspectratio=true,width=.45\textwidth]{pic_2dmacro_global_t080}}\\
\subfigure[Time $t=1.07$]{\includegraphics[keepaspectratio=true,width=.45\textwidth]{pic_2dmacro_global_t107}}
\subfigure[Time $t=1.4$]{\includegraphics[keepaspectratio=true,width=.45\textwidth]{pic_2dmacro_global_t140}}\\
\subfigure[Time $t=2.1$]{\includegraphics[keepaspectratio=true,width=.45\textwidth]{pic_2dmacro_global_t210}}
\subfigure[Time $t=2.75$]{\includegraphics[keepaspectratio=true,width=.45\textwidth]{pic_2dmacro_global_t275}}\\
\end{center}
\caption{Two-dimensional macroscopic dynamics : We simulate model \eqref{e:localmodel} with global visual perception $V_x=\Omega$. Two groups of pedestrians are initially placed in a corridor with a lower left and an upper right exit. The high density group separates according to the path optimization mechanism, as illustrated in several time snapshots of the density in (a) - (f). The right exit is vacated before the left exit and the final evacuation time (not shown here) is $\approx 3.4$.}
\label{fig-2dmacroglobal}
\end{figure}
We first study the case of global vision $L=\infty \Leftrightarrow
V_x=\Omega$ in Figure \ref{fig-2dmacroglobal}. In (b), the low
density group turns towards left and is quickly vacated. The high
density group on the other side splits along a curve of sonic
points. Pedestrians turning to the right cause a jam in front of
the right exit, whereas left-turning pedestrians occupy the
corridor in a rarefaction-type manner inherited by the physical
flux law \eqref{fluxlaw}. Upon arrival at the left exit,
pedestrians pile up and form a new jam (c). Hence, a fraction of
the density turns around again and heads for the right exit (d,
around $(0.5,.25$)), having to cross most of the corridor again
(e). However, most of the pedestrians are committed to the left
exit and do not turn, because the severeness of the left jam does
not compensate their expected travel time, and the left exit is
vacated later than the right exit (f).
Compared to the classical Hughes model, the relaxation term
\eqref{e-relaxation} and the conviction-based interaction
\eqref{e:localinteraction}-\eqref{e:conviction} allow for a smooth
turning behaviour. The wall-repulsion \eqref{e:wall} causes a
density gap, which is different to zero-flux conditions as in
\cite{HuangonHughes2009}, but prevents any spurious effects of the
boundary flux.
\begin{figure}
\begin{center}
\subfigure[Time $t=0.25$]{ \includegraphics[keepaspectratio=true,width=.45\textwidth]{pic_2dmacro_local_t025}}
\subfigure[Time $t=0.8$]{ \includegraphics[keepaspectratio=true,width=.45\textwidth]{pic_2dmacro_local_t080}}\\
\subfigure[Time $t=1.07$]{ \includegraphics[keepaspectratio=true,width=.45\textwidth]{pic_2dmacro_local_t107}}
\subfigure[Time $t=1.4$]{ \includegraphics[keepaspectratio=true,width=.45\textwidth]{pic_2dmacro_local_t140}}\\
\subfigure[Time $t=2.1$]{ \includegraphics[keepaspectratio=true,width=.45\textwidth]{pic_2dmacro_local_t210}}
\subfigure[Time $t=2.75$]{ \includegraphics[keepaspectratio=true,width=.45\textwidth]{pic_2dmacro_local_t275}}\\
\end{center}
\caption{Two-dimensional macroscopic dynamics : We simulate model \eqref{e:localmodel} with local visual perception $L=0.75$ and the same initial configuration as Figure \ref{fig-2dmacroglobal}. In the time snapshots of the density, we observe again an initial separation of the high density group (a). A waiting phenomena of up to two local groups is clearly observed (b-d). Crucially, waiting pedestrians are able to choose their exit later (d-e), which leads to a rather simultaneously clearing of both jams (f), as opposed to Figure \ref{fig-2dmacroglobal}. The final evacuation time (not shown here) is $\approx 3.025$.}
\label{fig-2dmacrolocal}
\end{figure}
It is clear that even the planning algorithm incorporated into the
classic Hughes' model does not lead to an optimal evacuation. One
reason for suboptimality in Figure \ref{fig-2dmacroglobal} is that
pedestrians \emph{have to} keep in motion constantly but cannot
predict the occurrence of future jams. Hence, pedestrians are
likely to walk towards an exit that will be blocked in the future,
as seen in the example.
In Figure \ref{fig-2dmacrolocal} we study the same initial
configurations with localised perception and a radial vision cone
of diameter $L=0.75$. The initial separation phase (a) is similar
to Figure \ref{fig-2dmacroglobal}. As pedestrians move from the
right to the left, the right jam gets out of sight and its
influence diminishes. At the same time, the density on the left
becomes visible. At a certain point a balance is achieved and
pedestrians locally accumulate around an area of equal walking
costs, where in this case they are able to stop (b-c). Hence, we
observe a waiting behavior which cannot be observed in classical
Hughes' type models. Looking from (c) to (d), a high density jam
forms at the left exit, which causes part of the left-walking
pedestrians to turn right after enough conviction is gathered.
Together with some outflow of the first waiting group, a second
waiting group is formed (d). Pedestrians in a waiting group choose
to move if one direction becomes favorable. As both jams at the
exits reduce at the same rate, the left waiting group walks to the
left and vice versa (e). Finally, the waiting groups dissolve and
the exits get vacated at a rather similar time (f), and the
evacuation time improves compared to Figure
\ref{fig-2dmacroglobal}.
\begin{figure}
\begin{center}
\includegraphics[width=0.6\textwidth]{evacuationplot}
\end{center}
\caption{Evacuation time of the macroscopic model for varying vision cone diameters $L$: The performance can improve with limited vision. $L=0$ corresponds to the eikonal case whereas $L>2.5$ implies unlimited vision in the given corridor.}
\label{fig-2dmacroevac}
\end{figure}
The fact that evacuation performance can improve under limited
perception of information is surprising at first glance. Our
simulations give an good explanation for the phenomena: As
pedestrians show a waiting behaviour, they are less likely to be
trapped in the jam arising at the left exit. In fact, the waiting
is made possible by the combined effect of multiple sonic points
due to local vision and the smoothed turning mechanism. Naturally,
this is not generally the case and cannot be a-priori answered.
For small vision lengths $L$, the dynamics will converge to the
velocity field given by the eikonal equation, which is our initial
configuration will exit almost all pedestrians using the right
exit and perform poorly. In Figure \ref{fig-2dmacroevac}, we study
the evacuation time of $99\%$ of the initial mass as a function of
the diameter $L$. We unexpectedly find in this case wo optimal
values of $L$ for which the evacuation time is minimal. The
classical Hughes' evacuation time ($L$ large) is always less or
equal than the eikonal case ($L=0$), however there is no way to
generally argue that there will always be a minimum in between.
\section{Conclusion}\label{s:conclusion}\label{s:conclusions}
In this work we introduced a localized smooth variant of Hughes's
model for pedestrian crowd dynamics. We regularised the original
model, composed by an eikonal equation and a continuity equation.
First by a local interaction term, which intermediates individual
pointwise path optimisation towards conviction-weighted walking
directions. Secondly, we allowed pedestrians to stop, if they are
undecided, using a smooth approximation of the normalization
condition. Most importantly, we restrict the information on the
global density each pedestrian can use for her planning algorithm
to a local surrounding area. This is a very realistic assumption
for large crowds that has not been considered in the literature so
far. We presented both a microscopic and a macroscopic version,
and illustrated the model components in the one-dimensional case.
In terms of analytical results, a rigorous theory for these kind
of equations in multiple dimensions is currently out of reach to
the best of our knowledge. However, we were able to identify some
qualitative properties of the dependence of the optimal path on
the vision cone that allow for a reduction of complexity.
The numerical approximation of the model on both levels has been discussed and utilizes several techniques including sweeping and marching methods, particle approximations and finite volume schemes. Though the numerical costs of computing a solution have increased due to the inhomogeneity of vision cones, we observe new effects and phenomena in the model based on our simulations.
First, local groups of pedestrians are able to change repeatedly their walking direction towards an exit. This 'multiple turn-around behaviour' can explained by the multiple sonic points of the estimated walking costs, which by construction cannot occur in the classical case.
We stress that the smoothening and conviction terms are crucial to allow a swift turning behavior, which is not trivial to model in first order equations.
Second, the model replicates a waiting behaviour in case of undecided pedestrians, i.e. in areas where locally estimated walking costs towards different exits are equal.
Surprisingly, we found that this waiting phenomena induced by localized information can improve the overall evacuation performance of the crowd. In our numerical example
we observed two local minima when varying the vision cone diameter.
To conclude, we have demonstrated that local vision effects can be implemented into first order models for crowd dynamics. This leads to new unforeseen phenomena and complex behavior, whose partial understanding via qualitative properties is important for the applicability of such equations to social-economic problems. On the other hand, this work illustrates the limitations to first order models such as Hughes', where planning decisions are instantaneously updated and no social or cognitive memory is taken into account. From our point of view Hughes' type equations constitute an important building block for crowd models and a mathematically important object of study, but it cannot be expected to be fully realistic.
\section*{Acknowledgment}
JAC acknowledges support from projects MTM2011-27739-C04-02 and
the Royal Society through a Wolfson Research Merit Award. JAC and
SM were supported by Engineering and Physical Sciences Research
Council (UK) grant number EP/K008404/1. MTW acknowledges financial
support from the Austrian Academy of Sciences \"OAW via the New
Frontiers Group NSP-001.
Preprint of an article submitted for consideration in Mathematical Models and Methods in Applied Sciences, \copyright 2015 World Scientific Publishing Company, http://www.worldscientific.com/worldscinet/m3as.
|
\section{Introduction}
\label{sec:intro}
Bose-Einstein condensates (BECs) with a multicomponent order parameter, and the topological defects
such systems support, represent a topic of great current interest in condensed matter
physics.\cite{Wieman97, Cornell98, Cornell99, Modugno2002, mueller02, Kasamatsu2003, Cornell2004,
Papp2008, Thalhammer2008, makoto2, Tojo2010, McCarron2011, mottonen, tsubota3, cipriani} Such
multicomponent condensates may be realized as mixtures of different atoms, mixtures of different
isotopes of an atom, or mixtures of different hyperfine spin states of an atom. The interest in such
condensates from a fundamental physics point of view is mainly attributed to the fact that one may
tune various interaction parameters over a wide range in a BEC.
This enables the study of a variety of physical effects which are not
easily observed in other superfluid systems such as He$^3$ and He$^4$.
The behavior of a single-component BEC under rotation is well
known. The ground state is a hexagonal lattice of vortex defects which melts to
a vortex liquid via a first-order phase transition. This is well described
by the London model, where amplitude fluctuations may be
ignored. Over the years, in the context of studying vortex lattice melting in high-$T_c$ superconductors,
many works have confirmed this through numerical Monte Carlo simulations for systems in the frozen gauge,
three-dimensional (3D) $XY$ and Villain approximations, \cite{Teitel91,Sudbo92,Teitel97,Hu97,Nguyen98_1,Nguyen98_2,Stroud98,Nguyen99,Chin99,Nguyen_EPL99} as
well as in the lowest-Landau-level approximation,\cite{Nordborg97} and by mapping it to a model of
2D bosons.\cite{MacDonald97} Single component condensates have been available experimentally
for quite some time\cite{Cornell95,Ketterle95}, and the hexagonal lattice ground state has been
verified.\cite{Abo-Shaeer2001}
Condensates with two components of the order parameter have also been studied extensively.
Analytical works focusing on determining the $T=0$ ground states have demonstrated interesting vortex solutions and a range of
unusual lattice structures\cite{mueller02,
Kasamatsu2003,makoto2,tsubota3,koba,mottonen,catelani}. By varying the ratio between inter- and
intra-component couplings, the ground-state lattice undergoes a structural change from hexagonal
symmetry through square symmetry to double-core lattices and interwoven sheets of vortices. Similar
systems with three components have also been studied.\cite{cipriani} Experimentally, spinor condensates
have been realized in two general classes of systems. The first option is to use one species of atoms,
usually rubidium, and prepare it in two separate hyperfine spin states.\cite{Wieman97,Cornell98}
Vortices\cite{Cornell99} and vortex lattices\cite{Cornell2004} have been realized in these binary mixtures,
where both hexagonal and square vortex lattice states were observed. The other option is to mix condensates of two different
species of atoms.\cite{Modugno2002,McCarron2011} The use of Feshbach resonances\cite{Feshbach,Inouye1998}
allows direct tuning of the scattering lengths, and by extension the inter- and intracomponent
interactions of multicomponent condensates.\cite{Papp2008,Thalhammer2008, Tojo2010}
In this paper, we consider a specific model of a two-component BEC, which has the full range of
fluctuations of the order parameter field included, as well as intercomponent density-density
interactions. We consider the model with $\mathrm{U(1)}\times\mathrm{U(1)}$- and as
$\mathrm{SU(2)}$ symmetries. For the $\mathrm{U}(1) \times \mathrm{U}(1)$ case, we find a succession
of square and hexagonal vortex ground-state patterns as the intercomponent interaction strength is
varied, along with the possibility of thermal reconstruction from a square to a hexagonal vortex
lattice as temperature is reduced.
The $\mathrm{SU}(2)$-symmetric case is interesting and experimentally realizable. In this case
$\mathrm{U}(1)$-vortices are no longer topological, in contrast to the $\mathrm{U(}1) \times
\mathrm{U}(1)$-symmetric case. In this case, when fluctuation effects are included we find a highly
unusual vortex state where there is no sign of any vortex lattice. Nonetheless, global phase
coherence persists. This state of vortex matter is a direct consequence of massless
amplitude fluctuations in the order parameter, when the broken symmetry of the system is
$\mathrm{SU}(2)$. At the $\mathrm{SU}(2)$ point, but at lower temperatures, we also observe
dimerized vortex ground-state patterns.
The paper is organized as follows. The model and definitions of relevant quantities are presented
in Section~\ref{sec:model}. The technical details of the Monte Carlo simulations are
briefly considered in Section~\ref{sec:MCdetails}. In Section~\ref{sec:results}, the results are
presented and discussed. In Section~\ref{sec:exp}, we discuss how to experimentally verify the results
we find. Some technical details, and the investigation of the order of the melting transitions
with full amplitude distributions included, for the cases $N=1$ and $N=2$, are relegated to Appendixes.
\section{Model and definitions}
\label{sec:model}
In this section we present the model used in the paper, first in a continuum description and then on a
three-dimensional cubic lattice appropriate for Monte Carlo simulations. The relevant quantities for
the discussion are also defined.
\subsection{Continuum model}
We consider a general Ginzburg-Landau(GL) model of an $N$-component Bose-Einstein condensate, coupled to a uniform
external field, which in the thermodynamical limit is defined as
\begin{equation} \mathcal{Z}=\int\prod_i^N\mathcal{D}\psi_i^\prime
\text{e}^{-\beta H},
\end{equation}
where
\begin{align}
H=\int
d^3r\Bigg[&\sum_{i=1}^N\sum_{\mu=1}^3\frac{\hbar^2}{2m_i}\left|(\partial_\mu-\mathrm{i}\frac{2\pi}{\Phi_0}A_\mu^\prime)\psi_i^\prime\right|^2\nonumber\\
&+\sum_i^N\alpha_i^\prime\abs{\psi_i^\prime}^2+
\sum_{i,j=1}^Ng_{ij}^\prime\abs{\psi_i^\prime}^2\abs{\psi_j^\prime}^2\Bigg]
\label{eq:genH}
\end{align}
is the Hamiltonian. Here, the field $A_\mu^\prime$ formally appears as a non-fluctuating gauge-field
and parametrizes the angular velocity of the system. The fields $\psi_i^\prime$ are dimensionful
complex fields, $i$ and $j$ are indices running from $1$ to $N$ denoting the component of the order
parameter (a ``color" index), $\alpha_i^\prime$ and $g_{ij}^\prime$ are Ginzburg-Landau parameters,
$\Phi_0=h/2e$ is the the coupling constant to the rotation induced vector potential, and $m_i$ is
the particle mass of species $i$. For mixtures consisting of different atoms or different isotopes
of one atom, the masses will depend on the index $i$, while for mixtures consisting of atoms in
different hyperfine spin states, the masses are independent of $i$. The inter- and intracomponent
coupling parameters $g_{ij}^\prime$ are related to real inter- and intracomponent scattering
lengths, $a_{ij}$, in the following way
\begin{align}
g_{ii}^\prime &= \frac{4\pi\hbar^2 a_{ii}}{m_i},\\
g_{ij}^\prime &= \frac{8\pi\hbar^2 a_{ij}}{m_{ij}}, (i\neq j)
\end{align}
where $m_{ij}=m_i~m_j/(m_i+m_j)$ is the reduced mass. In this paper we focus on using BECs of
homonuclear gases with several components in different hyperfine states, hence $m_i=m\,\forall\, i$.
Inter-component drag in BEC mixtures has been considered in previous works using Monte-Carlo simulation
(ignoring amplitude fluctuations), but we will not consider this case here.~\cite{Dahl2008_2,Dahl2008,Dahl2008_3,2004PhRvL..92c0403K,2004PhRvL..93w0402K}
We find it convenient for our purposes to rewrite~\eqref{eq:genH} on the following form, the details of which are relegated to
Appendix~\ref{app:rewrite},
\begin{equation}
H = \int d^3r\Bigg[\frac{1}{2}(D_\mu\Psi)^\dag(D_\mu\Psi)+V(\Psi)\Bigg].
\label{eq:Ham}
\end{equation}
Here, $\Psi$ is an $N$-component spinor of dimensionless complex fields, which consists of an
amplitude and a phase, $\psi_i=\abs{\psi_i}\exp{(i\theta_i)}$, $D_\mu =
\partial_\mu-\mathrm{i}\frac{2\pi}{\Phi_0}A^\prime_\mu$ is the covariant derivative, and summation over repeated spatial
indices is implied. We neglect, for simplicity, the presence of a trap and the centrifugal part of the
potential. We consider only the case where the vector potential is
applied to each component of $\Psi$, as follows from the fact that the masses are independent of
species-index $i$.
We have studied this model in detail with $N=2$, where we write the potential in the form
\begin{equation}
V(\Psi)=\eta(\left|\Psi\right|^2-1)^2+\omega(\Psi^\dag\sigma_z\Psi)^2.
\label{eq:N2pot}
\end{equation}
This formulation is more relevant for our discussion, as it immediately highlights the symmetry of
$\Psi$, as well as the soft constraints applied to it. The details of the reparametrization are shown
in Appendix~\ref{app:rewrite}.
Note that Eq. \ref{eq:N2pot} may also be rewritten in the form (correct up to an additive constant term)
\begin{equation}
V=(\eta+\omega)(\left|\psi_1\right|^4+\left|\psi_2\right|^4)+2(\eta-\omega)\left|\psi_1\right|^2\left|\psi_2\right|^2
\label{eq:inter_intra_g}.
\end{equation}
Comparing with Eq. \ref{eq:genH}, we have $g_{11}=g_{22}\equiv g=\eta+\omega$ and $g_{12}=\eta-\omega$. The model features
repulsive inter-component interactions provided $ \eta - \omega > 0$, and this is the case we will mainly focus on. We will
however briefly touch upon the case $\eta - \omega < 0$ corresponding to an attractive inter-component density-density
interaction, which leads to ground states with overlapping vortices in components $1$ and $2$. Normalizability of the individual
order parameter components, or equivalently boundedness from below of the free energy, requires that $\eta + \omega> 0$. Thus,
while $\omega > \eta$ makes physical sense, $\omega < -\eta$ does not. In this paper, we assume
$\eta > 0$ and {$\omega \geq 0$.}
Two-component BECs feature considerably richer physics than a single-component BEC. Since the gauge-field
parametrizing the rotation of the system is non-fluctuating, there is no gauge-field-induced current-current interaction
between the two condensates (unlike in multi-component superconductors). The only manner in which the two superfluid
condensates interact is via the inter-component density-density interaction $2 (\eta-\omega) |\psi_1|^2 |\psi_2|^2$. In
the limit where the amplitudes of each individual component are completely frozen and uniform throughout the
system, one recovers the physics of two decoupled 3D$XY$ models, with a global $\mathrm{U(1)}\times
\mathrm{U(1)}$ symmetry. The density-density interaction between $\psi_1$ and $\psi_2$ leads to
interactions between the topological defects excited in each component. As a result, a first
order melting of two decoupled hexagonal lattices is not the only possible phenomenon that could
take place. Previous experiments and numerical studies have reported a structural change of the ground state from a hexagonal
to a square lattice of vortices as the effective inter-component coupling is increased.\cite{mueller02,Cornell2004,Kasamatsu2003}
This corresponds to increasing the ratio $\eta/\omega$ in our case. As we will see below, other
unusual phenomena can also occur, notably when thermal fluctuations are included.
One special case of the model deserves some extra attention. If one takes the limit $\omega
\rightarrow 0$ in Eq.~\eqref{eq:N2pot} the symmetry of the model is expanded to a global
$\mathrm{SU(2)}$ symmetry. One may then shift densities from one component to the other with
impunity, as long as $|\psi_1|^2+|\psi_2|^2$ is left unchanged. This effectively leads to massless
amplitude-fluctuations in the components of the order parameter. Therefore, it is possible to unwind
a $2 \pi$ phase winding in one component by letting the amplitude of the same component vanish.
The introduction of this higher symmetry leads to very different vortex ground states than what are
found in the $\mathrm{U(1)}\times \mathrm{U(1)}$-symmetric case with $\omega \neq 0$.
\subsection{Separation of variables}
In multicomponent GL models complex objects, such as combinations of vortices
of different colors, are often of interest. In general, it is possible to
rewrite an $N$-component model coupled to a gauge field, fluctuating or not, in
terms of one mode coupled to the field and $N-1$ neutral
modes.\cite{Smiseth2005,Herland2010} For a more general discussion of charged
and neutral modes in the presence of amplitude fluctuations see Refs.
\citenumns{2002PhRvB..65j0512B} and \citenumns{CPN}. Considering only the
kinetic part of the two-component Hamiltionian, $H_k$, we have the following
expression:
\begin{align}
H_k=&\frac{1}{2\left|\Psi\right|^2}\left|\psi_1^*\partial_\mu\psi_1+\psi_2^*\partial_\mu\psi_2-\mathrm{i}A_\mu\left|\Psi\right|^2\right|^2\nonumber\\
+&\frac{1}{2\left|\Psi\right|^2}\left|\psi_1\partial_\mu\psi_2-\psi_2\partial_\mu\psi_1\right|^2.
\end{align}
Hence, the first mode couples to the applied rotation, while the second does not. This corresponds
to the phase combinations $\theta_1+\theta_2$ and $\theta_1-\theta_2$, respectively.
\subsection{Lattice regularization}
In order to perform simulations of the continuum model, we define the field $\Psi$ on a discrete set of
coordinates, \textit{i.e} $\Psi(\mathbf{r})\rightarrow\Psi_\mathbf{r}$, where
$\rvec\in(i\hat{\mathbf{x}}+j\hat{\mathbf{y}}+k\hat{\mathbf{z}} | i,j,k=1,\ldots,L)$. Here, $L$ is
the linear size in all dimensions; the system size is $V=L^3$. We use periodic boundary conditions
in all directions. By replacing the differential operator by a gauge-invariant forward difference
\begin{equation}
\bigg(\frac{\partial}{\partial
r_\mu}-\mathrm{i}A_\mu(\mathbf{r})\bigg)\Psi(\mathbf{r})\rightarrow\frac{1}{a}\bigg(\Psi_{\mathbf{r}+a\hat{\boldsymbol{\mu}}}\text{e}^{-\mathrm{i}\frac{2\pi}{\Phi_0}aA^\prime_{\mu,\mathbf{r}}}-\Psi_\mathbf{r}\bigg),
\end{equation}
and introducing real phases and amplitudes
$\psi_{\rvec,i}=\left|\psi_{\rvec,i}\right|e^{i\theta_{\rvec,i}}$ we can rewrite the
Hamiltonian.
\begin{align}
H=&\sum_{\substack{\rvec,\hat{\boldsymbol{\mu}}\\i}}\big|\psi_{\rvec+\hat{\boldsymbol{\mu}},i}\big|\big|\psi_{\rvec,i}\big|
\cos(\theta_{\rvec+\hat{\boldsymbol{\mu}},i}-\theta_{\rvec,i}-A_{\mu,\rvec})\nonumber\\
+&\sum_\rvec V(\Psi_\rvec).
\end{align}
The lattice spacing is chosen so that it is smaller than the relevant length scale of
variations of the amplitudes.
A dimensionless vector potential, $A_\mu$, has also been introduced.
See Appendix~\ref{app:rewrite} for details. We denote
the argument of the cosine as $\chi_{\rvec,i}^\mu$, as a shorthand.
\subsection{Observables}
An important and accessible quantity when exploring phase transitions is the specific heat of the system,
\begin{equation}
c_V=\beta^2\frac{\langle H^2\rangle-\langle H\rangle^2}{L^3}.
\end{equation}
While crossing a first order transition there is some amount of latent heat in the system, manifesting
itself as a $\delta$-function peak of the specific heat in the thermodynamic limit. On the lattice one
expects to see a sharp peak, or anomaly, at the transition. This is used to characterize the
transition as first order.
A useful measure of the global phase coherence of the system, is the helicity modulus, which is
proportional to the superfluid density. It serves as a probe of the transition from a superfluid to
a normal fluid. In the disordered phase, the moduli in all directions are zero, characterizing an
isotropic normal-fluid phase. The cause of this is a vortex loop blowout. Moving to the ordered
phase, all moduli evolve to a finite value. If we turn on the external field we still have zero
coherence in all directions in the disordered phase. In the ordered phase, however, the helicity
modulus along the direction of the applied rotation jumps to the finite value through a first order
transition. The value of the transverse moduli will remain zero. Formally, the helicity modulus is
defined as a derivative of the free energy with respect to a general, infinitesimal phase twist
along $r_\mu$,\cite{Fisher73}. That is, we perform the replacement
\begin{equation}
\theta_{\rvec,i}\rightarrow\theta_{\rvec,i}^\prime=\theta_{\rvec,i}-b_i\delta_\mu r_\mu
\end{equation}
in the free energy, and calculate
\begin{equation}
\Upsilon_{\mu,(b_1,b_2)} = \frac{\partial^2 F[\theta^\prime]}{\partial\delta_\mu^2}\bigg|_{\delta_\mu=0}.
\end{equation}
Here, $b=(b_1, b_2)$ represents some combination of the phases $\theta_1$ and $\theta_2$,
$b_1\theta_1+b_2\theta_2$.
To probe the individual moduli, $b_i$ is chosen as $b_i=(1,0)$ or $b_i=(0,1)$.
The composite, phase sum variable is represented by the choice $b_i=(1,1)$, while
$b_i=(1,-1)$ is the phase difference.
Generally, for a two-component model, the helicity modulus can be written as the sum of two indivudual moduli, and a
cross term,\cite{Dahl2008_3,Herland2010}
\begin{equation}
\Upsilon_{\mu,(b_1,b_2)} = b_1^2\Upsilon_{\mu,(1,0)}+b_2^2\Upsilon_{\mu,(0,1)} + 2b_1b_2\Upsilon_{\mu,12}.
\label{eq:mixedmodulus}
\end{equation}
For the model considered in this paper, the individual helicity moduli can be written as
\begin{align}
\langle\Upsilon_{\mu,i}\rangle=&\frac{1}{V}\bigg[\bigg\langle\sum_\rvec\psi_{\rvec_i}\psi_{\rvec+\hat{\boldsymbol{\mu}},i}\cos(\chi_{\rvec,i}^\mu)\bigg\rangle\nonumber\\
&-\beta\bigg\langle\bigg(\sum_\rvec\psi_{\rvec,i}\psi_{\rvec+\hat{\boldsymbol{\mu}},i}\sin(\chi_{\rvec,i}^\mu)\bigg)^2\bigg\rangle\bigg],
\end{align}
while the mixed term has the form
\begin{align}
\langle\Upsilon_{\mu,12}\rangle =
-\beta\bigg\langle\bigg(&\sum_\rvec\left|\psi_{\rvec,1}\right|\left|\psi_{\rvec+\hat{\boldsymbol{\mu}},1}\right|\sin(\chi_{\rvec,1}^\mu)\bigg)\nonumber\\
\bigg(&\sum_\rvec\left|\psi_{\rvec,2}\right|\left|\psi_{\rvec+\hat{\boldsymbol{\mu}},2}\right|\sin(\chi_{\rvec,2}^\mu)\bigg)\bigg\rangle.
\label{eq:moduluscross}
\end{align}
We denote the helicity modulus of the phase sum $\Upsilon_{\mu,(1,1)}$ as $\Upsilon^+_\mu$ as a short
hand.
The structure factor, $S_i(\mathbf{q_\perp})$, can be used to determine the underlying symmetry
of the vortex lattice. Square and hexagonal vortex structures will manifest themselves as four or
six sharp Bragg peaks in reciprocal space. In a vortex liquid phase one expects a completely
isotropic structure factor.
The structure factor is defined as the Fourier transform of the longitudinally averaged
vortex density, $\langle n_i(\mathbf{r}_\perp)\rangle$, which is subsequently thermally
averaged,
\begin{equation}
S_i(\mathbf{q_\perp})=\frac{1}{L_xL_yf}\bigg\langle\bigg|\sum_{\rvec_\perp}
n_i(\rvec_\perp)\text{e}^{-\mathrm{i}\mathbf{r}_\perp\cdot\mathbf{q_\perp}}\bigg|\bigg\rangle.
\end{equation}
Here $n_i(\rvec_\perp)$ is the denisity of vortices of color $i$ averaged over the $z$-direction
\begin{equation}
n_i(\rvec_\perp)=\frac{1}{L_z}\sum_z n_i(\rvec_\perp,z),
\label{eq:vdens}
\end{equation}
and $\rvec_\perp$ is $\rvec$ projected onto a layer of the system with a given $z$-coordinate. The
vortex density is calculated by traversing each plaquette of the lattice, adding the factor
$\chi_{i,\rvec}^\mu$ of each link. Each time we have to add (or subtract) a factor of $2\pi$ in
order to bring this sum back into the primary interval of $(-\pi,\pi]$ a vortex of color $i$
and charge +1(-1) is added to this plaquette.
In addition to the structure factor, we look at thermally averaged vortex densities, $\langle
n_i(\mathbf{r}_\perp)\rangle$, as well as thermally and longitudinally averaged amplitude
densities, $\langle\left|\psi_i\right|^2(\mathbf{r}_\perp)\rangle$, defined similarly to
Eq.~\eqref{eq:vdens},
\begin{equation}
\left|\psi_i\right|^2(\mathbf{r}_\perp)=\frac{1}{L_z}\sum_z
\left|\psi_i\right|^2(\rvec_\perp,z).
\label{eq:pdens}
\end{equation}
This provides an overview of the real space configuration of the system.
When including amplitude fluctuations, which, when the potential term is disregarded, are unbounded
from above, it is of great importance to make sure all energetically allowed configurations are
included. To this end, we measured the probability distribution of $\left|\psi_i\right|^2$,
$P(\left|\psi_i\right|^2)$ during the simulations by making a histogram of all field configurations
at each measure step, and normalizing its underlying area to unity in post-processing.
The uniform rotation applied to the condensates is implemented in the Landau gauge:
\begin{equation}
\mathbf{A}=(0,2\pi fx,0),
\end{equation}
where $f$ is the density of vortices in a single layer. Note that this implies a constraint $L f \in
(1, 2, 3, \ldots)$ due to the periodic boundary conditions. When probing a first order melting
transition, it is important to choose a filling fraction large enough that an anomaly in the
specific heat is detectable. However, if the filling fraction is too large, one may transition
directly from a vortex liquid into a pinned solid, completely missing the {\em floating solid}
\/phase of interest. This scenario is characterized by a sharp jump in not only the longitudinal,
but also the transverse helicity modulus.\cite{Teitel94,Wheatley94} One must therefore chose $f$
small enough, to assure that the vortex line lattice is in a floating solid phase when it melts.
\section{Details of the Monte Carlo simulations}
\label{sec:MCdetails}
The simulations were performed using the Metropolis-Hastings
algorithm.\cite{Metropolis53,Hastings70} Phase angles were defined as $\theta\in(-\pi,\pi]$, and
amplitudes as $\left|\psi\right|^2\in(0,1+\delta\psi]$. The choice of $\delta\psi$ will be discussed
further, as it is important to ensure inclusion of the full spectrum of fluctuations. Both the
phases and the amplitudes were discretized to allow the use of tables for trigonometric and square root
functions in order to speed up computations. We typically simulated systems of size $L^3=64^3$, with
sizes up to $L^3=128^3$ used to resolve anomalies in the specific heat. We used $10^6$ Monte Carlo
sweeps per inverse temperature step, and up to $10^7$ close to the transition. $10^5$ additional
sweeps were typically used to thermalize the system. In the simulations, we examined time series of the internal
energies taken during both the thermalization runs and the measurements runs to make sure the
simulation converged. One sweep consists of picking a new random configuration for each of the four
field variables separately in succession, at each lattice site. Measurements were usually performed
with a period of 100 sweeps, in order to avoid correlations. Ferrenberg-Swendsen multi-histogram
reweighting was used to improve statistics around simulated data points, and jackknife estimates of
the errors are used.
Fig.~\ref{fig:2compprob} shows the probability distribution of the amplitudes,
$\mathcal{P}(\abs{\psi_i}^2)$. We get a peaked distribution for finite $\omega$. On the other hand,
when $\omega=0$, this is no longer the case. The distribution now approaches a uniform distribution
on the interval $(0,1]$. In this case the parameter $\eta$ serves to control the approach to
uniformity, $\eta\rightarrow\infty$ corresponding to the $CP^1$ limit.
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{psiprob2comp_2}
\caption{(Color online) The probability distribution of the amplitudes, $\mathcal{P}(\abs{\psi_i}^2)$, for $N=2$, at inverse temperature $\beta=1.20$,
$f=1/32$, and $\eta=2$, with $\omega$ values from $0$ to $3$. The distribution is completely symmetric
in $i$.}
\label{fig:2compprob}
\end{figure}
With these initial simulation runs as a basis, we choose $\delta\psi$ appropriately in order to
capture the entire spectrum of fluctuations.
\section{Results of the Monte Carlo simulations}
\label{sec:results}
In this section, the $\eta-\omega$ phase diagram of ground states is explored by slow cooling and
examination of vortex and amplitude densities, as well as structure factors. In addition to the
expected hexagonal and square vortex ground states, several interesting regions of the parameter
space are investigated further. A special case between the square and the hexagonal region of the
phase diagram is discovered, where the lattice first forms a square structure, but thermally
reconstructs into a hexagonal lattice as the temperature is decreased further. Furthermore, we
consider in detail the $\omega=0$ line in the phase diagram, where we discover additional vortex
fluctuation effects. For $\omega=0$, the system features an $\mathrm{SU(2)}$ symmetry. An unusual
feature is an interesting state with global phase coherence, but without a regular vortex lattice.
In this case ordinary vortices do not have topological character due to $\mathrm{SU}(2)$ symmetry.
Additionally, we obtain several interesting vortex structures characterized by dimer-like
configurations at lower temperatures. Here, we observe honeycomb lattices, or double-core lattices,
and stripe configurations, consistent with previous $T=0$ results.\cite{Kasamatsu2003}
We also examine the melting transitions of the square and hexagonal lattices with the full amplitude
distribution included, as well as the melting of the hexagonal lattice in a model with $N=1$ as a
benchmark of the method. To classify the transition, we look at thermal averages of the specific
heat, helicity moduli, and vortex structure factors. These results are presented in
Appendixes~\ref{sec:N1} and~\ref{sec:latticemelting}.
\subsection{The $\eta-\omega$ phase diagram}
\label{sec:nwphases}
Adding a second matter field and inter-component density-density interactions results in a
considerably richer set of ground states than in the single-component case. In the
absence of a fluctuating part of the rotational "gauge-field" there will be no gauge-field-mediated inter-component
current-current interactions. For $\eta - \omega < 0$, $(\eta,\omega) > 0)$ the effective
inter-component density-density coupling $\eta-\omega$ is negative and the ground state of each
color of condensate has a hexagonal symmetry, as shown in Fig.~\subref*{fig:hexplot}. If, on the other hand $\eta
-\omega > 0$, the inter-component coupling becomes positive. Now, for sufficiently large ratios
$\eta/\omega$, the vortices arrange themselves into two inter-penetrating square lattices, shown in
Fig. \subref*{fig:squareplot}. The value of the ratio $\eta/\omega$ for which the lattice
reconstructs depends on the strength of the rotation, $f$.
If we neglect fluctuations, $\eta - \omega <
0$ is expected to result in a hexagonal lattice, while $\eta-\omega > 0$ leads to a square lattice
for sufficiently large $\eta/\omega$.
The physics of the reconstruction of the lattice can be explained by modulations of the amplitude
fields. The existence of static periodic amplitude-modulations (density-variations) is due to the
presence of vortices. Without vortices ($f=0$) and $\omega>0$, the ground state is one where both
amplitudes are equal and smooth. Vortices in one component tend to suppress locally the
corresponding amplitude, which in turn means that the term $\eta(|\psi_1|^2 + |\psi_2|^2-1)^2$
enhances the amplitude of the other component. At small $\omega$, \textit{i.e.}\ large
$\eta-\omega$ there is a strong tendency to form a square density lattice due to this intercomponent
density-density interaction. Conversely, if $\omega$ is large enough compared to $\eta$, the
density-density interaction is not strong enough to overcome the isotropic current-current
interactions between same-species vortices. In other words if the current-current interactions
dominate the interspecies density-density interactions, a hexagonal lattice is energetically
favoured over a square lattice, and vice versa. Note that similarly a square vortex lattice forms in
two-component London models with dissipationless drag when there are competing inter- and
intra-species current-current vortex interactions.\cite{Dahl2008,Dahl2008_2}
Figures.~\subref*{fig:phasesf_32} and~\subref*{fig:phasesf_64} show
the phase diagrams for filling fractions $f=1/32$ and $f=1/64$, respectively. The separation line
is approximate and drawn from several separate simulations.
\begin{figure}
\subfloat[]{\includegraphics[width=\columnwidth]{matrix_w5_converted}
\label{fig:hexplot}}\\
\subfloat[]{\includegraphics[width=\columnwidth]{matrix_w05_converted}
\label{fig:squareplot}}
\caption{(Color online) Representative configurations of the two main ordered phases in the $\mathrm{U(1)}\times\mathrm{U(1)}$
region. (a) shows a square structure at $(\eta,\omega)=(5.0,0.5)$, while (b)
illustrates the hexagonal structure at $(\eta,\omega)=(5.0,5.0)$. Each subfigure shows vortex
densities, $\langle n_i(\mathbf{r}_\perp\rangle$, in the left column, amplitude densities,
$\langle\left|\psi_i\right|^2(\mathbf{r}_\perp\rangle$, in the right column, and structurue
factors (insets) of each component as indicated. The induced vortex density and inverse
temperature are fixed to $f=1/64$ and $\beta=1.5$ in both subfigures.}
\label{fig:square_hex}
\end{figure}
To clarify what is going in Figs.~\subref*{fig:phasesf_32} and~\subref*{fig:phasesf_64}, we refer to
Figs.~\ref{fig:tableaux_l5} and~\ref{fig:tableaux_l3} in Appendix~\ref{app:tableaux}. Here, we show
tableaus to illustrate in more detail how the density and vortex lattices reconstruct at a
temperature well below any melting temperatures of the vortex (and density) lattices, as the
density-density interaction $2(\eta-\omega) |\psi_1|^2 |\psi_2|^2$ is varied. Specifically, we fix
the interaction parameter $\eta$, as well as the inverse temperature $\beta$ and filling fraction
$f$, while increasing the parameter $\omega$. This reduces the effective inter-component
density-density interaction which favors a square lattice, until the lattice reconstructs from
square to hexagonal symmetry.
\begin{figure}
\centering
\subfloat{
\includegraphics[width=\columnwidth]{Phases_32}
\label{fig:phasesf_32}}
\subfloat{
\includegraphics[width=\columnwidth]{Phases_64}
\label{fig:phasesf_64}}
\caption{(Color online) The $\eta-\omega$ phase diagram of the ground states for $f=1/32$ (top) and
$f=1/64$ (bottom). The simulations were performed for a range of $(\eta,\omega)$ pairs to determine
the zero temperature ground state. Approximate demarcation lines for the phase boundaries separating
hexagonal lattices, square lattices, and dimerized phases, were drawn from these results (solid lines).
I denotes the phase where the hexagonal vortex lattices in the two components are cocentric, II
denotes the case where the hexagonal lattices are intercalated, while III denotes the square lattice
phase. The dotted line is the line $\omega = \eta$ at which the intercomponent density-density
interaction $2(\eta-\omega) |\psi_1|^2 |\psi_2|^2$ changes sign. See also Figs.~\ref{fig:tableaux_l5}
and~\ref{fig:tableaux_l3} in Appendix~\ref{app:tableaux}.}
\end{figure}
When $\eta=\omega$, it is seen from Eq.~\eqref{eq:inter_intra_g} that the two components of the order parameter
decouple. For $\omega<\eta$ the inter-component density-density interaction is repulsive, while it is
attractive for $\omega>\eta$. For $\omega < \eta$, the vortex lattices (and the density lattices) are
intercalated, while for $\omega > \eta$ they are cocentric. In Figs.~\ref{fig:phasesf_32} and
\ref{fig:phasesf_64} we illustrate the demarcation line between the two situations as a dotted
line in the hexagonal phase.
Beyond the square and hexagonal lattices we also observe dimer configurations of vortices for $\omega=0$,
which will be discussed further below. The calculations are consistent with the ground states
obtained in Refs. \citenumns{Kasamatsu2003} and \citenumns{makoto2}.
\subsection{Thermally induced reconstruction of vortex lattices}
Now we move to discussion of the effects of thermal fluctuations in these systems.
Fig.~\ref{fig:crossover} shows the vortex-densities in component $1$ in reciprocal space, as $\beta$
is increased, \textit{i.e.},\ as temperature is reduced, in a temperature range below where the
lattice melts. The actual melting of the two component lattice is discussed in
Section~\ref{sec:latticemelting}. We fix the filling fraction $f=1/64$, as well as the interaction
parameters $\eta=2$ and $\omega=0.5$.
For the highest temperatures shown in Fig.~\ref{fig:crossover} the vortex lattice is square.
Upon cooling the system, the vortex lattice reconstructs into a hexagonal lattice, consistent with
the ground state phase-diagram of Fig.~\subref*{fig:phasesf_64}. The density-density interaction
term $2(\eta-\omega) |\psi_1|^2 |\psi_2|^2$ aids formation of a square lattice at higher
temperatures, while the current-current interactions drives the lattice towards a hexagonal
configuration when it is cooled further. This means that the free energy per vortex of the square
lattice, which is lower than that of the hexagonal lattice at $\beta=0.90$, has become larger than
that of the hexagonal lattice when $\beta=1.50$. This is essentially the combination of an energetic
and an entropic effect. We observe this reconstruction not too far away from the demarcation line
separating a square and a hexagonal vortex lattice. Deep inside the hexagonal phase in
Fig.~\subref*{fig:phasesf_64}, we observe direct vortex lattice melting from a hexagonal
lattice to a vortex liquid. We note that intermediate entropically-stabilized vortex lattice
phases were of a subject of interesting investigation in the different system of $U(1)\times U(1)$
superconductors \cite{kivelson}, however the vortex interaction form is different in this case.
\begin{figure}
\centering
\subfloat[]{\includegraphics[width=0.5\columnwidth]{sfrecon_beta080}}
\subfloat[]{\includegraphics[width=0.5\columnwidth]{sfrecon_beta090}}\\
\subfloat[]{\includegraphics[width=0.5\columnwidth]{sfrecon_beta120}}
\subfloat[]{\includegraphics[width=0.5\columnwidth]{sfrecon_beta130}}\\
\subfloat[]{\includegraphics[width=0.5\columnwidth]{sfrecon_beta134}}
\subfloat[]{\includegraphics[width=0.5\columnwidth]{sfrecon_beta138}}
\caption{(Color online) Thermally induced reconstruction from a square vortex lattice in either of the components
at $\eta=2,\omega=0.5$, to a hexagonal vortex lattice, as $\beta$ is increased. Here, $f=1/64$.
(a)-(f) show inverse temperatures $\beta=\{0.80, 0.90, 1.20, 1.30, 1.34,
1.38\}$, respectively
Each subfigure shows $S_1(\mathbf{q}_\perp)$ only; $S_2(\mathbf{q}_\perp)$ is identical.
The physical reason for the reconstruction originates with the inter-component density-density
interaction term $2( \eta-\omega ) |\psi_1|^2 |\psi_2|^2)$, and is explained in detail in the
text.}
\label{fig:crossover}
\end{figure}
\subsection{$\mathrm{SU(2)}$ vortex states}
\label{sec:su2lattices}
The limit $\omega\rightarrow 0$ is quite different from the $\mathrm{U(1)} \times
\mathrm{U(1)}$-symmetric case $\omega \neq 0$. From Eq.~\eqref{eq:N2pot}, it is seen that the
Hamiltonian is invariant under $\mathrm{SU(2)}$ transformations of $\Psi$. Vortices, which are
topological in a $\mathrm{U(1)} \times \mathrm{U(1)}$ model, are no longer topological in the
$\mathrm{SU(2)}$ case. One may unwind a $2 \pi$ phase winding by entirely transferring density of
one component to the other, which may be done at zero energy cost.
\begin{figure}
\centering
\subfloat{
\includegraphics[width=\columnwidth]{SU2Helicity_sum}
\label{fig:su2heli}}\\
\subfloat{
\includegraphics[width=\columnwidth]{SU2vortexsmoothing}
\label{fig:su2vortsmooth}}
\caption{(Color online) Illustration of the observed state with coherence along the direction of the rotation axis
without a regular vortex lattice, seen only with $\mathrm{SU(2)}$ symmetry. The parameters used
are $\omega=0.0$, $\lambda=5$, and $f=1/64$. The top panel shows the helicity modulus of the phase
sum, $\Upsilon^+_\mu$. The two bottom panels show the vortex densities, $n_1(\rvec_\perp)$, at
$\beta=0.94$, where the $z$-directed modulus is clearly finite. The bottom left and bottom right panels are
taken from simulations using $10^6$ and $10^7$ Monte-Carlo sweeps, respectively. No apparent
vortex line structure is seen here, and by increasing the number of Monte-Carlo sweeps the
variations of the vortex density are smoothed out further. Note how the
value of the average vortex density seems to converge towards $1/64$.}
\label{fig:su2obs}
\end{figure}
Fig.~\ref{fig:su2obs} shows one of the main results of our paper. These are simulations with
$\mathrm{SU(2)}$ symmetry, \textit{i.e.}, $\omega=0$, as well as $\eta=5.0$ and $f=1/64$. The top
panel show the phase stiffness associated with the phase sum, $\Upsilon^+_\mu$. This is the
physically relevant phase variable in this case, as it couples to the rotation.
We observe that the
stiffness along the $z$-direction becomes finite at an inverse temperature, $\beta\sim 0.9$. This is
what one would expect when a vortex lattice forms. However, the bottom panels, which shows the vortex
density of component $1$ at $\beta=0.94$, shows no apparent signs of vortex ordering. Hence, we have
an unusual situation. There is a relatively large $\beta$-range where we have a finite $z$-directed
helicity modulus of the phase sum, but no apparent ordering of induced vortices.
A finite helicity modulus generally means that there are straight vortex lines with very little
transverse fluctuations threading the entire system along the direction in question. In the
$\mathrm{U(1)}$ picture this corresponds to a regular vortex lattice. For an
$\mathrm{SU(2)}$-condensate, this is no longer the case. Large relative amplitude fluctuations can
occur since they have zero energy cost in the ground state as the energy is no longer minimized by
a preferential value of $\left|\psi_1\right|^2-\left|\psi_2\right|^2$. This results in many
(nearly) degenerate vortex states between which the system can fluctuate, thus greatly simplifying
the effort of moving an entire, almost straight, vortex line. We are left with a phase where we have
coherence along the $z$-direction, but no regular vortex lattice appears in thermal averages. Nearly
straight vortex lines will shift between a large number of degenerate, or nearly degenerate, states
at a time scale shorter than a typical Monte-Carlo run.
The bottom panels of Fig.~\ref{fig:su2obs} show some
inhomogeneities of the vortex densities, exemplifying that this is not an ordinary vortex liquid
with segments of vortex lines executing transverse meanderings along their direction, which would
yield zero helicity modulus along the direction of the field-induced vortices. Rather, what we have
is a superposition of many lattice-like states of nearly straight vortex lines, where the
fluctuations are largely collective excitations of entire nearly straight lines, rather than
fluctuations of smaller segments of lines.
We emphasize again that these collective excitations
originate with large amplitude fluctuations due to the $SU(2)$-softness of the amplitudes of the
components of the superfluid order parameter, rather than with phase fluctuations. Increasing the
number of Monte-Carlo sweeps by an order of magnitude smooths these variations out (without
noticably altering the value of $\Upsilon^+_z$), as seen in the bottom right panel of
Fig.~\ref{fig:su2obs}. Note how the average value of the vortex density seems to
converge towards $1/64$. This is what we expect for a vortex lattice or liquid in a
$\mathrm{U}(1)\times\mathrm{U}(1)$ symmetric model, as the density of thermal
vortices will average to zero, and $f$ is the average flux density per plaquette.
\begin{figure}
\centering
\subfloat[]{
\includegraphics[width=\columnwidth]{SU2vortices_beta084_converted}
\label{fig:su2vortices1}}\\
\subfloat[]{
\includegraphics[width=\columnwidth]{SU2vortices_beta150_converted}
\label{fig:su2vortices2}}
\caption{(Color online) Two examples of $\mathrm{SU(2)}$ vortex configurations from a single simulation, for two different inverse
temperatures. The parameters $\eta$, $f$, and $\omega$ are fixed in each subfigure, at $\eta=1.0$,
$\omega=0.0$, and $f=1/64$. (a) shows $\beta=0.84$, while (b) shows
$\beta=1.50$. Each subfigure shows vortex densities, $\langle n_i(\mathbf{r}_\perp\rangle$, in
the left column, amplitude densities, $\langle\left|\psi_i\right|^2(\mathbf{r}_\perp\rangle$, in
the right column, and structure factors (insets) of each component as indicated. This
illustrates the degeneracy of the vortex line lattice in the isotropic limit, as the
configurations evolve when $\beta$ is varied. See appendix~\ref{app:tableaux} for more
details.}
\label{fig:su2vortices}
\end{figure}
As the system is cooled further, the movements of large vortex lines cease, and a regular vortex
lattice appears. However, degeneracy must still be present, as the exact pattern formed by the
lattice is distinctively different between simulations (keeping all parameters equal). The lattice
also has a tendency to shift between configurations as the temperature is varied, below the
temperature of initial vortex lattice formation. We observe two distinct classes of vortex states,
illustrated in Fig.~\ref{fig:su2vortices}. The two are stripes (Fig.~\subref*{fig:su2vortices1}) and
honeycomb lattices (Fig.~\subref*{fig:su2vortices2}), both of which are seen in
Ref.~\citenumns{Kasamatsu2003}. Note that these vortex densities are taken from a single simulation,
after the lattice has formed. {Within the accuracy of our simulations the obtained states are
not metastable. The evidence of this is obtained by performing several independent runs from
different initial configurations.} Again, we refer to Appendix~\ref{app:tableaux}, where
Fig.~\ref{fig:tableaux_su2} illustrates the degeneracy in the vortex line lattices obtained in the
isotropic limit in further detail.
\section{Experimental considerations}
\label{sec:exp}
Hexagonal and square lattices have already been observed in binary condensates of rubidium.~\cite{Cornell2004} However,
an $\mathrm{SU(2)}$ condensate has not been realized experimentally. In this section, we briefly outline under what
circumstances an observation of an $\mathrm{SU(2)}$ vortex state may be feasible.
In order to experimentally realize $\mathrm{SU(2)}$ conditions, one requires a two-component BEC,
where both intra- and inter-component interactions are equal. As we have seen, the $\mathrm{SU(2)}$
physics crucially depends on this, since even minor deviations from this condition immediately
yield $U(1) \times U(1)$ physics. This corresponds to $\omega=0$ in our parametrization.
Intra- and inter-component density-density interactions are given in terms of scattering lengths. While tuning
of these in an experiment is possible with Feshbach resonances, it may still be a challenge to tune
two scattering lengths independently to be equal to a third, to arrive at the $\mathrm{SU(2)}$ point.
From what is known for scattering lengths of real systems, it appears that a mixture of two species
of the same atom, but in different hyperfine states, lends itself more readily to a
realization of an $\mathrm{SU(2)}$ condensate than a mixture of different atoms or a mixture of
different isotopes of the same atom. This is so, since in the former case, the relevant scattering
lengths typically {\it a priori} \/are much more similar to each other than what they are in more
heterogeneous mixtures.
One promising candidate therefore appears to be a condensate of $^{87}\text{Rb}$ prepared in the two
hyperfine states $|F=1, m_f=1\rangle\equiv|1\rangle$ and $|F=2, m_f=-1\rangle\equiv|2\rangle$. In
this system, the three relevant $s$-wave scattering lengths already have values close to the point of
interest, $a_{11} = 100.4 a_B$, $a_{22} = 95.00a_B$, and $a_{12} = 97.66a_B$, where $a_B$ is the
Bohr radius.~\cite{Cornell98,Merril2007} Reference \citenumns{Tojo2010} reports on a \textit{magnetic}
Feshbach resonance at a field of approximately $9.1$ Gauss, where control of $a_{12}$ of the order
of $10a_B$ is possible. Additionally, Reference \citenumns{Thalhammer2005} reports on an \textit{optical}
Feshbach resonance of the state $|F=1, m_F=-1\rangle$, able to tune the intracomponent scattering
length, using two Raman lasers, with detuning parameters approximately given by $\Delta_1=2\pi\times75$MHz and
$\Delta_2=2\pi\times20$MHz. Here, varying $\Delta_2$ tunes the value of the scattering length around
the Feshbach resonance, while varying $\Delta_1$ changes the width. Hence, greater control of the
resonance is possible with an optical Feshbach-resonance compared to a magnetic one. Presumably, there
should exist optical Feshbach resonances able to tune the scattering length of either the $|1\rangle$
or the $|2\rangle$ state, for instance the one reported to exist at $1007$G for the $|1\rangle$
state.~\cite{Volz2003} This resonance should be far enough away from the inter-component resonance
at $9.1$G to not cause any interference.
This suggests one possible setup. Namely, prepare a two-component condensate of $^{87}$Rb in the
$|1\rangle$ and $|2\rangle$ states under rotation, and tune $a_{12}$ to $a_{22}$ using a magnetic
field. Then, tune $a_{11}$ to the same value using optical techniques, while taking time-of-flight
images of the condensate. The prediction is that as the system is tuned through the optical Feshbach
resonance, one should observe a hexagonal composite vortex lattice at sub-resonance frequencies, the
non unique vortex ordering pattern, discussed above, at a frequency where all scattering lengths are equal, close to the
optical Feshbach resonance, and finally the reappearance of a hexagonal vortex lattice at
frequencies above the frequency where all lengths are equal Fig.~\ref{fig:su2vortices}. The
observation of a featureless rotating condensate would be a direct manifestation of the loss of
topological character of $U(1)$-vortices in the $\mathrm{SU(2)}$-symmetric case.
It would be interesting to study the dynamics of the vortex lattice in this case with methods like those used in
\cite{Freilich03092010}. For other discussions of SU($N$) models in cold atoms see Refs.
\citenumns{2014arXiv1403.2792C} and \citenumns{2010NatPh...6..289G}.
In actual experiments, a magnetic trap is used to confine the condensate in a given lateral region.
The effect of this on thermal fluctuations in vortex matter has been studied in detail in previous
theoretical works for the one-component case, without amplitude fluctuations \cite{PhysRevLett.97.170403,PhysRevA.77.043605}.
The effect of the trap is to yield a maximum overall condensate density at the center of the trap, while depleting it
towards the edge of the trap. As a result, the lattice melts more easily near the edge of the trap. As can be inferred from the work
on single-component melting \cite{PhysRevLett.97.170403,PhysRevA.77.043605}, the results of the present paper, where no
inhomogeneity due to a magnetic trap has been accounted for, is therefore most relevant to the region close to the center
of the trap.
\section{Conclusions}
\label{sec:conclusions}
In this paper, we have investigated a two-component $\mathrm{U}(1)\times\mathrm{U}(1)$ and
$\mathrm{SU}(2)$ Bose-Einstein
condensate with density-density interaction under rotation at finite temperature, thereby extending
previous works which calculated the zero-temperature ground state numerically. In the
$\mathrm{U}(1)\times \mathrm{U}(1)$ case we
report that thermal fluctuations can lead to a phase transition between hexagonal and square vortex
lattices with increased temperature.
In the isotropic, $\mathrm{SU(2)}$, limit, we have observed an intermediate state of global phase
coherence without an accompanying vortex lattice in the thermally averaged measurements. In
addition, we observe a variety of dimerized vortex states, such as dimerized stripes and
honeycomb-like lattices, which exist for a wide range of temperature. These lattices could be
observed in binary Bose-Einstein condensates in two separate hyperfine states, by precisely tuning the inter-
and intra-component scattering lengths to the $\mathrm{SU(2)}$ point through the use of Feshbach
resonances.
\begin{acknowledgments}
We thank Erich Mueller for useful discussions. P.~N.~G. thanks NTNU and the Norwegian Research
Council for financial support. E.~B. was supported by by the Knut and Alice Wallenberg Foundation
through a Royal Swedish Academy of Sciences Fellowship, by the Swedish Research Council Grants No.
642-2013-7837 and No. 325-2009-7664, and by the National Science Foundation under the CAREER Award
No. DMR-0955902, A.~S. was supported by the Research Council of Norway, through Grants No. 205591/V20 and
No. 216700/F20. This work was also supported through the Norwegian consortium for high-performance
computing (NOTUR).
\end{acknowledgments}
|
\section{Introduction}
\label{sec:introduction}
Ground-based observations and, in particular, spectroscopy can be
strongly affected by absorption caused by molecules in the Earth's
atmosphere. The wavelength ranges involved include, but are not
limited to, the near- and mid-infrared.
It is a standard practice to
correct as much as possible for such telluric absorption lines in order to recover
the spectrum (or photometry) of the target as it would
appear if the instrument was located above the atmosphere.
Different approaches have been proposed.
A very common method is to observe either a rapidly
rotating early-type or a G-type
star (depending on the specific purpose) soon before or after the
observations of the science target. This \textit{\emph{telluric}} star should
ideally be located close to the science target in the sky. Indeed,
early type stars are advantageous because they are mostly featureless,
except for hydrogen lines. Spectral regions close in wavelength to the
one of hydrogen lines, on the other hand, are therefore best corrected
by G-type stars. The reason to observe both science target and
telluric star close in time is to eliminate
differences caused by temporal variation in the atmosphere as much as the possible. On the other hand,
the need to observe both objects angularly close on the sky arises
because saturated and unsaturared telluric absorption lines do not
scale with the same function of airmass. In addition, the distribution
of molecules in the atmosphere (mainly water vapour) can depend on
the pointing direction.
Observations of such telluric standard stars require telescope
time that could otherwise be dedicated to other scientific
observations. This is a necessary but costly practice, especially
in the coming era of extremely large telescopes. Whenever a limited
time window of observability is available, or in the case of
high-cadence time series, observations of telluric stars further
reduce, sometimes considerably, the amount of time that can be
dedicated to the main science targets.
The quality of the correction can be variable and depends considerably on the
difference in airmass between the science and telluric star
observations, because in the best conditions, a difference of 1\% in
airmass already introduces a 1\% change in the optical depth of
unsaturated telluric lines. Examples of other problems
affecting observations of telluric stars are that
(i) suitable telluric stars that are close enough to the science
target may not be available; (ii) atmospheric conditions
may change by the end of the science exposures;
(iii) the signal-to-noise ratio of the obtained
telluric spectrum may be insufficient, either in comparison
with the quality of the science observations or with respect to
one's scientific aims; (iv) post-observation
analysis of the spectrum of the telluric star reveals that
it does not have the expected spectral type or
characteristics. Critically, telluric star observations may
be missing as a result of inclement weather, instrumental issues,
human errors, and/or incomplete calibration plan. The use of
archival data provides a particularly illustrative example
since observations of a telluric star may be missing or not
appropriate for the science goal pursued by the archive user as it may
significantly differ from the one originally intended.
Alternatively, the emergence of freely available radiative transfer
codes for atmospheric research provides the possibility of using
synthetic transmission spectra instead of telluric standard stars. One
of the first articles to demonstrate the ability of this approach is
\citetads{2007PASP..119..228B}, who used the radiative transfer model
SMART \citepads{1996JGR...101.4595M} for that purpose.
\citetads{2010A&A...524A..11S} refined the method incorporating the
radiative transfer code LBLRTM \citep{CLO05}, the line database HITRAN
\citep{ROT09}, a combination of meteorological data from various
sources to achieve synthetic transmission curves, and a model of the
line spread function. These were then used to successfully
perform telluric absorption corrections of CRIRES
\citepads{2004SPIE.5492.1218K} spectra up to an accuracy of
about 2 per cent. The code LBLRTM is also used by
\citetads{2013arXiv1312.2450H} who incorporated it into their own
code to fit an entire input spectrum.
A different approach is used by
\citetads{2014MNRAS.439..387C}. Since they investigate solar system
planetary atmospheres containing similar molecules to those in the Earth's
atmosphere, they derive the state of the latter by fitting spectral
features of telluric standard star observations. \cite{GAR13} used the
radiative transfer model\footnote{RTM is part of the SCIATRAN 2.2
software package, developed by Institute of Remote Sensing/Institute
of Environmental Physics of the University of Bremen, Germany}
developed for the SCIAMACHY instrument onboard the ENVISAT satellite
to fit water features visible in spectra taken at the Sierra Nevada
Observatory. In addition, they used it to determine the amount of
precipitable water vapour in the Earth's atmosphere at the time of
observation.
Unfortunately, the limited quality of these
corrections, the specialised design and/or restricted
availability has so far impeded any wide-spread use of such tools.
The deviation of the modelled absorption line from the observed one
can be caused by errors in the radiative transfer modelling, poor
representation of the line spread function, and/or limited
accuracy of the meteorological data. In practice, therefore, a
general tool allowing the use of simulated atmosphere transmission
spectra to correct for Earth's atmosphere telluric lines in a wide
variety of data is simply not available to the astronomical
community at large.
Here we describe\texttt{ molecfit}, a tool for modelling telluric
lines. \texttt{ Molecfit} retrieves the most appropriate atmospheric
profile (i.e., the variation in the temperature, pressure, and humidity
as a function of altitude) for the time of the given science
observations. It uses a state-of-the-art radiative transfer modelling
of the Earth's atmosphere, together with a comprehensive database of
molecular parameters. Finally, it allows for fitting
an analytical line spread function, as well as using a
user-provided numerical kernel. \texttt{ Molecfit} can be used from
0.3 to 30 $\mu$m (or more) and can be configured to
apply on data collected by a wide-range of optical and infrared
instruments from most observatories, making it very versatile.
The initial development of \texttt{ molecfit} was carried out mostly
as a set of IDL routines used as drivers for the Reference Forward
Model\footnote{\url{http://www.atm.ox.ac.uk/RFM/}} (hereafter, RFM)
radiative transfer
code. The initial goal was to measure the amount of
precipitable water vapour towards zenith (hereafter, PWV)
using the $\approx$ 19.5 $\mu$m emission line with VISIR, the mid-infrared
spectrograph and imager at the VLT \citepads{2008eic..work..433S}.
The code was then slightly modified to measure the amount of PWV using
sky emission line spectra obtained with the cryogenic high-resolution
infrared echelle spectrograph CRIRES, since the 5.038 to 5.063 $\mu$m
range is purely dominated by water vapour\footnote{These VISIR and
CRIRES measurements, together with UVES and X-shooter measurements
obtained using a different method, are available at
\url{http://www.eso.org/observing/dfo/quality/GENERAL/PWV/HEALTH/trend_report_ambient_PWV_closeup_HC.html}.}.
The good quality of the fits indicated that generalising the code to
absorption lines and other spectral ranges was promising and led to a
well-developed IDL prototype \citepads{2010HiA....15..533S}. Its
performances were similar to the one created by
\citetads{2010A&A...524A..11S} that only focused on CRIRES spectra.
This prototype was then ported in a robust way to CPL\footnote{The
Common Pipeline Library
\url{http://www.eso.org/sci/software/cpl/documentation.html} based
on ANSI C.} and further developed and optimised as part of the
Austrian in-kind contribution for joining ESO. In the process, the
LNFL/LBLRTM code \citep{CLO05} was chosen instead of RFM,
because it (i) was more frequently updated at the time of the code
selection, (ii) is used by a large community, (iii) incorporates a
code-optimised and up-to-date line list, and (iv) comprises broader
functionality (e.g., UV/optical ozone absorption, which was needed
in a parallel project \citepads{2012A&A...543A..92N}) and performs significantly
faster for most spectral ranges for similar or better precision.
This series of papers describes the results of these efforts.
In the present work, we describe the key ingredients of
\texttt{molecfit}, its applicability, and limitations, and we show a
variety of applications for ESO/VLT data covering a range of spectral
resolving power and wavelength regions. In a separate paper,
\citep[][Paper II]{KAUSCH14}, we have systematically investigated
the quality of the telluric line absorption correction by applying
it to the large sample of archival VLT/X-shooter near-infrared data,
and compared it with the standard method performed with the IRAF {\tt
telluric}\footnote{\url{http://iraf.net/irafhelp.php?val=telluric&help=Help+Page}}
task.
This paper is organised as follows. Section~\ref{sec:atmoabsorb}
describes the most important aspects regarding the absorption lines
formed in the Earth's atmosphere. Section~\ref{sec:description}
provides a description of the \texttt{molecfit} package.
A few characteristic cases are described in
Sect.~\ref{sec:usage}. Limitations to the use of the tool are
summarised in Sect.~\ref{sec:limitations}. Examples of successful
applications of \texttt{molecfit} on spectra obtained with several ESO
instruments are given in Sect.~\ref{sec:examples}. The most
important aspects of \texttt{molecfit} are summarised in the conclusion.
The technical aspects of \texttt{molecfit}
and additional user guidance, including a description of the
Graphical User Interface (GUI) that allows for
data interaction, are provided in the User Manual \citep{NOL13},
(hereafter, UM).
\section{Absorption arising in the Earth's atmosphere}
\label{sec:atmoabsorb}
The Earth's atmosphere consists of about $78\%$ of N$_2$, $21\%$ of
O$_2$, $1\%$ of Ar, and several trace gases and aerosols. Each of the
molecules and aerosols affects the light travelling through in
different wavelength regimes by absorption and
scattering. Figure~\ref{fig:trans} shows a model transmission curve
derived with the sky model developed in a parallel project
\citepads{2012A&A...543A..92N}. The main absorption regions of the eight major
contributing molecular species O$_2$, O$_3$, H$_2$O, CO, CO$_2$,
CH$_4$, OCS, and N$_2$O are marked. Thanks to the complexity of the
molecules, several ro-vibrational bands are clearly visible.
The optical regime is dominated by broad absorption bands from ozone
(Huggins bands shortwards of $\lambda \sim 0.4 \, \mu\mathrm{m}$,
\citetads{1890SidM....9..318H}; Chappuis bands at $\sim 0.5 <
\lambda/\mu\mathrm{m} < 0.7 $, \citetgallica{bpt6k30485}), narrow
oxygen bands (the prominent A, B and the weaker $\gamma$ bands at
$\sim 0.759 < \lambda/\mu\mathrm{m} < 0.772$, $ 0.686 <
\lambda/\mu\mathrm{m} < 0.695$, and $0.628 < \lambda/\mu\mathrm{m} <
0.634 $, respectively), and some comparably weak water vapour
features. The latter become dominant in the entire infrared regime
longwards of the $I$ band. In addition in the $JHK$ regime, the
transmission is significantly affected by CO$_2$, CH$_4$, and O$_2$.
Between the $K$ band and $\sim18\,\mu\mathrm{m}$ absorptions are mainly
caused by H$_2$O and CO$_2$, with some contribution from N$_2$O,
CH$_4$, and minor absorption features by CO and OCS. In the regime
from 18 to 30\,$\mu$m only water vapour is dominant. Over the whole
wavelength range, the gases listed in Table~\ref{tab:tracegases} lead
to absorption features with optical depth reaching up to 5\% at the
given spectra resolving power ($R \sim 10\,000$).
The Earth's atmosphere is highly variable on several time scales in
temperature, pressure, and chemical composition. The most variable
relevant molecule is water vapour, which is directly related to the
actual weather conditions at the time of observations. However,
seasonal, daily, and sometimes hourly variations in the volume mixing ratio
of various gases have been measured, in particular, for CO$_2$
\citepdoi{10.1029/JD094iD06p08549}, CH$_4$
\citepdoi{10.5194/acp-9-443-2009}, and O$_3$
\citepdoi{10.1038/172633a0}. In addition, following the World
Meteorological Organization\footnote{\url{http://www.wmo.int/gaw}}, the
fraction of CO$_2$ has increased by $10-15 $, CH$_4$ by $\sim10$, and
N$_2$O by $\sim8$ since 1985
\citepwmo{GHG_Bulletin_No.8_en.pdf} leading to global warming. Daily
columns for a number of molecules (O$_3$, N$_2$O, CH$_2$, etc.)
over most of the Earth surface can be obtained from \url{http://www.temis.nl/}.
\begin{figure*}
\centering
\includegraphics[clip=true]{transmission}
\caption[]{Synthetic absorption spectrum of the sky between 0.3 and
30~$\mu$m calculated with LBLRTM (resolution $R \sim 10\,000$) using the
annual mean profile for Cerro Paranal \citep{2012A&A...543A..92N}. The eight main
molecules O$_2$, O$_3$, H$_2$O, CO, CO$_2$, CH$_4$, OCS, and N$_2$O
contribute more than 5\% to the absorption in some wavelength
regimes. The red regions mark the ranges where they mainly affect
the transmission, minor contributions of these molecules are not
shown. The green regions denote minor contributions (see
Table~\ref{tab:tracegases}) from the following molecules: (1) NO,
(2) HNO$_3$, (3) COF$_2$, (4) H$_2$O$_2$, (5) HCN, (6) NH$_3$, (7)
NO$_2$, (8) N$_2$, (9) C$_2$H$_2$, (10), C$_2$H$_6$, and (11)
SO$_2$.}
\label{fig:trans}
\end{figure*}
\begin{table}
\caption[]{Species with a minor line-absorption contribution. The
specific spectra are calculated with the radiative transfer code
LBLRTM (resolution $R \sim 10\,000$) without continua to extract only the line contribution. The wavelength ranges are also shown in Fig.~\ref{fig:trans}.}
\centering
\vspace{5pt}
\begin{tabular}{c c c c}
\hline\hline
\noalign{\smallskip}
Species & Number in & Wavelength range & Optical\\
& Fig.~\ref{fig:trans} & [$\mu$m] & depth [\%]\\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
NO & 1 & 5.153--5.556 & $\la$0.3 \\
HNO$_3$ & 2 & 5.743--5.983 & $\la$1 \\
& & 7.404--7.825 & $\la$2 \\
& & 10.890--11.694 & $\la$1 \\
& & 20.452--23.478 & $\la$3 \\
COF$_2$ & 3 & 5.103--5.187 & $\la$0.01\\
& & 7.935--8.157 & $\la$0.01\\
H$_2$O$_2$ & 4 & > 19.705 & $\la$0.3\\
HCN & 5 & 2.959--3.092 & $\la$1\\
& & 6.801--7.370 & $\la$1\\
& & 12.819--15.467 & $\la$4\\
NH$_3$ & 6 & 5.755--6.530 & $\la$0.01\\
& & 8.904--12.571 & $\la$1\\
NO$_2$ & 7 & 3.412--3.487 & $\la$0.1\\
& & 6.051--6.383 & $\la$4\\
N$_2$ & 8 & 3.966--4.650 & $\la$2\\
C$_2$H$_2$ & 9 & 12.957--14.523 & $\la$0.1\\
C$_2$H$_6$ & 10 & 3.332--3.364 & $\la$2\\
SO$_2$ & 11 & 7.243--7.455 & $\la$0.01\\
\noalign{\smallskip}
\hline
\end{tabular}
\label{tab:tracegases}
\end{table}
\section{Description of the \texttt{molecfit} telluric correction tool}
\label{sec:description}
The \texttt{molecfit} package allows one to simulate and fit
tropospheric and stratospheric emission or absorption telluric lines
affecting user-selected region(s) of an observed (usually, science)
spectrum. It uses parameters provided in a configuration file and, if
present, the FITS header of the spectrum. It then returns a
transmission spectrum for the whole wavelength range of the spectrum.
\texttt{Molecfit} includes the following ingredients:
\begin{enumerate}
\item an automatic retrieval of two atmospheric profiles from the
Global Data Assimilation System (GDAS) web site.
These profiles bracket the time of the observation and can
be retrieved for any observatory in the world. These profiles
are linearly interpolated to match the
time of the observations and merged with a standard profile and
local meteorological data to create an input profile for the
radiative transfer code;
\item a radiative transfer code that simulates atmospheric emission and
transmission spectra;
\item a molecular spectroscopic database;
\item a simple grey-body calculator that simulates the contribution
of the continuum from the instrument and telescope;
\item a choice of possible line spread functions (also referred
to as \emph{\emph{kernels}});
\item an automatised fitting algorithm that adjusts the continuum
spectrum, wavelength calibration, spectral resolution, and column
density of each of the relevant molecules, with the possibility of
excluding spectral regions expressed either in pixels or in wavelength
regions.
\end{enumerate}
In the following, we describe important aspects of these components
in more detail. We also discuss specific choices made for the ESO
Paranal observatory.
\subsection{Atmospheric profile}\label{sec:atmprofile}
\begin{figure}[h]
\includegraphics[width=\columnwidth]{artprof.pdf}
\includegraphics[width=\columnwidth]{artprof_spec.pdf}
\caption{\textit{Top:} Two examples of the distribution of the water-vapour
volume-mixing ratio as a function of altitude. The black solid line
corresponds to a volume-mixing ratio of 0\,ppmv up to 3.1\,km
(Paranal altitude is 2.6\,km), while the red dashed line corresponds
to a constant relative humidity until 3.1\,km. The total amount of
precipitable water vapour, 2\,mm, is identical for both profiles.
\textit{Bottom:} sample of the corresponding transmission spectra at
$R = 100\,000$. The higher amount of water vapour in the lower
atmospheric layers above the observatory leads to both deeper cores
and stronger continuum absorption. }
\label{fig:pwv_artprof}
\end{figure}
An atmospheric profile typically describes the temperature $T$,
pressure $P$, and volume mixing ratio $x$ of several molecular species
as a function of altitude $h$ for a given location. These parameters
have a direct impact on the shape and strength of all the telluric
lines. For \texttt{{\tt molecfit}} we use a merged profile based on a
reference atmospheric profile, modelled 3-D data, and on-site
meteorological measurements.
The reference atmosphere profile is by default the equatorial
profile derived by J.
Remedios\footnote{\url{http://www.atm.ox.ac.uk/RFM/atm/}}. It contains
abundances of several molecular species up to an altitude of 120\,km,
divided into 121 altitude levels. Owing to the low latitude of Cerro
Paranal (Lat: -24.6$^\circ$), the equatorial profile is better suited
(Anu Dudhia, private communication) than the tropical or standard
profiles, which are also available in the molecfit package. We note here that
the molecular composition has changed since the profile was created
in 2001. In particular, the content of the greenhouse-gas carbon-dioxide content has increased by $\sim6\%$
\citepwmo{GHG_Bulletin_No.8_en.pdf}. \texttt{Molecfit} allows the user
to either fit or manually adjust the molecular content to take this
increase into account. Alternatively, the reference atmosphere used
by \texttt{molecfit} can be easily modified to another existing or
manually created profile.
The reference atmosphere is then merged with the modelled
GDAS\footnote{\texttt{http://www.ready.noaa.gov/gdas1.php}} 3-D data
provided by the National Oceanic and Atmospheric
Administration\footnote{\texttt{https://ready.arl.noaa.gov/gdas1.php}}
(NOAA). This model contains time-dependent information on the
temperature, pressure, relative humidity, wind speed, and direction up
to an altitude of $\sim26$\,km on a 3 h basis. Owing to the geographical
1$^\circ\times1^\circ$ grid, we created a set of profiles by means of
an interpolation of the four grid points closest to Cerro Paranal to
obtain local information. The GDAS profile derived for a specific
observation is then by default a linear interpolation of the two
resulting profiles closest in time to the observation. The {\tt
molecfit} package contains all the available GDAS profiles for Cerro
Paranal since December 1, 2004 at 00:00 UT and the release date of the
package. In addition, ESO will regularly maintain such a database for
Paranal. For observations obtained later or from another site, {\tt
molecfit} automatically retrieves the appropriate GDAS files.
It may still be possible that no GDAS profiles are available for the
time of observations. In the case of Cerro Paranal, \texttt{molecfit}
then use the best matching two-month average \citep{2012A&A...543A..92N}.
Alternatively, an ASCII profile with the same format as a GDAS profile
can also be provided by the user.
A further optional refinement of the final profile is achieved by
incorporating on-site measurements
obtained by the Astronomical Site Monitor (ASM)
as provided in the FITS header or through a parameter file. The
influence of the measured values at the observatory altitude is
gradually decreased up to a configurable upper mixing height,
set by default to $\sim5\,$km, where, for Paranal, the wind direction
changes by an average of $\sim90^\circ$ on average, as
revealed by an analysis of the GDAS wind data. Thus, it can be assumed
that at this altitude the influence of the local environment (as
determined from the ASM data) has diminished. More information on the
merging procedure can be found in the UM.
Water vapour is the molecule showing the most variations
because its total amount (PWV) and its altitude
distribution can change significantly and on a short time scale. In
particular, the largest amount of PWV may not be located in the lowest
layers of the atmosphere.
Figure~\ref{fig:pwv_artprof} illustrates
the importance of having a decent profile for water vapour.
The total amount of PWV can optionally be set if the value can
be retrieved independently (for example, from radiometer
\texttt{measurements}, see \citetads{2012SPIE.8446E..3NK}). In this case, the merged profile
composed by the reference atmosphere, the GDAS data, and the local
meteo information will be scaled to the requested PWV.
Alternatively, the atmospheric profile provided by the radiometer
-- soon retrievable from the ESO archive -- can be used directly.
\begin{figure}[h]
\includegraphics[width=\columnwidth]{H2O_temp11_Pfg.pdf}
\includegraphics[width=\columnwidth]{H2O_pres742p8_Pfg.pdf}
\caption{ Effect of ground temperature or pressure variation on the
H$_2$O telluric spectrum for two different spectral ranges. Each
graph shows the ratio between telluric spectra computed with the
indicated ground temperature or pressure. \textit{Top: }
variation of 1\,$\degr$C and 10\,$\degr$C. The region
close to 3.68\,$\mu$m shows opposite behaviours. \textit{Bottom:}
variation of 1~mb and 10~mb.}
\label{fig:gtemppres}
\end{figure}
Figures~\ref{fig:gtemppres} illustrates the effect of the
variation of the ground temperature and pressure on the
H$_2$O telluric spectrum. Depending on line parameters, different
lines behave differently with temperature or pressure.
\subsection{Radiative transfer code}\label{sec:lblrtm}
The resulting merged atmospheric profile is used as input in the
radiative transfer code LNFL/LBLRTM \citep{CLO05}. The
version used at the time of the first release of the
\texttt{molecfit} package is the latest
version at that time, V12.2. This code package is widely used in atmospheric
sciences and is publicly available. It calculates radiance and
transmission spectra with a spectral resolving power of about four million. More
information can be found at the web
site\footnote{\url{http://rtweb.aer.com/lblrtm_frame.html}}.
\subsection{Database of molecular parameters}
\label{sec:database}
In addition to the profile, the radiative transfer code requires a
line database as input. The first release of
\texttt{molecfit} uses the database \texttt{aer\_v\_3.2}, which
is delivered with the LNFL/LBLRTM code package: it is based on an
updated version of the HITRAN 2008 database \citep{ROT09} and contains
information on more than 2.7 million spectral lines of 42 molecular
species (see Table \ref{tab:molecs}). The \texttt{ molecfit} line
database can be easily updated when a new version becomes available.
Only molecules for which an atmospheric volume mixing ratio is
available in the standard atmosphere profile \texttt{provided with} \texttt{molecfit} can be
used by default. These molecules are identified in Table
\ref{tab:molecs}.
\begin{table}[!Ht]
\begin{center}
\caption{List of molecules as provided by the \texttt{ aer} line
parameter database.}
\vspace{6pt}
\centering
\begin{tabular}{r | c | c | c | c}
\hline\hline
\#
& Molecule
& Chemical name
& In standard
& GUI\\
&
&
& profile?
&\\
\hline
1 & H$_2$O & Water & X & f \\
2 & CO$_2$ & Carbon dioxide & X & f \\
3 & O$_3$ & Ozone & X & f \\
4 & N$_2$O & Nitrous oxide & X & f \\
5 & CO & Carbon monoxide & X & f \\
6 & CH$_4$ & Methane & X & f \\
7 & O$_2$ & Oxygen & X & f \\
8 & NO & Nitric oxide & X & c \\
9 & SO$_2$ & Sulfur dioxide & X & c \\
10 & NO$_2$ & Nitrogen dioxide & X & c \\
11 & NH$_3$ & Ammonia & X & c \\
12 & HNO$_3$ & Nitric acid & X & c \\
13 & OH & Hydroxyl & & \\
14 & HF & Hydrogen fluoride & & \\
15 & HCl & Hydrogen chloride & & \\
16 & HBr & Hydrobromic acid & & \\
17 & HI & Hydrogen iodide & & \\
18 & ClO & Chlorine monoxide & X & c \\
19 & OCS & Carbonyl sulfide & X & c \\
20 & H$_2$CO & Formaldehyde & & \\
21 & HOCl & Hypochlorous acid & X & c \\
22 & N$_2$ & Nitrogen & X & c \\
23 & HCN & Hydrogen cyanide & X & c \\
24 & CH$_3$Cl & Chloromethane & & \\
25 & H$_2$O$_2$ & Hydrogen peroxide & X & c \\
26 & C$_2$H$_2$ & Acetylene & X & c \\
27 & C$_2$H$_6$ & Ethane & X & c \\
28 & PH$_3$ & Phosphine & & \\
29 & COF$_2$ & Carbonyl fluoride & X & c \\
30 & SF$_6$ & Sulfur hexafluoride & X & c \\
31 & H$_2$S & Hydrogen sulfide & & \\
32 & HCOOH & Formic acid & & \\
33 & HO$_2$ & Hydroperoxyl & & \\
34 & O & Oxygen & & \\
35 & ClONO$_2$ & Chlorine nitrate & X & c \\
36 & NO+ & Nitrosonium & & \\
37 & HOBr & Hypobromous acid & & \\
38 & C$_2$H$_4$ & Ethylene & & \\
39 & CH$_3$OH & Methanol & & \\
40 & BrO & Bromine Monoxide & & \\
41 & C$_3$H$_8$ & Propane & & \\
42 & C$_2$N$_2$ & Cyanogen & & \\
\hline
\label{tab:molecs}
\end{tabular}
\tablefoot{
Col. 1: reference
number of the molecule as defined in the HITRAN 2008 database; Col. 2: its chemical formula, followed by its chemical
name. Col. 3 indicates if an atmospheric profile of the
molecular volume mixing ratio is available in the equatorial standard profile
as included in\texttt{ molecfit}. Last column:
functionality available through the GUI: \textit{f} the
corresponding column density can be fitted, \textit{c} the
corresponding transmission spectrum is only calculated using the user
provided abundance, relative to the one fixed in the reference
atmosphere.}
\end{center}
\end{table}
\subsection{Fitting algorithm}
\label{sec:fittinglagorithm}
\texttt{Molecfit} makes use of the C version of the least-squares fitting
library MPFIT \citepads{2009ASPC..411..251M} based on the FORTRAN fitting routine
MINPACK-1 \citepdoi{BFb0067700}. It uses the Levenberg-Marquardt
technique to solve the least-squares problem leading to a fast
convergence even for the complicated functions involved in
\texttt{molecfit}, while being numerically robust and self-contained
and allowing upper and lower limits or
parameters tied to each other that can be individually set.
\subsection{Inclusion and exclusion regions}
\label{sec:IncludingAndExcludingRegions}
A spectrum shows a number of features that can be
characterised in the following way:
\begin{itemize}
\item intrinsic features from the science object,
\item telluric features,
\item intrinsic features from the instrument caused either by the
spectrograph or the detector.
\end{itemize}
Intrinsic features either from the science object or from the instrument can
interfere with the fitting process.
Therefore, \texttt{molecfit} allows one to:
\begin{itemize}
\item select inclusion regions, also called \emph{\emph{fitting ranges}}, whose main
purposes are to allow\texttt{ molecfit} to determine the column density
of the relevant molecules and the line spread function
(or kernel). Such regions must therefore show a good enough
continuum level for the fitting accuracy and include telluric absorption lines
for determining the line spread function. One or several
such regions must be defined so that at least some telluric lines from each
molecule to be corrected are covered.
\item select exclusion regions within the inclusion regions that are
affected by the presence of intrinsic features from the science
objects;
\item mask pixels in the inclusion regions that are affected purely
by instrumental or detector effects.
\end{itemize}
Details can be found in the UM.
\subsection{Continuum contribution}
\label{sec:continuum}
The model spectrum, $F_\mathrm{out}$, is scaled by a polynomial of degree $n_\mathrm{c}$:
\begin{equation}
F_\mathrm{out}(\lambda) =
F_\mathrm{in}(\lambda) \, \sum_{i = 0}^{n_\mathrm{c}} a_i \lambda^i.
\end{equation}
For $a_0 = 1$ and all other $a_i = 0$, the model spectrum remains
unchanged. This is the default configuration for the initial
coefficients.
The continuum correction is carried
out independently for each fitting range.
A fitting range (or the full spectrum) is split further
if it is distributed over more than one chip.
If the input spectrum is an emission line spectrum, an optional flux
conversion can be carried out in order to match the unit of the
synthetic spectrum. Further details are available in the UM.
\subsection{Telescope background}
\label{sec:backgrounds}
For sky emission modelling, the telescope background is assumed to be
a grey body (black body times emissivity). The parameters of this grey
body correspond to the telescope main mirror temperature and an
effective emissivity of the telescope and instrument.
\subsection{Wavelength calibration}
\label{sec:wavelength}
The science spectra to analyse are expected to have an accurate
wavelength calibration. However, small errors can affect
the determination of the line spread function and column densities negatively.
Therefore in each inclusion region, the synthetic spectrum being fitted can
be adjusted to match the input spectrum.
A Chebyshev polynomial of degree $n_\mathrm{w}$ is used for this purpose:
\begin{equation}
\label{eq:wavelength}
\lambda = \sum_{i = 0}^{n_\mathrm{w}} b_i t_i,
\end{equation}
where
\begin{equation}
t_i = \left\{
\begin{array}{ll}
1 & \textrm{for\ } i = 0 \\
\lambda & \textrm{for\ } i = 1 \\
2 \, \lambda^\prime \, t_{i-1} - t_{i-2} & \textrm{for\ } i \ge 2
\end{array} \right.
\end{equation}
where the wavelength range
is normalised so that $\lambda^\prime$ ranges from $-1$ to $1$
over the entire spectral range of the spectrum or, in the case of
multi-chip instruments, over the range covered by an individual chip.
The conversion of the wavelength grid to a fixed interval
results in coefficients $b_i$ independent of the wavelength range and
step size of the input spectrum. For $b_1 = 1$ and all other $b_i =
0$, the model spectrum remains unchanged. This is the default
configuration for the initial coefficients.
On the other hand, in some cases, one wishes to use the telluric lines
as references to improve an unsatisfactory initial wavelength
calibration. This is the case for CRIRES in particular. For such a
purpose, \texttt{molecfit} actually outputs the inverse of the
wavelength calibration described in Eq. \ref{eq:wavelength}, in the
sense that the input spectrum is then adjusted to match the synthetic
one. Such wavelength calibration is only meaningful when
the inclusion range covers the whole spectrum and is sampled well by
telluric lines.
\subsection{Line spread function}
\label{sec:kernel}
The model spectrum is convolved with up to three different profiles to
determine the shape of the line spread function. They are:
\begin{enumerate}
\item a simple boxcar
\begin{equation}
F_\mathrm{box}(\lambda) = \left\{ \begin{array}{ll}
1 & \textrm{for\ } -w_\mathrm{box}/2 \le \lambda \le w_\mathrm{box}/2 \\
0 & \textrm{for\ } \lambda < -w_\mathrm{box}/2 \ \mathrm{or} \lambda > w_\mathrm{box}/2
\end{array} \right.
,\end{equation}
which is adapted to the pixel scale and normalised to an integral of 1.
This kernel is particularly useful for objects fully covering the
entrance slit; and\\
\item a Gaussian
\begin{equation}
F_\mathrm{gauss}(\lambda) = \frac{1}{\sigma \sqrt{2 \pi}}
\exp{\Bigg(-\frac{\lambda^2}{2 \sigma^2}\Bigg)}
\end{equation}
centred on 0, where
\begin{equation}
\sigma = \frac{w_\mathrm{gauss}}{2 \sqrt{2 \ln{2}}}.
\end{equation}
The full width at half maximum (FWHM), $w_\mathrm{gauss}$, is given in
pixels. \\
\item a Lorentzian
\begin{equation}
F_\mathrm{lorentz}(\lambda) = \frac{1}{\pi} \,
\frac{\lambda}{\lambda^2 + (w_\mathrm{lorentz}/2)^2}
\end{equation}
centred on 0, where $w_\mathrm{lorentz}$ is the FWHM in pixels. Compared to a
Gaussian, the Lorentzian approaches the 0-level flux significantly more slowly.
\end{enumerate}
A kernel consisting of all three components is usually too complex
to be constrained by a fit. To avoid unrealistic best-fit FWHM, the user should
reduce the number of degrees of freedom by fixing individual fit
components. A FWHM of zero pixel corresponds to a unity convolution;
in other words, it does not change the input spectrum.
Alternatively, the user has the possibility of choosing
a single Voigt profile kernel directly, which is then calculated by an
approximate formula that takes the FWHM of Gaussian and Lorentzian as
inputs.
The user can also select a kernel whose size in pixels
linearly increases with wavelength. It is suitable for an
instrumental setup with nearly constant resolving power, fixed
wavelength step size, and a wide wavelength range, such as
cross-dispersed echelle spectrographs.
Finally, a user-defined kernel can be provided in the form of an ASCII
table of fixed kernel elements. This option overrules the
parameterised kernel discussed above.
An inappropriate kernel can cause deviations in the determination of
molecular column densities of more than 10\%.
\subsection{Molecular volume-mixing ratio}
\label{sec:columndensity}
The integrated volume mixing ratio, $X(h_0)$, of a molecule for a
column of air from the observer at an altitude $h_0$ to the
top of atmosphere can be derived by
\begin{equation}
X(h_0) = \frac{\int_{h_0}^{\infty}{\frac{x(h)\,P(h)}{T(h)}\,\mathrm{d}h}}
{\int_{h_0}^{\infty}{\frac{P(h)}{T(h)}\,\mathrm{d}h}}
\end{equation}
if the mixing ratio (or mole fraction) $x$, air pressure $P$, and temperature $T$
are given depending on altitude $h$ by an atmospheric profile. Volume-mixing
ratios are usually given in part per million in volume (ppmv).
For $\mathrm{H_2O}$, it is also common to provide the column height of
PWV in mm (also referred to as amount of PWV). It can be calculated by
\begin{equation}
\mathrm{PWV} = \frac{m_\mathrm{mol, H_2O}}{\rho_\mathrm{H_2O}\,R}
\int_{h_0}^{\infty}{\frac{x_\mathrm{H_2O}(h)\,P(h)}{T(h)}\,\mathrm{d}h},
\end{equation}
where the mole mass of water $m_\mathrm{mol,\,H_2O} = 0.0182$\,kg, the density
of liquid water $\rho_\mathrm{H_2O} \approx 10^3\,\mathrm{kg\,m}^{-3}$, and the
gas constant $R = 8.31446\,\mathrm{J\,mol^{-1}\,K^{-1}}$. The required units of
$x$, $P$, $T$, and $h$ are ppmv, Pa, K, and km, respectively. {\tt
Molecfit} outputs the column height for other gases using a similar formula.
Since the GDAS profiles and the ESO MeteoMonitor data provide an $\mathrm{H_2O}$
volume-mixing ratio as relative humidity $RH$ in percent, it has to be converted to
$x$. For this purpose, we have implemented the approximations for vapour
pressures of ice and supercooled water as function of temperature as described
by \citetdoi{10.1256/qj.04.94}. The minimum of
both then corresponds to the saturated vapour pressure $P_\mathrm{sat}$, which
is used to estimate the altitude-dependent $x$ in ppmv by
\begin{equation}
x(h) = 10^6 ~ \frac{P_\mathrm{sat}(T(h))}{P(h)}\,\frac{RH(h)}{100}.
\end{equation}
\texttt{Molecfit} allows the user to adjust this quantity based on the observed telluric
lines, and we note that the GUI only allows one to fit the column density for
the molecules indicated in the last column of Table \ref{tab:molecs}.
This adjustment is done by multiplying the overall atmospheric
profile by a single constant. For most molecules, the impact is
actually limited to the lowest atmospheric layers.
\subsection{Supported instruments}
\label{sec:supportedinstruments}
One-dimensional (1-D) spectra of any visible, near-, or mid-infrared
instrument could in principle be used by\texttt{ molecfit}. Details on
the required information can be obtained in the User Manual
\citep{NOL13}. Although the tool has been tailored to ESO-Paranal
instruments -- in particular, the GDAS profiles are available as a
tarball file only for this site -- there are no specific limitations
either in the terms of format or in geographical location for a
ground-based observatory.
\subsection{Inputs and outputs}
\label{sec:outputs}
\texttt{ Molecfit} accepts 1-D spectra provided as ASCII tables, FITS
tables, and images. The FITS tables can have several extensions
corresponding to different chips (e.g., CRIRES). Required keywords are
either taken from the FITS header or directly from the parameter
file. Finally, ASCII or FITS files for inclusion and exclusion regions
can be provided.
\texttt{ Molecfit} fits the synthetic spectrum in the inclusion regions
alone. These regions can cover just a small fraction of the
wavelength range or all of it. It returns the best-fit model only for the
inclusion regions in both an ASCII and a FITS table, and the best-fit
parameters and molecular column densities (including the PWV) in a
results ASCII file.
The \texttt{molecfit} package then outputs a transmission spectrum and the
division of the input spectrum by the transmission spectrum over
the full spectral range of the input spectrum.
Apart from results tables in ASCII and FITS format, the corrected spectrum is
written into a file in the same format as the input spectrum (see the UM for
more details)
In addition, \texttt{molecfit} can provide a new wavelength calibration
based on the telluric lines as described in Sect.~\ref{sec:wavelength}.
This two-step approach greatly improves the speed of the overall process.
\subsection{Expert mode}
\label{sec:expert}
In most cases \texttt{molecfit} is able to determine the different
user-selected parameters automatically with a minimum of
configuration. However, in a few cases, mostly for CRIRES, a much
greater flexibility is needed.
An expert mode is therefore also provided. It allows the user to fit
the spectra on each inclusion range individually and provides him/her
the access to the coefficients of the polynomials for the continuum
and wavelength correction. Moreover, the parameter file with the
best-fit values for all parameters is saved and can be used for
another iteration of molecfit.
\subsection{Fitting sky emission spectra}
\label{sec:fittingskyemission}
The tool is able to fit sky emission spectra for molecules and
physical processes handled by the radiative transfer code, with the
exclusion of lines produced by chemiluminescence (such as OH, see,
e.g., \citetads{2012A&A...543A..92N}). In other words \texttt{molecfit}
can in principle determine the line spread function and column densities
of the relevant molecules and therefore derive the transmission
spectrum of the atmosphere based the sky emission spectrum if
the spectral range is located in the thermal infrared.
\section{Impact on observing strategy}
\label{sec:usage}
Although powerful,\texttt{ molecfit} is not necessarily suited to
telluric corrections in all cases. In this section, we briefly discuss
a few situations to consider while deciding on the details of the
observing strategy for the absorption telluric correction.
\subsection{Main cases}
\label{sec:maincases}
Two main cases for the use of \texttt{molecfit} can be considered, depending
on whether the wavelength calibration of the spectrum is reliable and
accurate (root-mean-squared residuals $\approx 0.1$ pixel or better) or not.
In the case of accurate wavelength calibration, the main requirement
is that the spectrum to be corrected has at least one spectral region
that includes medium-strength molecular lines of the most variable
relevant species and that these lines are isolated enough so that
their column density and the line spread function can be determined.
If the wavelength calibration is not reliable but needs to be
corrected by\texttt{ molecfit} based on the telluric lines
themselves, a -- possibly delicate -- selection of inclusion and
exclusion regions must be carried out, such that only telluric lines
appear in the inclusion regions and that any intrinsic feature of the
object is masked out in the exclusion regions. In case the continuum
needs to be refined by \texttt{molecfit}, the fitting regions need to
present enough coverage to be able to determine either an overall
continuum or local continua.
\subsection{Simple continuum shape and small number of intrinsic features}
\label{sec:simple}
The easiest spectra that can be modelled by the tool are the ones for
objects with a well-defined and simple continuum and with limited
confusion between telluric and intrinsic features. In such
conditions, the tool can properly model the continuum spectrum of the
object, and the parameters of the telluric features can be well
constrained.
\subsection{Large number of intrinsic features}
\label{sec:numerousintrinsic}
If the number of intrinsic features is large in the spectral domain of
interest, several options are possible. The choice of the best option
depends on the line crowding: a first constraint is the number of
points that can be used to determine the spectral continuum; a second
constraint is the number of spectral ranges where telluric lines can
be used to determine the column density of the molecules contributing
in this spectral domain. An additional constraint arises if the
wavelength calibration is not reliable.
Therefore a possible strategy is first to observe a telluric star with
the same setup -- or retrieve corresponding data from the archive --
and use \texttt{ molecfit} on its reduced spectrum to determine which
molecules are relevant, the wavelength calibration or continuum
determination. Then, \texttt{molecfit} can be applied to the science
target using the derived wavelength calibration (if relevant),
changing the airmass, or fitting the continuum and/or the column
densities of the molecules involved. An example of such a strategy is
shown below in Sect.~\ref{sec:crires} and is illustrated in
particular in Fig.~\ref{fig:crires_4889p5_e}. Another possible
strategy is to ensure that the spectral setup includes spectral
regions containing representative telluric lines but free of
intrinsic lines.
\subsection{Emission line spectra}
\label{sec:nocontinuum}
The telluric transmission spectrum cannot be fitted on the spectrum of
an object showing little or no continuum, such as a faint cometary
spectrum or an emission line object. A suggested strategy in this
case is either to retrieve a spectrum of a telluric star from the
archive or observe a telluric star in the spectral set-up of interest.
Its spectrum can be modelled by \texttt{molecfit}. The model can then be
modified easily by changing the airmass such that it corresponds to
the airmass of the science target. If the water vapour at the site is
monitored, the model can also be modified accordingly. The number of
times telluric stars need to be observed is therefore reduced
significantly.
\section{Limitations}
\label{sec:limitations}
In this section, we briefly discuss the various factors that limit
the accuracy of the \texttt{molecfit} correction of
telluric spectra. Such factors can be external to \texttt{molecfit},
and are related to the accuracy with which instrumental parameters
are taken into account or are internal to the method itself.
\subsection{External}
\label{sec:externallimitations}
\paragraph{Molecular parameters' accuracy and completeness: }
A first limitation arises owing to the completeness of the incorporated
line database and on the accuracy of the contained molecular
parameters. Currently, the AER line database delivered with
the LBLRTM code is included by default with \texttt{molecfit}. It is
usually more frequently updated than the underlying main
HITRAN\footnote{\url{http://www.cfa.harvard.edu/HITRAN/}} and is
therefore recommended. However, although \texttt{ molecfit} \emph{\textit{\emph{a
priori}} }allows one to use the original HITRAN database, this
functionality is subject to the stability of the
database structure.
\paragraph{Molecules in atmospheric profile:}
\texttt{Molecfit} only offers the possibility to fit the overall volume
mixing ratio along the line of sight in the atmosphere: the overall
shape of the profile is not adjusted. If the actual profile differs
from the modelled one in some atmospheric layers, systematic errors
can occur. In particular such a situation can be caused by the specific, temporary
volume-mixing ratios of a molecule and is relatively frequent for
water vapour. Figure~\ref{fig:pwv_artprof} illustrates the possible
impact in an extreme case.
In particular, the accuracy of the transmission spectrum derived
from sky emission as described in Sect. \ref{sec:fittingskyemission}
strongly depends on the precision with which the retrieved atmospheric
profile actually represents the true one.
\paragraph{Atmospheric profile stability: }
Any correction of telluric absorption requires a relatively stable
atmosphere. The arrival of an atmospheric front is a clear case where
the atmospheric profile changes suddenly. Even if the GDAS profiles
would show such a front, its precise timing and characteristics are
usually uncertain and surely depend on the pointing direction.
Fortunately, few optical or infrared astronomical observations are
executed in such conditions.
\paragraph{Radiative transfer code accuracy:}
The reader is referred to the publication list available on the
LBLRTM web
page\footnote{In particular, \url{http://rtweb.aer.com/lblrtm_ref_pub.html}} for
references regarding the validation of the radiative transfer code.
In particular, it is worth repeating here that ``the algorithmic
accuracy of LBLRTM is approximately 0.5\% and the errors associated
with the computational procedures are of the order of five times less
than those associated with the line parameters so that the limiting
error is that attributable to the line parameters and the line shape''.
\subsection{Instrumental}
\label{sec:instrumentallimitation}
\paragraph{Grating scattering: }
Internal diffusion strongly limits the achievable accuracy of the
modelling. Villanueva et al. (2009, internal communication) found
that the CRIRES line spread function is best reproduced by a Voigt
profile that needs to be calculated over $\approx 1\,000$ pixels to account for the grating scattered light within the
instrument. As a consequence, it often appears that the core of
heavily saturated absorption lines does not reach zero intensity. The
amount of flux in the core actually depends on the spectral energy
distribution of the light reaching the grating and cannot be modelled
accurately by\texttt{ molecfit}, because\texttt{ molecfit} only applies the
convolution of the line spread function to the transmission spectrum
and not to the science target one. In addition, later investigation
(Uttenthaler 2010, private communication) indicates that the spatial
profile of the grating scattered light is different from the one of
the direct light.
The observed spectrum therefore combines different effects that
cannot be easily modelled by \texttt{ molecfit}. Therefore they limit
the accuracy of the correction such that residuals can reach an
estimated 1\% of the continuum depending on the significance of the
grating scattering.
\paragraph{Instrumental background: }
When modelling sky emission spectra, \texttt{molecfit}
uses a background defined by a grey body at the temperature
of the main mirror of the telescope. It assumes that the
thermal background from the instrument is negligible.
\paragraph{Instrumental response curve shape:}
\texttt{Molecfit} fits a low-order polynomial to
the observed continuum of the input (object spectrum) for each inclusion region. However, it
does not make a difference between the intrinsic continuum shape and the
effect of the response curve of the instrument. Good results are
obtained if the observed continuum of each inclusion region can
be modelled well by such polynomials. Alternatively, the user may
provide a normalised spectrum as input.
\paragraph{Line spread function: } The line spread function should behave
well over the wavelength range of interest, either being
constant or having a width that is linearly dependent on wavelength, as expected
in cross-dispersed spectrographs.
When the transmission spectrum is derived from the sky
emission spectrum (Sect. \ref{sec:fittingskyemission}), one should
note that the parameters of the line spread function are likely to be
different from the one of the science target given that the image of
the star does not uniformly illuminate the entrance slit.
\subsection{Internal}
\label{sec:internallimitations}
\paragraph{Good internal guess values:}
\label{sec:goodguessvalues}
By essence of the Levenberg-Marquardt fitting approach, \texttt{molecfit}
needs and is sensitive to the initial guess solution. In particular,
\texttt{molecfit} is designed to refine the wavelength calibration and
continuum determination, but it is not designed to measure these values
from an uncalibrated spectrum. Adequate initial estimates improve
the robustness of the fit and the speed of the convergence
process. Knowledge of the line spread function and spectral resolving
power are also important, although the latter can be adjusted by {\tt
molecfit}.
\paragraph{Finite pixel size:}
\texttt{Molecfit} results are best when the spectral resolution is at
least Nyquist sampled by the instrument. In any case, \texttt{molecfit}
takes the finite pixel size of the spectrum into account: \texttt{molecfit}
estimates the highest pixel resolution in a spectrum by evaluating the
wavelength grid. An oversampling factor of 5 is then applied on the
input to the radiative transfer code.
\begin{figure*}[ht]
\sidecaption
\includegraphics[width=12cm]{grb080310.pdf}
\caption{$\mathrm{H}_2\mathrm{O}$ telluric absorption lines
affecting absorption lines associated with the $z = 2.42743$ DLA
in the UVES spectrum of \object{GRB080310}. The telluric lines -
modelled by \texttt{ molecfit} - are shaded in yellow; the models for
the different components ($a,b,c,d$) of the absorption system are
represented with different colours (blue, cyan, green, orange),
while the combined model is in red. See
\citetads{2012A&A...545A..64D} for details.}
\label{fig:grb080310}
\end{figure*}
\section{Examples}
\label{sec:examples}
In the following, we illustrate the versatility of \texttt{molecfit}
through its application to data sets obtained with various ESO VLT
instruments covering a range of spectral resolving powers and
wavelength domains.
\subsection{UVES}
The prototype version of \texttt{ molecfit} was manually
adjusted to determine the
location and strength of telluric absorption lines in VLT UV--Visual
\'Echelle Spectrograph (UVES, \citeads{2000SPIE.4008..534D}) Rapid
Response Mode or Target-of-Opportunity programmes of $\gamma$-ray
bursts (UVES is a high-resolution cross-dispersed spectrograph in the
visible): indeed, some $\mathrm{H}_2\mathrm{O}$ lines mimic lines
associated with high-redshift damped Ly-$\alpha$ (DLA) lines. For
example, in the spectrum of \object{GRB050730}
\citepads{2009A&A...506..661L} obtained under Programme ID
075.A-0603 (P.I.: Fiore), $\mathrm{H}_2\mathrm{O}$ lines appear
coincident in wavelength with the \ion{Si}{ii} $^2\mathrm{P}_{3/2}$
$\lambda1816$ line, and affect the \ion{Fe}{ii} $^6\mathrm{D}_{7/2}$
$\lambda$1267 and \ion{Fe}{ii} $^4\mathrm{F}_{7/2}$ $\lambda1659$
lines, all associated with the $z = 3.69857$ DLA. Similarly, as shown
in Fig. \ref{fig:grb080310}, using\texttt{ molecfit} allowed the location and strength of the $\mathrm{H}_2\mathrm{O}$
lines affecting the \ion{Fe}{ii} $\lambda2382$, $\lambda2374$,
\ion{Fe}{ii} $^4\mathrm{D}_{7/2}$, and $\lambda2389$ lines to
be determined (amongst
others) associated with the $z = 2.42743$ DLA in the spectrum of
\object{GRB080310} \citepads{2012A&A...545A..64D} obtained
under programme ID 080.D-0526 (PI. Vreeswijk).
\subsection{FLAMES}
Figure~\ref{fig:ngc3603} presents a Fibre Large Array Multi Element
Spectrograph (FLAMES)/Giraffe spectrum of the O4~V star \object{NGC
3603-117} observed with the LR8 setting and the ARGUS integral field
unit (data courtesy of M. Gieles; Prog. ID: 079.D-0374(A)), providing a
spectral resolving power of $R \approx 10\,400$. The presented spectrum
corresponds to the brightest spaxel on the object.
The presence of water-vapour telluric absorption lines impedes any
precise determination of the width and centroid of the Paschen
hydrogen lines at 901 and 923~nm. However, there is usually no
observation of telluric stars for FLAMES. \texttt{Molecfit} can be used
to determine the PWV and to correct the science spectrum by the
transmission spectrum. As a result, a precise determination of the
width and position of the lines can be obtained by multiplying the
number of stellar lines available by three for, e.g., radial velocity
measurements in that spectral region.
\begin{figure}[ht]
\includegraphics[width=\columnwidth]{ngc3603.pdf}
\caption{FLAMES/Giraffe Spectrum of NGC~3603-117. \textit{(Black:)} Original spectrum.
\textit{(Red:)} telluric line corrected spectrum. The spectral
range covered by the shaded area was used to determine the amount
of PWV. }
\label{fig:ngc3603}
\end{figure}
\pagebreak
\subsection{SINFONI}
\label{sec:sinfoni_j}
Archival data of the field of the binary star \object{2MASS
J01033563-5515561} obtained with the Spectrograph for INtegral Field
Observations in the Near Infrared (SINFONI,
\citeads{2003SPIE.4841.1548E}) for programme ID 290.C-5022(A) (P.I.:
Delorme) were retrieved and reduced with the SINFONI pipeline and
default parameters. Only 20\,s Detector Integration Time (DIT)
$J$-band exposures obtained at a median airmass of 1.28 are shown
here. The telluric star \object{Hip 024337} was observed at an
airmass of 1.33, approximately 66 min after the start of the first
science exposure. The spectra of the secondary star (B) and
of the telluric star were extracted using a three-pixel radius centred on
the photo-centre of each source. Hot pixels or cosmics in the reduced
spectrum affected were manually edited out.
Inclusion regions were selected to cover telluric lines caused by
H$_2$O and O$_2$. Exclusion regions were defined for the pixels affected
by cosmics. Figure~\ref{fig:sinfoni_j} shows the original spectrum and
compares its correction by the transmission spectrum calculated by
\texttt{molecfit} and by the telluric star. The spectrum corrected by the
transmission spectrum calculated by \texttt{molecfit} is slightly less
noisy than the one corrected by the telluric star spectrum, and it does
not show the effect of the \ion{H}{i} Paschen $\beta$ line. On the
other hand, the correction of the lines in the $\sim 1.13~\mu$m water vapour band
is slightly less good probably because of a small change in the
line spread function.
\begin{figure}[ht]
\includegraphics[width=\columnwidth]{sinfoni_j.pdf}
\caption{SINFONI $J$ band spectrum of 2MASS J01033563-5515561
B. \textit{(Bottom:)} Pipeline reduced spectrum. The yellow shaded areas
show the spectral ranges of the inclusion regions. Exclusion
regions appear in blue. \textit{(Medium:)} Spectrum divided by
the spectrum of the telluric star. \textit{(Top:)} Spectrum
divided by the transmission spectrum calculated by \texttt{molecfit}.}
\label{fig:sinfoni_j}
\end{figure}
\subsection{X-shooter}
\label{sec:xshooter}
A reduced X-shooter \citepads{2011A&A...536A.105V} spectrum of the
luminous blue variable star \object{R71} was kindly provided to us by
Andrea Mehner. X-shooter is an instrument composed of a UV, a visible,
and a near-infrared cross-dispersed medium-resolution spectrograph.
The data were obtained on February 1, 2014 as part of the programme ID
092.D-0024(A) (P.I.: Baade). The slit width used was
$5 \arcsec \times 11 \arcsec$ and the exposure time 20 s. Figure~\ref{fig:r71}
shows the visible part of the original spectrum together with
the spectrum corrected by the transmission spectrum calculated by
\texttt{molecfit}, based on the fitting made in the regions marked in
in the shaded regions.
The wide slit used means that the line spread function is dominated by
the image quality at the time of the observations (FWHM $\approx$
2\arcsec\, in this case) and not by the slit width. No telluric star
was observed with the same instrumental setup because the observation was
obtained to correct the flux loss suffered by spectra obtained later with
a narrow slit. Even if a telluric star had been obtained, its
usefulness for telluric line correction would have strongly depended
on the stability of the seeing.
Given the overheads associated with such a short integration time, the
time spent executing the observation of the telluric star would have
been roughly 100\%\, of the time spent executing the observation of
the science target: therefore, \texttt{molecfit} also allows one to
correct telluric features in spectra obtained with a slit larger than
the image quality.
\begin{figure}[ht]
\includegraphics[width=\columnwidth]{R71_VIS_0p6.pdf}
\includegraphics[width=\columnwidth]{R71_VIS_0p8.pdf}
\caption{Visible arm X-shooter spectrum of \object{R71} LBV star.
The pipeline reduced spectrum is shown in black. The shaded areas
show the inclusion regions
used to model the line spread function and determine the column
densities of H$_2$O and O$_2$. The spectrum corrected by the derived
transmission spectrum is shown in red.}
\label{fig:r71}
\end{figure}
\subsection{VISIR}
\label{sec:visir}
\begin{figure}[ht]
\includegraphics[width=\columnwidth]{hd104327_h2s2.pdf}
\caption{VISIR high-resolution $\lambda$12.279~$\mu$m spectrum of
\object{HD 1043\,27}. \textit{Bottom:} pipeline reduced
spectrum. In red, the result of the fit obtained with
\texttt{molecfit}. \textit{Medium:} Spectrum divided by the telluric
star. \textit{Top:} Spectrum divided by the transmission
spectrum calculated by \tt{molecfit}.}
\label{fig:visir}
\end{figure}
Archival data of \object{HD 104\,237} obtained with the VLT Imager and
Spectrometer for mid InfraRed (VISIR) for programme 076.C-0129
(P.I.: van den Ancker) were retrieved and reduced with the VISIR
pipeline v.\ 3.5.1 and default parameters. Only the 12.279$\mu$m
HRG spectrum is shown here. The airmass ranged from 1.68 to
1.72.
A first attempt to use \texttt{molecfit} on the spectrum of \object{HD
104\,237} did not provide a good correction. Analysis of the
residuals showed the presence of low frequency variations of the
response curve, on a scale similar to the size of the telluric lines,
and therefore difficult to correct by the continuum fitting
implemented within \texttt{molecfit}. Indeed, the VISIR calibration
plan does not include any flat fields that could correct an effect
normally taken care of by the observation of telluric standards. To
alleviate this problem, \texttt{molecfit} was used on the associated
telluric star \object{HD 92\,305}, which actually shows photospheric
absorption lines. The same low-frequency variations can be seen in
the spectrum of the telluric star divided by the molecfit model, which
was then fitted by a seventh-degree polynomial. This continuum model was
then applied to the reduced spectrum of \object{HD 104\,237} before
another attempt with \texttt{molecfit}. This time the fit is very
satisfactory, as can be seen in Fig.\ref{fig:visir}, which can be
compared with the one for the same object shown in Fig. 1 of
\citetads{2008A&A...477..839C}.
\texttt{Molecfit} offers the advantage of (a) less telescope time
used for observation of a telluric star: in this case, the execution
of the observation for the telluric star took 36 minutes compared to
the two hours used for the science target; (b) the telluric star was
observed with an airmass difference of about 0.1, which is sufficient
to lead to significant differences in the wings of the telluric lines
where the molecular H$_2$ 0--0 S(2) ($J = $4--2) emission at 12.278
$\mu m$ emission line was expected; (c) the telluric star shows
photospheric lines that must be properly identified and corrected
for.
\subsection{KMOS}
\label{sec:kmos}
\begin{figure}[t]
\includegraphics[width=\columnwidth]{kmos_YJ_40.pdf}
\includegraphics[width=\columnwidth]{kmos_YJ_36.pdf}
\includegraphics[width=\columnwidth]{kmos_YJ_8.pdf}
\caption{KMOS spectra of 3 red supergiant stars in \object{NGC~6822,}
each feeding a different spectrograph. From top to bottom: star
\#40, star \#36, and star \#8. In each figure, the bottom graph
shows the reduced data in black, while the spectra corrected by the
synthetic spectrum determined by \texttt{molecfit} is shown in
red; the top graph shows the star spectrum corrected by the
telluric star observed just after. The molecular column densities
were determined using only the spectrum of star \# 8: the results
were used on the spectra of stars \# 36 and \# 40. The line spread
functions were, however, determined on each spectrum
independently. The yellow shaded areas indicate the inclusion
regions, while the blue shaded areas show the regions excluded from
the fit because they correspond to spectral features intrinsic to
the stars. }
\label{fig:kmos}
\end{figure}
\begin{table}[h]
\centering
\caption{Correspondence between star ID, coordinates, KMOS arms, and spectrographs.}
\begin{tabular}[]{ccccc}
\hline
\hline
Star ID & $\alpha(J2000)$ & $\delta(J2000)$ & Arm & Spectrograph\\
\hline
8 & 19 44 45.98 & $-$14 51 02.4 & 11 & 2\\
36 & 19 44 55.93 & $-$14 47 19.6 & 1 & 1\\
40 & 19 44 57.31 & $-$14 49 20.2 & 17 & 3\\
\hline
\end{tabular}
\label{tab:kmos}
\end{table}
Reduced spectra (incl. sky subtraction) of three red supergiant stars of
Barnard's galaxy \object{NGC~6822} simultaneously obtained during the
K-band Multi Object Spectrograph (KMOS,
\citetads{2013Msngr.151...21S}) science verification programme
60.A-9452 (P.I.: C. Evans) were kindly provided to us by Lee Patrick
(Patrick et al., in preparation). KMOS deploys 24 arms to
user-selected $2.8\arcsec \times 2.8\arcsec$ areas at a spatial
sampling of $0.2\arcsec \times 0.2\arcsec$ in a 7.2\arcmin\ diameter
field feeding 24 integral-field units sending light to three near-infrared
medium-resolution spectrographs. The three stars were chosen such that
their spectra were obtained on different spectrographs. The YJ
filter/grating setup was used. See Table \ref{tab:kmos} for the
correspondence between star IDs, coordinates, arms, and
spectrographs. For the purpose of this paper, the determination of the
column densities for the relevant molecules was done using only the
spectrum of star \#8, and the determined values were then used as constant
for the two other stars. On the other hand, the parameters of the line
spread function were determined independently on each spectrum.
The execution time required for the spectra shown in
Fig. \ref{fig:kmos} is slightly more than 1h. Observation of the same
telluric star (\object{HD 187\,439} of spectral type B6III) for each
arm was carried out right after and took 15 min, i.e. $\approx$ 25\%
of the time devoted to the science. The telluric stars were chosen by
the user, but the difference in airmass between the science target
(ranging from 1.45 to 1.98) and telluric star (ranging from 2.00 to
2.13) observations make the corrections by the telluric star
challenging. Instead, \texttt{molecfit} here again allows one to
avoid such problems.
Therefore in a number of cases \texttt{molecfit} produce very
satisfactory telluric corrections for KMOS. It would be interesting to
determine if the line spread function can be modelled well enough as a
function of rotation angle and temperature to avoid the need for
systematic observations of telluric stars. Such a study is outside the
scope of this paper.
\subsection{MUSE}
\label{sec:muse}
\begin{figure}[t]
\includegraphics[width=\columnwidth]{muse.pdf}
\caption{MUSE spectra of 3 stars in the globular cluster
\object{NGC~6\,397}. The
reduced data appear in black, while the spectra corrected by the
synthetic spectrum determined by \texttt{molecfit} are shown in
red. The molecular column densities and line spread function were determined using only
star \# 5326 (\textit{bottom}) located on IFU \#7: the results were used on the spectra
of star \# 10327, also located on IFU \#7 (\textit{middle}), and
\# 10893 (\textit{top}), located on IFU \# 13. The yellow shaded areas indicate the inclusion regions.
No regions were excluded from the fit.}
\label{fig:muse}
\end{figure}
\begin{table}[h]
\centering
\caption{Correspondance between star ID, coordinates, and MUSE IFUs.}
\begin{tabular}[]{cccc}
\hline
\hline
Star ID & $\alpha(J2000)$ & $\delta(J2000)$ & IFU \\
\hline
5326 & 17 40 43.51 & $-$53 40 26.0 & 7 \\
10327 & 17 40 42.46 & $-$53 40 11.4 & 7 \\
10893 & 17 40 40.66 & $-$53 40 13.4 & 13 \\
\hline
\end{tabular}
\label{tab:muse}
\end{table}
Reduced spectra of three stars of the globular cluster \object{NGC~6\,397}
obtained during the Multi-Unit Spectroscopic Explorer (MUSE)
\citepads{2010SPIE.7735E..08B} commissioning 2B (programme ID:
60.A-9100(C)) were kindly provided to us by Roland Bacon and Sebastian
Kamann. MUSE is an integral field spectrograph composed of 24
identical IFU modules sampling a $1\arcmin \times 1\arcmin$
field-of-view at ($0.2\arcsec \times 0.2\arcsec$) spatial resolution
and up to $R \sim 3\,000$ spectral resolving power. The three stars were
chosen such that their signal-to-noise ratio is greater than 100 and so that
two of them fall on the same IFU and the third one on a different IFU.
See Table \ref{tab:muse} for the correspondance between star ID,
coordinate, and IFU. The column densities for the
relevant molecules (O$_2$, H$_2$O) and line spread function were determined
using only the spectrum of star \#5326. The derived transmission
spectrum was then applied to the other stars, \# 10327 and \# 10893.
The very good wavelength calibration and the similarity of the line
spread functions between IFUs explain that the same transmission
spectrum can be used for these IFUs. \texttt{Molecfit} can therefore
potentially be used to correct for telluric absorption for all 90\,000
spectra obtained by MUSE at once, provided a suitable star falls in
the field of view. The spectral resolution of the instrument seems
low enough not to be affected by the details of the atmospheric
profiles described in Sect.~\ref{sec:atmprofile}, so that
\texttt{molecfit} can be used on a spectrum resulting from the
combination of multiple observations spread over several, but not
necessarily successive nights.
\subsection{CRIRES}
\label{sec:crires}
\begin{figure}[t]
\includegraphics[width=\columnwidth]{gammaGem_4889p5.pdf}
\includegraphics[width=\columnwidth]{HD20010_4889p5.pdf}
\includegraphics[width=\columnwidth]{Barnard_4889p5.pdf}
\caption{Extract (detector \#2) of CRIRES $4889.5$ nm setting
spectra of \object{$\gamma$ Gem} \textit{(top)}, \object{$\alpha$
For} \textit{(middle),} and \object{Barnard's star}
\textit{(bottom)}. In each case, the bottom graph shows the
original, reduced spectrum and the blue-shaded exclusion regions;
the red spectrum shows the \texttt{molecfit} derived transmission
spectrum scaled by a constant factor. The top graph shows the {\tt
molecfit} corrected spectrum.}
\label{fig:crires_4889p5}
\end{figure}
Spectra obtained with the Cryogenic high-resolution IR \'Echelle
Spectrograph (CRIRES, \citeads{2004SPIE.5492.1218K}) before mid-July
2014 -- when the instrument was removed from operations to undergo an
upgrade -- covered a narrow spectral range ($\Delta\lambda \approx
\lambda/70$) at a resolution $R \sim 50\,000 ~ \mathrm{to}
\, 100\,000$. Although the line spread function is usually well
modelled by a Gaussian function \citepads{2010A&A...524A..11S}, its
wings are more Lorentzian. Therefore a Voigt profile is usually best
for representing it (Villanueva et al. 2009, internal communication), in
particular, to attempt to model the light scattered on the grating
that causes even heavily saturated lines to appear with residual light
in their core. The kernel width, on the other hand, depends on the
quality of the adaptive optics (AO) correction -- when the AO is used
-- which in turn depends on the `AO star' magnitude and on the
turbulence profile of the atmosphere. On the other hand, for non-AO
observations, it may also depend on the actual slit width; one should
note here that the slit mechanism lacked reliability before the
installation of fixed slit widths in September 2011 \citep{CRIRES_UM}.
As mentioned in the CRIRES User Manual \citep{CRIRES_UM}, CRIRES
spectra are often affected by imprecise wavelength calibration, owing to
a lack of calibration lines. \texttt{Molecfit} can solve this problem in
a number of situations using the telluric lines themselves when they
are distributed well over the spectral range and not too mixed with
lines that are intrinsic to the science target.
Figure~\ref{fig:crires_4889p5} shows an extract (detector \#2 only) of
the 4889.5\,nm setting spectra for three stars obtained as part of the
CRIRES-POP programme \citepads{2012A&A...539A.109L}: the A0~IV star
$\gamma$ Gem, observed on October 20, 2010 (programme ID
086.D-0066, P.I.: Lebzelter); the F8~V star
\object{$\alpha$ For}, observed on October 31, 2009 (programme ID
084.D-0912, P.I.: Lebzelter); and the M4~V
\object{Barnard's star}, observed on July 7, 2010 (programme ID
086.D-0066, P.I.: Lebzelter). No telluric stars
were observed for this programme. The number of intrinsic lines
increases for late-type stars, thus requiring an increase in the
number and size of exclusion regions; consequently, an F8V star is
approximately the latest star for which \texttt{molecfit} can be
used in a non-expert mode.
\begin{figure}[h]
\includegraphics[width=\columnwidth]{HD134453_4889p5.pdf}
\caption{Extract (detector \#2) of CRIRES $4889.5$ nm setting
spectra of \object{X TrA}; the red spectrum shows the \texttt{molecfit}-derived transmission spectrum scaled by a constant
factor. The top graph shows the corrected
spectrum. \texttt{Molecfit} was used in expert mode in this case.}
\label{fig:crires_4889p5_e}
\end{figure}
For later stars, the expert mode (see Sect.~\ref{sec:expert})
is required. The variable
carbon star \object{X~TrA} was observed on September 22, 2010
as part of the same programme.
Regarding the molecular content of the atmosphere, \texttt{molecfit} was
configured to fit only the column density for H$_2$O and CO$_2$, while
the O$_3$ and OCS column densities were fixed. The polynomial coefficients
for the continuum and wavelength calibration were copied from
the results of the fit to the \object{$\gamma$~Gem} spectrum. The wavelength
solution found for \object{$\gamma$~Gem} was kept fixed. This method allowed one to
obtain very good results for the telluric line correction, as shown
in Fig.~\ref{fig:crires_4889p5_e}.
\begin{figure*}[h]
\includegraphics[]{HD61421_2332p2.pdf}
\includegraphics[]{HD61421_2368p7.pdf}
\caption{Extract (detector \#3) of CRIRES $2332.2$
\textit{(top)}
and $2368.7$ nm \textit{(bottom)} setting
spectra of \object{Procyon}. In each figure the top graphs
show generic (not model) transmission spectra for the relevant molecules:
H$_2$O in black, CH$_4$ in red, CO$_2$ in green.
The bottom graphs show the original pipeline reduced spectra
in black, as well as the molecfit-corrected
spectra, in red. The dotted lines show 2\% deviation from the
median value of the corrected spectra.}
\label{fig:crires_highsnr}
\end{figure*}
To illustrate the quality of the correction on high signal-to-noise
spectra, we used \texttt{molecfit} on the 2332.2 and 2368.7 nm setting spectra
of the F5~IV star Procyon observed on November 7, 2011 (programme ID
088.D-0109, P.I.: Lebzelter). The signal-to-noise ratios - estimated
on clean regions of the \texttt{molecfit}-corrected spectra - are
$\sim 350$ and 400, respectively. Figure~\ref{fig:crires_highsnr}
clearly indicates a very good correction, well within 2\% except for
the core of the moderately saturated lines. A slight
difference in the quality of the correction can also be seen between
H$_2$O and CH$_4$ lines, most likely caused by slight errors in the
atmospheric profiles.
Observations of telluric stars with CRIRES can easily take longer than
for the science targets. Given the high signal-to-noise of the spectra
obtained, differences in observing conditions and in airmass may
significantly affect spectra corrected by telluric standards. The use
of a tool such as \texttt{molecfit} therefore allows one to
significantly increase the efficiency of the instrument. Some
programmes, such as CRIRES-POP \citepads{2012A&A...539A.109L} have
deliberately decided to only rely on synthetic telluric correction.
\subsection{Accuracy of the telluric correction}
\label{sec:summary}
To estimate the quality of the correction, we compared the standard
deviation $\sigma$ of the \texttt{molecfit}-corrected
spectrum normalized by the continuum in two separate regions for
several of the examples above: (1) a reference region with no telluric
lines and (2) a region affected by telluric lines.
The \texttt{molecfit}-corrected range is a region affected by
unsaturated or moderately saturated telluric lines but not
covered by any inclusion region, except for VISIR and CRIRES, for which
the inclusion region corresponds to the whole spectrum. Both regions
are free -- as much as possible -- of intrinsic features (either
in emission or absorption). To avoid being affected by
differences in signal-to-noise ratios, these two ranges are close to
each other. The continuum was fitted locally by a low-degree
polynomial on each of the two ranges of the telluric-corrected spectrum.
Representative values of $\sigma$ are reported in Table
\ref{tab:accuracy}.
We could not define reference regions in the spectra shown in
Fig.~\ref{fig:crires_highsnr}. The value listed in Table
\ref{tab:accuracy} therefore corresponds to most of the spectra
(avoiding an intrinsic line in the 2332.2 setting spectrum) and can
be compared to the signal-to-noise ratios given above.
\begin{table*}[h]
\caption{Representative measurements of the accuracy of the telluric
correction in spectral ranges affected by unsaturated or
moderately saturated lines.}
\centering
\begin{tabular}[]{l|llc|llc}
\hline\hline
& \multicolumn{3}{c}{Reference} & \multicolumn{3}{c}{Telluric-corrected} \\
Instrument & $\lambda_\mathrm{min}$ &$\lambda_\mathrm{max}$& $\sigma$
&$\lambda_\mathrm{min}$ &$\lambda_\mathrm{min}$ & $\sigma$ \\
& ($\mu m$) & ($\mu m$) & ($\times
100)$&($\mu m$) & ($\mu m$) & ($\times 100$)\\
\hline
FLAMES & 0.88823 & 0.89111 & 1.2 & 0.89555 & 0.89913 & 1.0\\
SINFONI & 1.2230 & 1.2420 & 2.1 & 1.2940 & 1.3045 & 1.6\\
SINFONI & 1.2230 & 1.2420 & 2.1 & 1.1500 & 1.1600 & 6.2\\
X-shooter VIS & 0.878 & 0.885 & 1.0 & 0.894 & 0.890 & 1.5\\
X-shooter VIS & 0.806 & 0.810 & 0.8 & 0.810 & 0.818 & 1.1\\
VISIR & 12.263 & 12.273 & 2.0 & 12.283 & 12.285 & 2.5\\
KMOS \# 40 & 1.162 & 1.170 & 4.7 & 1.140 & 1.150 & 5.4\\
MUSE \# 10893 & 0.708 & 0.714 & 0.4 & 0.716 & 0.733 & 1.0\\
CRIRES $\alpha$ For & 4.8485 & 4.8500 & 2.5 & 4.8516 & 4.8524 & 2.7\\
CRIRES Procyon & & & & 2.3270 & 2.3350 & 0.6\\
CRIRES Procyon & & & & 2.3635 & 2.3740 & 1.2\\
\hline
\end{tabular}
\label{tab:accuracy}
\end{table*}
These measurements show that for the examples shown in this paper,
\texttt{molecfit} is able to correct unsaturated lines to a
standard deviation better than 2\% of the continuum. In Paper II we
systematically investigate the quality of the telluric corrections by
\texttt{molecfit} over a large sample of X-shooter near-infrared
spectra and by comparison with the IRAF \texttt{telluric} task.
\section{Conclusion}
\label{sec:conclusion}
\texttt{ Molecfit} is a versatile tool for modelling and correcting telluric
absorption lines. In its most common use, \texttt{molecfit} fits a
spectrum of the transmission of the Earth's atmosphere over narrow,
user-selected ranges (inclusion regions) of the spectrum of a science
target. This allows one to determine the column densities of the
relevant molecules causing telluric lines and the parameters of the
line spread function. Alternatively, the fit can partially or fully
make use of parameters determined on spectra of other objects obtained
at a different time. In all cases, \texttt{molecfit} can take the best information of temperature, pressure, and humidity into
account in
the atmosphere above the observatory at the time of observing
the science target. \texttt{Molecfit} then derives a transmission
spectrum for the whole range of the input spectrum and corrects for
it. This approach allows one to regularly reach an accuracy of the
order of 2\% of the continuum or better in correcting unsaturated
telluric absorption lines.
In most cases, the input spectrum must meet the following conditions
for a good correction: (1) its wavelength calibration must be
accurate, (2) the continuum in the narrow ranges used as inclusion
regions is already corrected or can be modelled well by low-degree
polynomials, (3) the inclusion regions cover unsaturated lines of
the relevant molecules whose column density varies significantly with
time (mostly water vapour, but also CO$_2$, O$_3$) and whose shapes
allow one to determine the line spread function with sufficient
precision.
However, the versatility of the tool allows several departures from
these conditions. For example,\texttt{ molecfit} can use the
telluric lines themselves to provide an improved wavelength
calibration, provided that they are well spread over the spectral
range of interest and are reasonably unaffected by intrinsic spectral
lines of the science target. Additionally, the user can indicate to
\texttt{molecfit} that absorption or emission lines in the inclusion
regions are intrinsic to the object and exclude them from the fit.
Finally, it is important to remember that any correction method cannot
recover with accuracy any science target feature coincident in
wavelength with $\tau \gtrsim $ 2 telluric absorption lines. Such
\textnormal{saturated} lines can be easily identified in high-resolution
spectra (as obtained by CRIRES) where these lines are often
spectrally resolved; however, in low-to -medium-resolution spectra
(such as X-shooter), their shape is mainly determined by the
instrument's spectral resolution. In other words, even apparently weak
telluric absorption lines can be caused by saturated lines.
\begin{acknowledgements}
A prototype version of \texttt{molecfit} was built around the Reference
Forward Model, a GENLN2-based line-by-line radiative transfer model
originally developed at the Atmospheric, Oceanic and Planetary
Physics Laboratory, Oxford University, to provide reference spectral
calculations for the MIPAS launched on the ENVISAT satellite in
2002. We warmly thank Anu Dudhia for his help in various phases of
the prototype development. We also thank Andreas Seifahrt for his
help in usingf LNFL/LBLRTM.
This study was carried out in the framework of the Austrian ESO
In-Kind project funded by BM:wf under contracts
BMWF-10.490/0009-II/10/2009 and BMWF-10.490/0008-II/3/2011. This
publication is also supported by the Austrian Science Fund (FWF):
P26130 and by the project IS538003 (Hochschulraumstrukturmittel)
provided by the Austrian Ministry for Research (bmwfw).
A.G. acknowledges support from FONDECYT grant 3130361.
We would like to thank the participants to an internal ESO
mini-workshop who provided us with sample spectra from various
instruments obtained for various scientific purposes: their help
allowed us to identify a number of problems in a previous version of
the \texttt{molecfit} package. We are also grateful to Andrea Mehner
and Patrick Lee for providing spectra ahead of publications, to
Annalisa De Cia for providing Fig.\,\ref{fig:grb080310}, and to Andr\'e
M\"uller for various comments. We thank the anonymous referee whose
constructive comments improve the quality of this article.
\end{acknowledgements}
\bibliographystyle{aa}
|
\section{JEP and Pure AEC}\label{pure}
\numberwithin{Thm}{section}
One can trivially augment any AEC $\bK$ by adding structures below
the L\"{o}wenheim-Skolem number LS($\bK$) which have no extensions; to avoid such
trivialities the following assumption applies to the rest of the
paper.
\begin{assumption}\label{BasicAssumption}
For each AEC $(\bK,\subm)$, we work at or above the L\"{o}wenheim-Skolem number LS($\bK$).
\end{assumption}
In this section we spell out the parameterized notions of joint
embedding and introduce the notion of pure and hybrid AEC. We then
show that there is no real theory possible if hybrid AEC are allowed.
\begin{Def}\label{jepdef} The AEC $(\bK,\subm)$ has the {\em joint embedding
property}
at the infinite cardinal $\kappa$ (JEP$(\kappa)$) if for any two
models $A,B$ of cardinality $\kappa$ have a common $\subm$ extension
$C$.
If this condition holds for models $A,B$ of any cardinality $\leq
\kappa$ ($<\kappa$) we write JEP$(\leq\kappa)$ (JEP$(<\kappa)$). In particular, $|A|$ and $|B|$ can be
different.
The \emph{full-joint embedding property} (full-JEP) is the JEP with
no restriction on the sizes of $A,B$.
The AEC $(\bK,\subm)$ has the {\em amalgamation property}
at the infinite cardinal $\kappa$ (AP$(\kappa)$) if for any three
models $A,B,C$ each of cardinality $\kappa$ there are $\subm$-maps
of $B$ and $C$ into an
extension $D$ which agree on $A$. We use cardinal parameters as for
joint embedding.
\end{Def}
By Assumption \ref{BasicAssumption}, $\kappa$ is greater or equal to LS($\bK$) and without loss of generality we can assume that $C$ in the
definition of JEP and $D$ in the definition of AP, both have size $\kappa$.
The following easy consequences of the definitions show there are
some subtleties in the relation between joint embedding, maximal
models and arbitrarily large model.
\begin{Lem}\label{basicjep} Let $(\bK,\subm)$ be an AEC
\begin{enumerate}
\item If there are no maximal models then $\bK$ has arbitrarily
large models.
\item If $(\bK,\subm)$ satisfies JEP$(\le\kappa)$ and has a
model in power $\kappa$ then any model extends to one of
size of $\kappa$; thus
\item If $(\bK,\subm)$ has full-JEP and has arbitrarily large
models then $\bK$ has no maximal models.
\item If $\bK$ has two non-isomorphic maximal models in power
$\kappa$, JEP$(\kappa)$ fails.
\end{enumerate}
\end{Lem}
We will construct an AEC with at least two maximal models in a
cardinal $\lambda^+$. Condition 4) says the most JEP possible is
JEP$(\leq \lambda)$. Some of our examples will satisfy this. For
others we settle for JEP$(< \lambda)$.
We show that without the hypothesis of full-JEP the implication from
arbitrarily large models to no maximal models fails on various
countable sets of cardinals. There are some trivial examples for this
(see Corollary \ref{Cor:TrivialExample}), where one just takes
disjunctions of sentences (in disjoint vocabularies). However, the
disjunction of two sentences introduces properties that are clearly
artificial. In particular, one can find sentences with maximal models
in any countable set of cardinals by putting them in disjoint
vocabularies and taking a disjunction. Such a class does not have
JEP in any cardinal. We eliminate these trivial examples using the
following definition. Moreover, our examples also satisfy JEP$(\le\lambda)$, for some infinite $\lambda$.
\begin{Def} \label{Def:PureAEC}
\begin{itemize}
\item Let $\bK$ be a collection of $\tau$-structures and $\tau_1$ be a subset of $\tau$.
Then $\bK_{\tau_1}$ is the subcollection of $\bK$ of
models where all symbols not in $\tau_1$ have the empty interpretation.
\item An AEC $(\bK,\subm)$ in a vocabulary $\tau$ is called {\em
hybrid} if $\tau =\tau_1 \cup \tau_2$, $\bK = \bK_{\tau_1}\cup
\bK_{\tau_2}$ and at least one of $\tau_1$, $\tau_2$ is not equal to $\tau$.
If $\bK$ is not hybrid then it is \emph{pure}.
\end{itemize}
\end{Def}
The most trivial example of a hybrid AEC is defined by the
disjunction of sentences in disjoint vocabularies. The definition
allows a more subtle version where the vocabularies can overlap but
one of the classes forces some of the relations to be empty. Lemma \ref{Lem:FullJEP} provides an example of how hybrid AEC that are not just
disjunctions of sentences in disjoint vocabularies give trivial counterexamples.
Note that trivially JEP$(<\kappa)$ for all $\kappa$, or AP$(<\kappa)$
for all $\kappa$, imply full-JEP and full-AP respectively. On the contrary, Lemma \ref{Lem:FullJEP} proves that the assumption JEP$(<\kappa)$, for
all $\kappa$, can not be replaced by the assumption JEP$(\kappa)$, for all $\kappa$. In addition (see Corollary \ref{Cor:TrivialExample}) the
example in Lemma \ref{Lem:FullJEP} has AP$(\kappa)$, for all uncountable $\kappa$, but fails AP$(\aleph_0)$ and thus it fails AP$(<\kappa)$, for all
uncountable $\kappa$.
So there is a genuine distinction between these types of conditions. This is an important distinction since the definition of a good-$\kappa$
frame requires the weaker AP$(\kappa)$ and not AP$(<\kappa)$.
\begin{Lem}\label{Lem:FullJEP} The full-Joint Embedding Property is \emph{not} equivalent to the conjunction of JEP$(\kappa)$,
for all infinite $\kappa$.
\begin{proof}
Let $\tau$ be the vocabulary $\{V,U,E,<\}$, where the $V,U$ are unary
predicate symbols, and $E,<$ are binary predicate symbols. Consider
$\phi_1$ to be the conjunction of $\tau$-sentences asserting:
\begin{enumerate}
\item $V,U$ partition the universe;
\item $(U,<)$ is well-ordered in order type $\omega$; and
\item $E$ defines a bijection from $V$ onto $U$.
\end{enumerate}
Let $\phi_2$ be the conjunction of
\begin{enumerate}
\item $U$ is empty, $V$ is infinite and equals the universe; and
\item $<,E$ are empty
\end{enumerate}
Let $\bK$ be the collection of models of $\phi_1 \vee \phi_2$, and
let $\subm$ be the substructure relation.
If $\M_1,\M_2$ are two models of $\phi_2$ of the same cardinality
$\kappa$, then they are isomorphic so can be jointly embedded. If
$\M_1,\M_2$ are two (necessarily countable) models of $\phi_1$, then
$\M_1$ and $\M_2$ are isomorphic. If $\M_1$ is a model of $\phi_1$
and $\M_2$ is a countable model of $\phi_2$, then $\M_2$ can be
embedded in $\M_1$. So, for all infinite $\kappa$, JEP$(\kappa)$
holds.
However, if $\M_1$ is a countable model of $\phi_1$ and $\M_2$ is an
uncountable model of $\phi_2$,
$\M_1,\M_2$ have no common extension in
$\bK$ (as $\M_1$ is maximal in $\bK$). So, full-JEP fails.
\end{proof}
\end{Lem}
\begin{Cor} \label{Cor:TrivialExample} The AEC $(\bK,\subm)$ defined in Lemma \ref{Lem:FullJEP}
is a hybrid AEC that satisfies JEP$(\kappa)$ for all infinite $\kappa$, has maximal models
in $\aleph_0$, and has arbitrarily large models, but fails JEP$(\le\aleph_1)$. Moreover, it satisfies AP$(\kappa)$, for all uncountable
$\kappa$, but it fails AP$(\aleph_0)$ and AP$(\le\aleph_1)$.
\end{Cor}
\begin{question} Is there a `pure' example (according to Definition \ref{Def:PureAEC}) to illustrate the distinction between full-JEP and
JEP$(\kappa)$ for all $\kappa$? \end{question}
We contrast this result with a less complicated version of a result
of Shelah (Theorem 2.8 of \cite{Sh88}) which was originally stated
without proof.\footnote{The proof is sketched on page 134 of
\cite{Shaecbook}. A weaker form of this result was reproved in
\cite{BLS} (Theorem 3.4), inadvertently without citation. The result
also occurs in Rami Grossberg's master's thesis.}.
\begin{Fact}\label{AP} If an AEC has AP$(\kappa)$ for every $\kappa$, then it
has the (full-) Amalgamation Property.
\end{Fact}
An easy variation of the proof shows:
\begin{Cor} If $\lambda<\kappa$ and an AEC satisfies JEP$(\lambda)$ and AP$(\le\kappa)$, then it also satisfies JEP$(\kappa)$ and even
JEP$(\le\kappa)$.
\end{Cor}
Thus, for the distinction made in Lemma \ref{Lem:FullJEP} between JEP$(\le\aleph_1)$ and the conjunction of JEP$(\aleph_0)$ and JEP$(\aleph_1)$, it
is imperative for AP$(\le\aleph_1)$ to fail.
We can modify the example of Lemma \ref{Lem:FullJEP} to allow
$(U,<)$ to be an infinite well-order of order type $\le\kappa$ (with
strong substructure as end-extension), for some cardinal $\kappa$.
The resulting AEC will satisfy JEP$(\lambda)$, for all infinite
$\lambda$, and even JEP$(\le\kappa)$, but fail JEP$(\le\kappa^+)$.
\begin{Cor} For all infinite $\kappa$,
\begin{enumerate}
\item JEP$(\le\kappa^+)$ is \emph{not} equivalent to the conjunction of JEP$(\lambda)$, $\lambda\le\kappa^+$.
\item JEP$(\le\kappa^+)$ is \emph{not} equivalent to the conjunction of JEP$(\le\kappa)$ and JEP$(\kappa^+)$.
\end{enumerate}
\end{Cor}
We will see with more difficulty below that there are pure AEC which exhibit the behavior of Corollary \ref{Cor:TrivialExample}, i.e.
they have both maximal models and arbitrarily large models.
\section{Basic combinatorics}\label{section:intro}
\numberwithin{Thm}{section}
In this section we set up a first-order template of bipartite graphs on
sets $A,B$ with colors from $C$. We introduce the requirement that
there is no monochromatic $K_{2,2}$ subgraph (a complete bipartite graph on
points $a_1, a_2 \in A, b_1, b_2 \in B$ with all edges the same
color). Then we show restrictions on the cardinality of $A$ and $B$
that are imposed by restrictions on the number of colors. In later
sections, we will impose the restrictions on $|C|$ by characterizing
them by sentences of $\lomegaone$ in the following sense.
\begin{Def} \label{Def:CharacterizableCard} An $\lomegaone$-sentence $\phi$ \emph{characterizes} an infinite cardinal $\kappa$,
if $\phi$ has models in all cardinalities $\le\kappa$, but no model of size $\kappa^+$. In this
case we say that the cardinal $\kappa$ is \emph{characterizable}.
\end{Def}
Since the Hanf number for $\lomegaone$-sentences is $\beth_{\omega_1}$, it follows that all characterizable
cardinals are strictly less than
$\beth_{\omega_1}$.
\begin{Not}\label{sigma1def} Let $\tau_0=\{A,B,C,E\}$ where $A,B,C$ are unary
predicates and $E$ is a ternary relation. Let $\sigma_0$ be the
conjunction of the following statements:
\begin{itemize}
\item $A,B,C$ are non-empty and partition the universe.
\item $E\subset A\times B\times C$ defines a total function
from $A\times B$ into $C$.
\end{itemize}
As a notation, let $F(a,b)$ the unique value $c$ such that
$E(a,b,c)$ holds.
$A$ and $B$ should be regarded as the two sides of a bipartite
graph and $C$ as the set of edge-labels. $E$ assigns a unique
label to any pair from $A\times B$.
Let $\sigma_1$ be the conjunction of $\sigma_0$ and
\begin{itemize}
\item[($*$)] for all distinct $a_1,a_2$ in $A$ and $b_1,b_2$ in
$B$, the four values $F(a_i,b_j)$ ($i,j\in\{1,2\}$) are
not all identical.
\end{itemize}
\end{Not}
We will also refer to $(*)$ as ``there are no monochromatic $K_{2,2}$ subgraphs''.
If $a_1,a_2$ are two distinct elements in $A$ and for some $b\in B$, $F(a_1,b)=F(a_2,b)$, we will say that there exists a \emph{monochromatic path} of
length $2$, or \emph{monochromatic $2$-path}, on $a_1,a_2$.
\begin{Lem}\label{Lem:BasicCombinatorics}
In any model of $\sigma_1$, if $|A|>|C|^+$ then $|B|\leq|C|$. By
symmetry, the same is true if we switch the roles of $A$ and $B$.
\end{Lem}
\begin{proof}
Toward a contradiction, assume that $|B|>|C|$. For any subset $D$
of $B$ and any element $a\in A$, define $S(a,D)=\{F(a,d)|d\in D\}$
(which is a subset of $C$).
Now given any such $D$ of size $|C|$ and any $b\in B\setminus D$,
we observe that for all but $|C|$ many elements $a\in A$,
$S(a,D)\subsetneq S(a,D\cup\{b\})$. Indeed, if $S(a,D)=S(a,D\cup\{b\})$, then we have some $c\in C$ and $d_a\in D$ with $F(a,d_a)=F(a,b)=c$.
If $a,a'$ are distinct elements of $A$ and $F(a,d_a)=F(a,b)=c=F(a',d_{a'})=F(a',b)$, then $d_a$ and $d_{a'}$ have to be distinct.
If not, $a,a',d_a,b$ witness a violation to ($*$). So, for every color $c$, the set $\{a\in A|F(a,b)=c\}$ has size at most $|D|$. Since there
exist $|C|$ many colors, there are at most $|D|\cdot|C|=|C|$ many elements such that $S(a,D)=S(a,D\cup \{b\})$.
Now let $(b_i|i<|C|^+)$ be a sequence of distinct elements in
$B\setminus D$ and set $D_i=D\cup\{b_j|j<i\}$.
For each $i<|C|^+$, let $A_i\subset A$ be the set of elements $a$
such that $S(a,D_i)\subsetneq S(a,D_{i+1})$. Since
$|A|>|C|^+$ and all $A\setminus A_i$ have size at most $|C|$,
$A^*=\bigcap\limits_{i<|C|^+}A_i$ is non-empty (in fact its complement
has size at most $|C|^+$). But for any element $a\in A^*$,
$S(a, D_i)$ grows at each step $i<|C|^+$ which is impossible since
$S(a, D_i)\subset C$ for all $i$.
\end{proof}
\section{Maximal models in many cardinalities}\label{incom}
\numberwithin{Thm}{subsection}
In this section we prove that one can
have interesting spectra of maximal models for AEC that are \emph{pure} (see Definition \ref{Def:PureAEC}). Specifically, we
construct sentences in $L_{\omega_1,\omega}$ that are not just
disjunctions of complete sentences. In Section~\ref{com}, we show limitations on getting such results for complete sentences compatible with
$\sigma_1$.
In \cite{Sh300} Shelah defines a {\em universal class} as one that
is closed under substructure, union of chains, and isomorphism. He
remarks that by a result of Tarski, if the vocabulary is finite, then
such a class is axiomatized by a set of universal first order
sentences. This generalizes to: If the vocabulary has cardinality
$\kappa$, the class is axiomatized in $L_{\kappa^+,\omega}$. For
simplicity here we use only countable vocabularies and
$L_{\omega_1,\omega}$-sentences.
\subsection{Maximal Models}\label{maxmod}
\medskip
Throughout this section $\M$ is a model of $\sigma_1$ of cardinality
$\kappa$. Since we discuss in this subsection only the construction
of extensions of a single model we are free to assume that $C
\subseteq \kappa$. We write $C^{\M} = C \subset \kappa$ to assert
that the interpretation of the predicate $C$ is a subset of $\kappa$
and use $|C|$ when we mean cardinality.
In this section we build models $\M$ of $\sigma_1$ that are
$C$-maximal in the following sense.
\begin{Def} Let $|C| = \kappa$; $\M$ is a $C$-maximal model of $\sigma_1$ if $C^{\M} =
C$ and there is no proper extension of $\M$ to a model $\M'$ of
$\sigma_1$ with $C^{\M'} = C$.
\end{Def}
In applications we require that we will expand
$\tau_0$ to a vocabulary $\tau' = \tau_0 \cup \tau_1$ and study
$\tau'$-models $M$ such that $M\restriction \tau_0 \models \sigma_1$
and $C^M \restriction \tau_1$ belongs to an AEC
$(\bK_0,\subKzero)$ for the vocabulary $\tau_1$, and $\bK_0$ has models in
cardinality $\kappa$ but no larger, and thus it has a maximal model in
$\kappa$.
\begin{Not}
If $\M\models\sigma_1$, $|A^\M|=\kappa$ and $|B^\M|=\lambda$
then we say that $\M$ is a $(\kappa,\lambda)$-model.
\end{Not}
In the following construction, we reverse the procedure of
Lemma~\ref{xxmodel} and build a model from a function on cardinals.
\begin{Lem}\label{xxmodel} For any $\kappa$, there is a $(\kappa^+,\kappa^+)$ model $\M\models\sigma_1$ such that $C^\M=\kappa$.
\end{Lem}
\begin{proof} Let $A^\M$ and $B^\M$ be two copies of $\kappa^+$.
Fix a function $F$ from $\kappa^+\times \kappa^+$ to $\kappa$ such that a)\footnote{This requirement is not
needed now but is used in the proof of Lemma~\ref{lem:maximal1}.} for all
$\alpha$,
$F(\alpha,\alpha) = 0$ and b) for all $\alpha\in A$, $F(\alpha,\cdot)$ is a
one-to-one function when restricted to the set $\{\beta\in
B|\beta\le\alpha\}$. Symmetrically, demand that for all $\beta\in B$,
$F(\cdot,\beta)$ is a one-to-one function when restricted to the set
$\{\alpha\in A|\alpha\le\beta\}$. Both conditions are possible
because all initial segments have size $\kappa=|C^\M|$. Then define a
graph as in Notation~\ref{sigma1def} using this function.
Towards contradiction, assume that there are distinct $\alpha_1,\alpha_2$ in $A$ and $\beta_1,\beta_2$ in $B$ with all
four values $F(\alpha_i,\beta_j)$ ($i,j\in\{1,2\}$) identical.
Without loss of generality assume that
$\max\{\alpha_1,\alpha_2,\beta_1,\beta_2\}=\alpha_1$. By the choice
of $F$, $F(\alpha_1,\beta_1)$ must be
different than $F(\alpha_1,\beta_2)$. Contradiction.
\end{proof}
\begin{Cor}\label{getmax}
For any infinite cardinal $\kappa$, the class of all models
$\N$ of $\sigma_1$ with $C^{\N} = C$ where $C=\kappa$ is
fixed, contains a $(\kappa^+,\kappa^+)$-model $\M$ that is $C$-maximal.
\end{Cor}
\begin{proof} By fixing $C=\kappa$ we have an $(\kappa^+,\kappa^+)$ model $\M$ by Lemma~\ref{xxmodel}. But there is no
extension of $\M$ with either A or B of cardinality $>\kappa^+$ by
Lemma~\ref{Lem:BasicCombinatorics}. The collection of extensions
$\N$ of $\M$ that satisfy $\sigma_1$ with $C^{\N} = C$ is closed
under union since $\sigma$ is $\forall_1$. So some extension of $\M$
with cardinality $\kappa^+$ must have no extension.
\end{proof}
We can in fact give two explicit constructions that yield
nonisomorphic maximal models. The first proof uses Fodor's theorem,
which we state for the sake of completeness.
\begin{Fact}[Fodor] If $f$ is a regressive function on a stationary set $S\subset \kappa$, then there is a stationary set $T\subset S$ and some
$\gamma<\kappa$ such that
$f(\alpha) = \gamma$, for all $\alpha\in T$.
\end{Fact}
\begin{Lem}\label{lem:maximal1} If we modify the construction of Lemma \ref{xxmodel} to require that
\begin{center}
\begin{minipage}{.85\textwidth} $(\dag)$ for all $\alpha\in A$, $\alpha\ge \kappa$,
the function $F(\alpha,\cdot)$ restricted to $\{\beta\in B|\beta<\alpha\}$ is onto $C-\{0\}$,
\end{minipage}
\end{center}
then we obtain a $C$-maximal model.
\begin{proof}
Recall for each $\alpha$, $F(\alpha,\alpha) = 0$. First note that if
we extend $B$ by a new point $b$, then there must exist some $i\in C$
and a stationary subset $S_i$ of $A$ such that all the edges between
$s\in S_i$ and $b$ are colored $i$. Without loss of generality assume
that $S_i\subset \kappa^+\setminus\kappa$.
Now, define a function $g$ from $\kappa^+\setminus\kappa$ to $\kappa^+$ by
\[g(\alpha)=\text{ least $\beta<\alpha$ such that $F(\alpha,\beta)=i$.}\]
By $(\dag)$, $g$ is well-defined on all $\kappa^+\setminus\kappa$, and by the definition, $g$ is regressive, i.e. $g(\alpha)<\alpha$.
By Fodor's
Theorem
we get a stationary $T_i\subset S_i$ and a $\gamma_i$ such
that for each $t \in T_i$, $F(t,\gamma_i) = i$. But this contradicts $(*)$ of Notation~\ref{sigma1def}.
\end{proof}
\end{Lem}
\begin{Not}Consider the following condition $(\ddag)_A$.
\begin{center}
\begin{minipage}{.85\textwidth} $(\ddag)_A$ For any pair $(a,a')\in A^2$ and for any color $c$, there exists some $b\in B$ such that $F(a,b)=F(a',b)=c$.
\end{minipage}
\end{center}
Similarly, define $(\ddag)_B$ by exchanging the role of $a$'s and
$b$'s in $(\ddag)_A$, and let:
\begin{center}
\begin{minipage}{.85\textwidth} $(\ddag)$ is the conjunction of $(\ddag)_A$ and
$(\ddag)_B$.
\end{minipage}
\end{center}
\end{Not}
\begin{Lem}\label{lem:maximal2}
If $\M$ is a ($\kappa^+,\kappa^+$)-model of $\sigma_1\wedge(\ddag)_A$,
then $B^\M$ can not be extended, and symmetrically, $A^\M$ can not be extended from a ($\kappa^+,\kappa^+$)-models
of $\sigma_1\wedge(\ddag)_B$.
Thus, if $\M$ is a model of $\sigma_1\wedge (\ddag)$ with $C^{\M} =
\kappa$ and $C^\M=\kappa$, then $\M$ is
$C$-maximal.
\begin{proof}
Assume a model satisfies $(\ddag)_A$ with $|C^\M| =\kappa$ and $C \subseteq \kappa$. Towards contradiction,
assume we can extend $B$ by one element,
say $b$. Since there are $\kappa$ many colors
and $\kappa^+$ many
elements $a\in A$ to connect to $b$, there will be two elements $a_1,a_2\in A$ so that both
edges $(a_1,b),(a_2,b)$ get the same color $c$. Then
$(\ddag)_A$ gives a contradiction to $(*)$.
\end{proof}
\end{Lem}
\begin{Lem}\label{Lem:ddagModel} There exists a $(\kappa^+,\kappa^+)$-model of $\sigma_1$ that satisfies $(\ddag)$ and $C=\kappa$.
\begin{proof} Proceed as in the proof of Lemma \ref{xxmodel}. At every stage $\alpha$ choose either a pair $a_1,a_2<\alpha$ in $A$ or
a pair $b_1,b_2<\alpha$ in $B$, and some color $c\in C$. Organize the
construction so that every combination of a pair and a color appears
at exactly one stage. This is possible, since there are $\kappa^+$
stages and $\kappa^+$ such combinations.
Assume $(a_1,a_2)$ and $c$ are chosen at stage $\alpha$. If there is a $2$-path on $(a_1,a_2)$ colored by $c$, then do nothing more than what the
proof of Lemma \ref{xxmodel}
requires. If there is no such pair, require that the new edges $(a_1,\alpha)$ and $(a_2,\alpha)$ are both colored by $c$.
This is a small violation
of the requirement that $F(\cdot,\alpha)$ is $1$-$1$; demand that
this is the only violation. Make the analogous choice when
$(b_1,b_2)$ and $c$ are chosen.
We claim that the resulting construction satisfies $(*)$ and
obviously satisfies $(\ddag)$. Towards contradiction, assume that
there are distinct $\alpha_1,\alpha_2$ in $A$ and $\beta_1,\beta_2$
in $B$ with all four values $F(\alpha_i,\beta_j)$ ($i,j\in\{1,2\}$)
equal to the same value $c$. Without loss of generality assume that
$\max\{\alpha_1,\alpha_2,\beta_1,\beta_2\}=\alpha_1$. Observe that
$F(\alpha_1,\beta_1)=F(\alpha_1,\beta_2)=c$ is possible only if the
pair $\beta_1,\beta_2$ and the color $c$ were chosen at stage
$\alpha_1$. Split into two cases:
Case 1. $\alpha_2>\beta_1,\beta_2$. Then the same observation (for
$\alpha_2$) proves that the same pair $\beta_1,\beta_2$ and the same
color $c$ that were chosen at stage $\alpha_1$ were also chosen at
stage $\alpha_2$. But it is impossible for the same combination of
pair and color to appear more than once. Contradiction.
Case 2. $\alpha_2\le\max\{\beta_1,\beta_2\}$. Then at stage
$\alpha_1$, there already exists a $2$-path on $(\beta_1,\beta_2)$
colored by $c$. The construction requires in this case that
$F(\alpha_1,\beta_1)$ be different than $F(\alpha_1,\beta_2)$ which
again yields a contradiction. So, $(*)$ holds.
\end{proof}
\end{Lem}
The requirements of Lemma \ref{lem:maximal1} and
Lemma~\ref{lem:maximal2} are contradictory, so there are two
nonisomorphic $C$-maximal models of $\sigma_1$.
\begin{Cor}\label{cor:TwoNoniso} For all infinite cardinals $\kappa$,
there is a model $\M$ with $C^{\M} = \kappa$ that has two
non-isomorphic extensions that are $C$-maximal.
\end{Cor}
We can vary the constructions and get still other maximal models;
these construction will used in the next section.
\begin{Cor}\label{cor:maximal1} If $A^{\M} =\kappa^+$, $A_0$ is a club in $\kappa^+$ with
$A_0\cap\kappa=\emptyset$, and $C^\M \subset \kappa$ then condition
$(\dag)$ in Lemma \ref{lem:maximal1} can be relaxed to the following
condition.
\begin{center}
\begin{minipage}{.85\textwidth} $(\dag)_{A_0}$ For all $\alpha\in A_0$, the function $F(\alpha,\cdot)$ restricted to the set $\{\beta\in
B|\beta<\alpha\}$ is onto $C-\{0\}$.
\end{minipage}
\end{center}
and we still get that $\M$ is $C$-maximal.
\end{Cor}
\begin{Cor}\label{cor:maximal2} If $A^{\M} =\kappa^+$, $A_0$ is a subset of $A$ of size $\kappa^+$, and $C^\M \subset \kappa$, then
$(\ddag)_A$ can be relaxed to
\begin{center}
\begin{minipage}{.85\textwidth} $(\ddag)_{A_0}$ For any pair $(a,a')$, $a,a'\in A_0$, and for any color $c$,
there exists some $b\in B$ such that
$F(a,b)=F(a',b)=c$.
\end{minipage}
\end{center}
and we still get that $\M$ is $C$-maximal.
\end{Cor}
Notice that while condition $(\ddag)$ can be expressed by a
first-order sentence in the same vocabulary as $\sigma_1$, this is
not the case for $(\dag)$ and $(\dag)_{A_0}$. The latter conditions
make use of the ordering $<$ that we used during the proof which is
not part of the vocabulary.
Corollary \ref{cor:maximal2} will be used to construct infinitely
many nonisomorphic maximal models of $\sigma_1$ in Section
\ref{sec:maximalmodel}.
The existence of maximal models is complemented by the following lemma.
\begin{Lem}\label{getmod}
For any $\kappa$, there is a model $M\models\sigma_1$ with
$|A|$ arbitrary large, $|B|\le\kappa$ and $|C|=\kappa$.
\end{Lem}
\begin{proof}
Let $A$ be an arbitrary set, $B=\{b_\alpha|\alpha<\gamma\le\kappa\}$, and $C=\{c_\alpha|\alpha<\kappa\}$
such that $A,B,C$ are pairwise disjoint. For any $a\in A$ and $b_\alpha\in B$, set $F(a,b_\alpha)=c_\alpha$.
We cannot have a contradiction to ($*$) since each element in $B$
is connected only by edges of a fixed color and distinct elements
in $B$ get distinct colors.
\end{proof}
\begin{Cor}\label{cor:LargerThanKappa}
For any infinite cardinal $\kappa$, the class of all models
of $\sigma_1$ with $|C|=\kappa$ has arbitrary large models.
Moreover, in any model larger than $\kappa^+$, exactly one of
$A$ or $B$ has to be no larger than $\kappa$.
\end{Cor}
\subsection{Failure for Complete Sentences}\label{com}
\numberwithin{Thm}{subsection}
We show that our main combinatorial idea does not support the maximal
model spectra given above, if the $\lomegaone$-sentence is required
to be complete. For this we need to formalize the consequences of
our two types of constructions of maximal models. The next lemma
proves that the models of $\sigma_1$ given in Lemma \ref{getmod} are
typical of $(\lambda,\kappa)$-models, where $\lambda\ge\kappa^+$. We
need one definition first.
\begin{Def} Let $\M= (A,B,C,E)$ be colored by $F$. For $a\in A$, let $C_a=range(F(a,\cdot))$, and for $c\in C_a$ let
$$B_{a,c}=\{b\in B|F(a,b)=c\}.$$
\end{Def}
\begin{Lem} Let $\M$ be a $(\lambda,\kappa)$-model of $\sigma_1$, $\lambda\ge\kappa^+$, such that $|C^\M|=\kappa$. For all but $\kappa$ many $a\in A$
and for all $c\in C_a$, $|B_{a,c}|=1$.
\begin{proof} Assume otherwise, i.e. there are at least $\kappa^+$ many $a\in A$ such that there exists some $c_a\in C_a$ so that $|B_{a,c_a}|\ge 2
$. Call $A_0$ the set of these $a$'s. Since $A_0$ has size $\kappa^+$
and $C$ has size $\kappa$, we can restrict $A_0$ to some subset $A_1$
of size $\kappa^+$ such that $c_a=c$, for all $a\in A_1$. Then, for
each $a\in A_1$ choose a $2$ element subset $B'_{a,c}$ of $B_{a,c}$.
Since there are only $\kappa$ many $2$-element subsets of $\kappa$,
there exist $a_1,a_2\in A_1$, $B'_{a_1,c}=B'_{a_2,c}$. But $a_1,a_2$
witness that $(*)$ is violated. Contradiction.
\end{proof}
\end{Lem}
Now we formalize this distinction.
\begin{Lem}\label{cor:OneColor} Let $\tau_1$ be the (first-order) statement: ``There exists some $a\in A$ so that for all $c\in C_a$,
$|B_{a,c}|=1$''. If $|C|=\kappa$ and $\lambda\ge\kappa^+$ , then any
$(\lambda,\kappa)$-model of $\sigma_1$ satisfies $\tau_1$, while
$\tau_1$ is obviously false in $(\kappa^+,\kappa^+)$-models.
\end{Lem}
\begin{Cor} There is no model $\M$ of size $\kappa^{++}$ such that $|C^\M|=\kappa$ and
$\M$ satisfies the conjunction $\sigma_1\wedge \neg\tau_1$.
\end{Cor}
Now we show this combinatorics will not give a complete sentence with
two maximal models.
\begin{Thm}\label{nocomplete} For each $\kappa$ and each $\tau' \subseteq \tau_0$, there is no complete $\tau'$-sentence $\phi_\kappa$ such
that (a) $\phi_\kappa$ allows at most $\kappa$ colors, (b) $\phi_\kappa$ is consistent
with $\sigma_1$, (c) $\phi_\kappa$ has maximal models in some cardinal $\lambda\ge\kappa^+$ and
(d) $\phi_\kappa$ has arbitrarily large models.
\begin{proof} By Lemma \ref{Lem:BasicCombinatorics}, if $\phi_\kappa$ has a model of
cardinality $\lambda\ge\kappa^{++}$, this is a $(\lambda,\kappa)$-model. Then by Corollary \ref{cor:OneColor},
$\phi_\kappa$ is consistent with $\tau_1$. In particular, $\phi_\kappa$ does not have any $(\kappa^+,\kappa^+)$- models.
So, all models of $\phi_\kappa$ of size $\lambda\ge\kappa^+$ are $(\lambda,\kappa)$-models and by
Corollary \ref{cor:NotMaximal}, any such model can not be maximal.
\end{proof}
\end{Thm}
We see that all sufficiently large models are extendible.
\begin{Cor}\label{cor:NotMaximal} If $\M$ is a $(\lambda,\kappa)$-model of $\sigma_1$, $\lambda\ge\kappa^+$, and
$|C^\M|=\kappa$, then the $A$-side of $\M$ is extendible, while keeping the $B$-side of $\M$ and $C$ the same.
In particular, $\M$ is not
maximal.
\begin{proof} By Corollary \ref{cor:OneColor}, $\tau_1$ holds and let $a$ be an element that witnesses $\tau_1$. Extend $A$ by adding a new element
$a'$ and letting
$F(a',b)=F(a,b)$, for all $b\in B$. It is immediate that $(*)$ holds in the new model.
\end{proof}
\end{Cor}
\begin{obs} Before we move to the next section note that the requirement that
the requirement of $C$-maximality that appears in the results of this
section can be replaced by the requirement that $C$ is the universe
of a maximal model of an $\lomegaone(\tau')$-sentence $\phi$ for some
$\tau'$ disjoint from $\tau_0$ and $\phi$ characterizes $\kappa$ in
the sense of Definition \ref{Def:CharacterizableCard}. More
generally, we can require that $C$ belongs to an AEC with models in
cardinality $\kappa$, but no larger.
\end{obs}
\subsection{The Maximal Model Functor and JEP}\label{maxfn}
In this section we define a functor which takes us from an AEC
which has models of size $\kappa$ but no models in $\kappa^+$, to an AEC with a maximal model in $\kappa^+$
but arbitrarily large models.
{\em For the rest of the paper ${\hat \bK_0}$ is taken to depend on
$\bK_0$ as in the next definition.} We build the construction using
$\sigma_1$ from Notation~\ref{sigma1def}.
\begin{Def}\label{funcdef} Let $(\bkzero,\subKzero)$ be an AEC. The vocabulary
of ${\hat \bkzero}$ is ${\hat \tau}_0 =\tau_0 \cup \tau_{{\mbox{\scriptsize ${\boldmath {K_0}}$}}}$. Let ${\hat \bkzero}$
be the collection of models of $\sigma_1$ with the color sort $C$ the
domain of a model in $\bkzero$. Define for $M,N\in {\hat \bkzero}$,
$M\subhat N$, if $M\subset N$ and $C^M\subKzero C^N$.
\end{Def}
\begin{Lem}\label{Lem:LS(K)} $({\hat \bK_0},\subhat)$ is an AEC with the same L\"{o}wenheim-Skolem number as $\bkzero$.
\begin{proof} Since $\sigma_1$ is a $\forall^0_1$-first-order sentence, ${\hat \bK_0}$ is closed under
direct limits. The coherence axiom is straightforward. So the only
issue is to check the L\"{o}wenheim-Skolem number. Let $M$ be a model
in ${\hat \bK_0}$ and $X$ be a subset of $M$. Find some $C_1\in
\bkzero$ such that $X\cap C^M\subset C_1$, and $|C_1|=|X\cap
C^M|+LS(\bkzero)$. Let $M_0=X\cup C_1$. In particular, $C^{M_0}=C_1$
and $M_0$ belongs to ${\hat \bK_0}$. Indeed, $M_0$ satisfies
$\sigma_1$, since any violations of $(*)$ in $M_0$ would be violations
of $(*)$ in $M$ too. Contradiction. Considering that
$|M_0|\le|X|+|C_1|\le|X|+|X|+LS(\bkzero)=|X|+LS(\bkzero)$, it follows
that $LS({\hat \bK_0})=LS(\bkzero)$.
\end{proof}
\end{Lem}
\begin{Thm}\label{Thm:UncountableJEPAP} Let $\kappa$ be an uncountable cardinal,
$\bkzero$ and ${\hat \bK_0}$ be as in Definition~\ref{funcdef}, and suppose $\bkzero$
has models in cardinality $\kappa$, but no larger.
Then ${\hat \bK_0}$ is an AEC that satisfies the following
\begin{enumerate}
\item If $\lambda\leq \kappa$ then $\bkzero$ satisfies
JEP$(\leq\lambda)$ if and only if ${\hat \bK_0}$ satisfies
JEP$(\leq\lambda)$. The equivalence extends to
JEP$(<\lambda)$ and JEP$(\lambda)$.
\item AP fails in all infinite cardinals;
\item ${\hat \bK_0}$ has at least 2 maximal models in $\kappa^+$
and none in any $\lambda \neq\kappa^+$; moreover, ${\hat
\bK_0}$ fails JEP$(\le\lambda)$, even JEP$(\lambda)$, for
$\lambda \ge \kappa^+$.
\item ${\hat \bK_0}$ has arbitrarily large models; and
\item $LS({\hat \bK_0})=LS(\bkzero)$.
\end{enumerate}
Moreover, ${\hat \bK_0}$ is a pure AEC, in the sense of Definition
\ref{Def:PureAEC} if and only if $\bkzero$ is pure.
\begin{proof} First observe that since $\bkzero$ characterizes $\kappa$ it must contain some maximal models in $\kappa$.
(1) Clearly if ${\hat \bK_0}$ satisfies
JEP$(\leq\lambda)$ then $\bkzero$ satisfies
JEP$(\leq\lambda)$. For the converse, fix $\lambda \leq \kappa$ and suppose $\bkzero$
satisfies JEP$(\leq\lambda)$; we show ${\hat \bK_0}$ satisfies
JEP$(\leq\lambda)$. The other two cases (JEP$(<\lambda)$, JEP$(\lambda)$) are similar. Let $\M_1=(A_1,B_1,C_1)$,
$\M_2=(A_2,B_2,C_2)$ be two models in ${\hat \bK_0}$ such that both
$M_1,M_2$ have size $\leq \lambda$. Use JEP on $\bkzero$ and
Lemma~\ref{basicjep}.2 to find a common extension $C$ of both
$C_1,C_2$ with cardinality at most $\lambda$. Then consider the
structure $(A_1\cup A_2, B_1\cup B_2, C)$. By identifying $C_1$ and
$C_2$ with subsets of C, we can consider all existing edges as
$C$-colored. Then assign colors to new edges in a one-to-one way.
This is possible, since that there are no more than $\lambda$ many
edges and $\lambda$ many colors. Towards contradiction assume there
is a violation of $(*)$ witnessed by the edges $(l,l',r,r')$. If
there were three old edges among these then all four vertices would
be in $\Mscr_1$ or $\Mscr_2$. So there are two new edges, but the new
edges were colored, so there can not be two new edges among
(l,l',r,r') with the same color. Contradiction.
(2) For any $\lambda$, let $\M_i=(A_i,B_i,C_i)$, $i=1,2,3$, be three
models of $\hat \bK_0$ with cardinality $\lambda$ such that
$\M_1\subset \M_2,\M_3$ and there exist distinct $a_0,a_1,a_2\in
A_1$, distinct $c,c',c''\in C_1$, $b_2\in B_2$, and $b_3\in B_3$ such
that $F(a_k,b_l)=c$, for $k=0,1$ and $l=2,3$, $F(a_2,b_2)=c'$, and
$F(a_2,b_3)=c''$. (Extensions $\M_2$ and $\M_3$ of any $\M_1$ must
exist). Thus, in the disjoint amalgam of $\M_2$ and $\M_3$,
$a_0,a_1,b_2,b_3$ witness a violation to $(*)$. But in any amalgam of
$M_1,M_2,M_3$, the images of $b_2$ and $b_3$ must be distinct, and
thus, $a_0,a_1,b_2,b_3$ witness a violation to $(*)$ in any amalgam.
(3) By Corollary \ref{cor:TwoNoniso} there exist two maximal models
in $\kappa^+$; by Corollaries \ref{cor:LargerThanKappa} and
\ref{cor:NotMaximal} no model of size $>\kappa^+$ can be maximal. By
(1) no model of cardinality $\leq\kappa$ can be maximal.
Since there are two maximal models in $\kappa^+$, ${\hat \bK_0}$ fails
JEP$(\kappa^+)$ and JEP$(\le\lambda)$ for $\lambda \ge \kappa^+$.
To see that ${\hat \bK_0}$ fails JEP$(\lambda)$ for $\lambda>\kappa^+$,
consider a model $\M_0$ of type $(\lambda,\kappa)$ and a model $\M_1$ of type $(\lambda,\kappa)$.
By Corollary \ref{cor:LargerThanKappa}, $\M_0$ and $\M_1$ can not be jointly embedded into a model
in ${\hat \bK_0}$.
(4) is established by Corollary \ref{cor:LargerThanKappa}. The proof
of (5) is from Lemma \ref{Lem:LS(K)}
\end{proof}
\end{Thm}
Suppose $\bkzero$ has models in cardinality $\kappa$, but no larger,
and $\bkzero$ satisfies JEP$(\leq \kappa)$. It follows from Theorem
\ref{Thm:UncountableJEPAP} that ${\hat \bK_0}$ will satisfy
JEP$(\le\kappa)$ and have a maximal model in $\kappa^+$. This
condition on $\bkzero$ is very strong: there is a unique maximal
model in $\kappa$. However, examples of this sort (e.g. the
well-orderings of order type at most $\omega_1$ under end-extension)
are well-known.
\begin{question} Is there a complete sentence of
$L_{\omega_1,\omega}$? that has more than one maximal
model?\end{question}
\subsection{$ \mathbf 2^{\kappa^+}$ Nonisomorphic Maximal Models}
\label{sec:maximalmodel}
In this section we prove that the AEC given by Theorem
\ref{Thm:UncountableJEPAP} actually has $2^{\kappa^+}$ many
nonisomorphic maximal models in $\kappa^+$. We will build a family
of models of $\sigma_1$, each one starting with sets $A,B,C$, the
first two ordered as $\kappa^+$, $C$ ordered as $\kappa$, and with a
subset $C_0$ of $C$ that also has order type $\kappa$, and with $C
\setminus C_0$ has cardinality $\kappa$.
We are building models of the AEC $\hat \bK_0$ with vocabulary
$\hat \tau_0$ using an input AEC $\bK_0$ with vocabulary $\tau_0$ to
control the cardinality of the color sort. The key step in the
construction is to add new relations to the vocabulary $\hat \bK_0$
and use them to construct many models (in the expanded vocabulary).
But then, we show these relations are definable in
$L_{\kappa,\omega}({\hat \tau}_0)$ and deduce many ${\hat
\tau}_0$-models.
The proof goes in two steps. At the first step we ``code'' a linear
order of order type $\kappa$ on $C_0$. At the second step
we make use of this linear order to ``code'' $\kappa^+$ many subsets
of $\kappa$ into $A$. By varying the construction we get
$2^{\kappa^+}$ many nonisomorphic maximal models.
Recall that there exists a monochromatic 2-path (based) on some
$a_1,a_2\in A$, if there exists some $b\in B$, such that both edges
$(a_1,b)$ and $(a_2,b)$ have the same color.
\medskip
{\bf Step I} Code order:
Let $C$ be the set of colors and assume $C=\kappa$. Extend the
vocabulary ${\hat \tau}_0$ to ${\hat \tau}_1$ by including a unary
symbol $C_0$ and a binary symbol $<$. $C_0$ will be a subset of $C$
and $<$ will be a linear order on $C_0$ of order type $\kappa$. The
goal is to build a model as in Lemma \ref{Lem:ddagModel}, but this
time certain 2-paths are disallowed. In particular, for all
$\alpha<\kappa$ there exist two elements $l^\alpha_1,l^\alpha_2\in A$
and the 2-paths on $l^\alpha_1,l^\alpha_2\in A$ can not use any of
the colors $\{\beta|\beta\le\alpha\}$. Any other color is allowed.
The resulting model is maximal, as seen by Corollary
\ref{cor:StepImax}.
\begin{Lem}\label{Lem:Step1} There is a ${\hat \tau}_1$-model $\M$ that satisfies all the following conditions.
\begin{enumerate}\setcounter{enumi}{-1}
\item $C^\M \restriction \tau_{\bK_0} \in \bK_0$.
\item $\M \restriction \tau_0$ is a $(\kappa^+,\kappa^+)$-model
of $\sigma_1$ and $C=\kappa$.
\item $C_0$ is a subset of $C$ such that $|C_0|=|C\setminus C_0|=\kappa$ and $<$ is an order on $C_0$ of order type $\kappa$. We may refer to the
elements of
$C_0$ using ordinals $<\kappa$.
\item $<$ is void outside $C_0$.
\item For every $\alpha\in C_0$, there exist two elements $l^\alpha_1,l^\alpha_2\in A$ such
that there exists a 2-path on $l^\alpha_1,l^\alpha_2$ colored by $c$ if and only if $c>\alpha$ or $c \in C\setminus C_0$.
\item For distinct $\alpha,\alpha'\in C_0$, the elements $l^\alpha_1,l^\alpha_2, l^{\alpha'}_1,l^{\alpha'}_2$ are all distinct.
\item For every pair $(a_1,a_2)$ in $A$ and for all $c\in C$, there exists a 2-path on $a_1,a_2$ colored by $c$, unless it is forbidden by clause $(4)$.
\end{enumerate}
\begin{proof} We now construct the model. Let $A,B,C, C_0$ be as in the first paragraph of
Section~\ref{sec:maximalmodel} and order $C_0$ by $<$ so that the
requirements of clauses $(2)$ and $(3)$ are met. For every
$\alpha<\kappa$, select two elements $l^\alpha_1,l^\alpha_2\in
\kappa$ so that $\alpha\neq\alpha'$ implies
$\{l^\alpha_1,l^\alpha_2\}\cap\{l^{\alpha'}_1,l^{\alpha'}_2\}=\emptyset$.
The rest of the proof is similar to the proof of Lemma
\ref{Lem:ddagModel}, the only difference is that for every
$\alpha<\kappa$, the pair $l^\alpha_1,l^\alpha_2$ given by clause
$(4)$ do not have a 2-path with any color $c\le\alpha$. The rest of
the argument remains the same.
\end{proof}
\end{Lem}
A priori, the conditions in Lemma~\ref{Lem:Step1} are not ${\hat
\tau}_0$-invariant. We show in Lemma~\ref{Lem:LoneIso} that $C_0$ and
$<$ are definable in $L_{\kappa,\omega}({\hat \tau}_0)$ so they are.
\begin{Cor}\label{cor:StepImax} The models that satisfy the requirements of Lemma \ref{Lem:Step1} are
$C$-maximal.
\begin{proof} Since the set of all $l^\alpha_1,l^\alpha_2$ has size $\kappa$, the result follows from Corollary \ref{cor:maximal2}.
\end{proof}
\end{Cor}
\begin{Lem}\label{Lem:LoneIso} Let $\M_1$ and $\M_2$ be two ${\hat \tau}_1$-models
that satisfy the conditions of Lemma \ref{Lem:Step1} and let
$\M_1|_{{\hat \tau}_0},\M_2|_{{\hat \tau}_0}$ be their reducts to
vocabulary ${\hat \tau}_0$. Then any isomorphism $i$ between
$\M_1|_{{\hat \tau}_0}$ and $\M_2|_{{\hat \tau}_0}$ is also an
isomorphism of $\M_1,\M_2$ (as ${\hat \tau}_1$-structures).
We will refer to this property as ``the ${\hat \tau}_0$-isomorphisms
respect $C_0,<$''.
\begin{proof} We claim that both $C_0,<$ are definable in the original structure $\Mscr$ by a
sentence of an appropriate infinitary language in vocabulary ${\hat
\tau}_0$, and therefore, preserved by ${\hat \tau}_0$-isomorphisms.
First, $C_0(x)$ is defined by ``there exists a pair
$(l^{\alpha}_1,l^{\alpha}_2)\in A$ such that there is no 2-path on
$l^\alpha_1,l^\alpha_2$ colored by $x$''.
Second, for each ordinal $\alpha<\kappa$, let $\alpha^{\M}$ denote the $\alpha^{th}$ element of the order $<^{\M}$.
Since $<$ has order type $\kappa$, the specification makes sense. We prove by induction on $\alpha<\kappa$
that $\alpha^{\M}$ is defined by a formula $\phi_\alpha(x)$ in
$L_{{\kappa},\omega}$ in the vocabulary ${\hat \tau}_0$.
$\phi_0(x)$: There exist two points $a,a'$ with no 2-path colored
$x$. But for every other color $c\neq x$, there is a $c$-colored 2-path on $a,a'$.
$\phi_{\alpha}(x)$: There exist two points $a,a'$ with no 2-path
colored by $x$ or by any color $y$ satisfying $\bigvee_{\beta <\alpha}
\phi_\beta(y)$. But for every other color $c\neq x$ and $\bigwedge_{\beta<\alpha} \neg\phi_\beta(c)$, there is a $c$-colored 2-path on $a,a'$.
Now $<$ is defined by a $L_{\kappa,\omega}$-formula in the vocabulary ${\hat \tau}_0$.
$$ x < y \mbox{ \rm if and only if \ } \bigvee_{\alpha < \beta <
\kappa} \phi_{\alpha}(x) \wedge \phi_{\beta}(y).$$
Since each $\phi_\alpha(x)$ is a formula in vocabulary ${\hat
\tau}_0$, this proves the result.
\end{proof}
\end{Lem}
It also follows by a similar argument that the elements
$l^\alpha_1,l^\alpha_2$ are definable by a formula in
$L_{\kappa,\omega}$ in the vocabulary ${\hat \tau}_0$. Consider the
formula $\phi(x,y)$: ``there exists a 2-path on $x,y$ colored by $c$
if and only if $\neg \bigvee_{\beta\le\alpha} \phi_\beta(c)$. ''. By
clauses $(4)$ and $(6)$ of Lemma \ref{Lem:Step1}, $\phi(x,y)$ holds
if and only if $\{x,y\}=\{l^\alpha_1,l^\alpha_2\}$. So, any ${\hat
\tau}_0$-isomorphism must preserve the two-element subset
$\{l^\alpha_1,l^\alpha_2\}$, for all $\alpha<\kappa$.
\medskip { \bf Step II} Code subsets:
Recall that ${\hat \tau}_1={\hat \tau}_0\cup\{C_0,<\}$ and extend
${\hat \tau}_1$ to ${\hat \tau}_2$ by including a new binary symbol
$S$. $S$ will be a binary relation that codes subsets
$A_0=\{m_\alpha|\alpha<\kappa^+\}$ of $A$ by elements of $C_0$.
We
also require that the set $\{l^\alpha_i|\alpha<\kappa,i=1,2\}$ from
Step I and the set $A_0$ from Step II are disjoint. Using $S$ we can
assign to each $m_\alpha\in A_0$ a distinct subset $S_\alpha$ of
$C_0$. The goal is to build a model that satisfies all the
restrictions from Step I, plus more 2-paths are forbidden. In
particular, for each $\alpha<\kappa^+$ the 2-paths based on
$m_0,m_\alpha$ can not use any of the colors in $S_\alpha$. Every
other color is allowed. Once again, the resulting model is maximal
(see Corollary \ref{cor:StepIImax}).
We again add predicates, this time to code models, and then prove
they are $L_{\kappa,\omega}({\hat \tau}_0)$ definable.
\begin{Lem}\label{Lem:Step2} There is an ${\hat \tau}_2$-model $\N$ that satisfies all the following conditions.
\begin{enumerate}
\item Clauses $(1)-(5)$ from Lemma \ref{Lem:Step1} hold.
\item There is a set $A_0 =\{ m_\alpha:{\alpha<\kappa^+}\}\subset A$ such that $|A\setminus A_0 |=\kappa^+$ and $A_0$ is disjoint from $\{l^\alpha_i|\alpha<\kappa, i=1,2\}$.
\item $S(x,y)$ is a binary relation on $A_0 \times C_0$. Denote the set $\{y\in C_0|S(m_\alpha,y)\}$ by $S_\alpha$.
\item The $S_\alpha$'s are distinct subsets of $C_0$. For all $\alpha$, $|S_\alpha|=|C_0\setminus S_\alpha|=\kappa$, and $0$ does not belong to any $S_\alpha$.
\item For all $0<\alpha<\kappa^+$, there exists a 2-path on $m_0,m_\alpha$ colored by $c$ if and only if $c\in S_\alpha$.
\item For all $a_1,a_2\in A$ and for all $c$, there exists a 2-path on $a_1,a_2$ colored by $c$, unless it is forbidden by clause $(5)$ of this Lemma
or by clause
$(4)$ of Lemma\ref{Lem:Step1}.
\end{enumerate}
\begin{proof} The proof is similar to the proof of Lemma \ref{Lem:Step1} and is left to the reader.
\end{proof}
\end{Lem}
\begin{Cor}\label{cor:StepIImax} The models that satisfy the requirements of Lemma \ref{Lem:Step2} are $C$-maximal.
\begin{proof} Since $|A\setminus (A_0\cup \{l^\alpha_i|\alpha<\kappa, i=1,2\})|=\kappa^+$, the result follows from Corollary \ref{cor:maximal2}.
\end{proof}
\end{Cor}
\begin{Lem}\label{Lem:LtwoIso} Let $\N_1$ and $\N_2$ be two ${\hat \tau}_2$-models that satisfy the conditions of Lemma \ref{Lem:Step2} and let
$\N_1|_{{\hat \tau}_0},\N_2|_{{\hat \tau}_0}$ be their reducts to
vocabulary ${\hat \tau}_0$. Then any isomorphism $i$ between
$\N_1|_{{\hat \tau}_0}$ and $\N_2|_{{\hat \tau}_0}$ is also an
isomorphism of $\N_1,\N_2$ (as ${\hat \tau}_2$-structures).
\begin{proof} From Step I we know that $C_0$ and $<$ are definable by $L_{\kappa,\omega}({\hat \tau}_0)$-formulas.
We prove that the same is true for the set $A_0
=\{m_\alpha|\alpha<\kappa^+\}$ and the sets $S_\alpha$,
$\alpha<\kappa^+$. The element $m_0$ is defined by the following
$L_{\kappa,\omega}({\hat \tau}_0)$-formula $\psi_0(x)$.
$\psi_0(x)$: there exist two distinct elements $m_1,m_2\in A$ and two distinct colors $c_1,c_2\in C_0$ and
there is no 2-path based on $x,m_1$ colored
by $c_1$,
and there is no 2-path based on $x,m_2$ colored by $c_2$.
We now show the set $\{m_\alpha|0<\alpha<\kappa^+\}$ is defined by
the formula $\psi_1(x)$.
$\psi_1(x)$: there exists some $y\neq x$ such that $\psi_0(y)$, i.e. $y$ equals $m_0$, and there exists a color $c\in
C_0$ and there is no 2-path based on $y,x$ colored by $c$.
Then $\psi_1(x)$ holds if and only if $x$ belongs to $\{m_\alpha|0<\alpha<\kappa^+\}$. Note that $\psi_1$ defines the whole set
$\{m_\alpha|0<\alpha<\kappa^+\}$,
but not the order of the $m_\alpha$'s in this set. Nevertheless, for every $\alpha<\kappa^+$, the set $S_\alpha$ is definable by the following formula
$\psi_2$ which uses $m_\alpha$ as a parameter; $\psi_2$ is a reformulation of clause $(5)$ from Lemma \ref{Lem:Step2}.
$\psi_2(x,m_\alpha)$: there exists some y such that $\psi_0(y)$ and there exists a 2-path on $y,m_\alpha$ colored by $x$.
Then $\psi_2(x,m_\alpha)$ holds if and only if $x\in S_\alpha$.
Since all these sentences are in vocabulary ${\hat \tau}_0$, this
finishes the proof.
\end{proof}
\end{Lem}
Now fix $Y$ to be some subset of $\kappa^+$ and vary the construction
of Lemma \ref{Lem:Step2} so that for each $0<\alpha<\kappa^+$, $0\in
S_\alpha$ if and only if $\alpha\in Y$. Call the corresponding
${\hat \tau}_2$-structure $\N_Y$. If $Y_1,Y_2$ are two distinct
subsets, then $\N_{Y_1}$ and $\N_{Y_2}$ are easily seen to be
nonisomorphic as ${\hat \tau}_2$-structures. By Lemma
\ref{Lem:LtwoIso}, their ${\hat \tau}_0$-reducts are also
nonisomorphic, which proves the following.
\begin{Thm}\label{Thm:ManyMaxModels} If $\bkzero$ is an AEC that has models in cardinality $\kappa$ but no larger, then ${\hat \bK_0}$ from Theorem
\ref{Thm:UncountableJEPAP} has $2^{\kappa^+}$-many nonisomorphic maximal models of type $(\kappa^+,\kappa^+)$.
\end{Thm}
In the next section we give three applications of Theorem \ref{Thm:ManyMaxModels}.
\subsection{Maximal Models in Many
Cardinalities}\label{sec:ManyCardinalities}
If $\kappa<\lambda<\beth_{\omegaone}$ are two characterizable
cardinals (Definition \ref{Def:CharacterizableCard}) and
$\bK_{\kappa}$, $\bK_{\lambda}$ the corresponding AEC (in disjoint
vocabularies), then the union\footnote{By the union of AEC's with
disjoint vocabularies we mean the collection of structures in the
union of the vocabularies, where the obvious symbols have the empty
interpretation, and one model is a strong substructure of another if
the same is true for their reducts to the vocabulary where the structures are non-trivial.} with an AEC with arbitrarily large models is an AEC (
with strong substructure being in the appropriate vocabulary) with
maximal models in $\kappa$ and $\lambda$
and arbitrarily large models. However, the union is a hybrid AEC which fails JEP in all cardinals.
If $<\lambda_i|i\le\alpha<\aleph_1>$ is a strictly increasing sequence
of characterizable cardinals (Definition
\ref{Def:CharacterizableCard}), we provide an example of a pure
(Definition~\ref{Def:PureAEC}) AEC with maximal models in
cardinalities $<\lambda_i^+|i\le\alpha<\aleph_1>$, arbitrarily large
models, and JEP$(<\lambda_0)$ holds.
For any triple $(A,B,C)$ there is a first order sentence $\sigma_1$
asserting that $A,B$ form a bipartite graph with $C$ many colors that
contains no monochromatic $ K_{2,2}$ subgraph (see property $(*)$).
The structures constructed below will contain many substructures
satisfying this requirement. Rather than cluttering the paper with a
careful description of the formal sentence (with different ternary
relations for each colored graph) we will just assert where
$\sigma_1$ holds.
We begin with the case of two cardinals.
\begin{Lem}\label{2max} Let $\kappa<\lambda$ and let $(\bkzero^k,\prec_k)$ be an AEC in vocabulary $\tau^k$ with models in $\kappa$ but no higher,
and let $(\bkzero^\ell,\prec_\ell)$ be an AEC in vocabulary
$\tau^\ell$ with models in $\lambda$ but no higher. If both
$(\bkzero^k,\prec_k)$ and $(\bkzero^\ell,\prec_\ell)$ satisfy
JEP$(<\kappa)$, then there is an AEC $(\bK^*,\prec_{\bK^*})$ which
\begin{enumerate}
\item satisfies JEP$(<\kappa)$;
\item fails AP in all infinite cardinals;
\item has $2^{\kappa^+}$ non-isomorphic maximal models in
$\kappa^+$, $2^{\lambda^+}$ non-isomorphic maximal models in
$\lambda^+$, but no maximal models in any other cardinality,
while JEP fails in all $\lambda \geq \kappa$;
\item has arbitrarily large models; and
\item $LS(\bK^*)=\max\{LS(\bkzero^k),LS(\bkzero^\ell)\}$.
\end{enumerate}
If both $(\bkzero^k,\prec_k)$ and $(\bkzero^\ell,\prec_\ell)$ are
pure, then $(\bK^*,\prec_{\bK^*})$ is pure. If both
$(\bkzero^k,\prec_k)$ and $(\bkzero^\ell,\prec_\ell)$ are definable
by an $\lomegaone$-sentence, then the same is true for
$(\bK^*,\prec_{\bK^*})$.
\begin{proof} Let $\bK^*$ be the AEC defined by the following construction.
The construction contains 4 bipartite graphs entangled together.
Recall that a bipartite graph is a $\tau_0$ structure and
$\sigma_1$ is a $\tau_0$ sentence.
\begin{enumerate}[a)]
\item $A_1,A_2,A_3, C_1,C_2$ are non-empty and partition the universe.
\item The structures $(A_1,A_2,C_1)$, $(A_1,A_3,C_1)$,
$(A_1,C_2,C_1)$, and $(A_2,A_3,C_2)$ are colored bipartite
graphs satisfying $\sigma_1$.
\item $C_1$ is a model in $\bkzero^k$ and $C_2$ is a model in $\bkzero^\ell$. In particular $|C_1|\le\kappa$ and
$|C_2|\le\lambda$
\end{enumerate}
Define for $\M,\N\in \bK^*$, $\M\prec_{\bK^*}\N$, if $\M\subset \N$
with respect\footnote{We abuse notation here; depending on the exact
location the colored graph will be with respect to a different
ternary relation; but we will think of it as a structure modeling the
appropriate translation of $\sigma_1$.} to $\tau_0$, $C_1^{\M}\prec_k
C_1^{\N}$ and $C_2^{\M}\prec_\ell C_2^{\N}$.
(1) Fix $\chi < \kappa$ and let
$\M_1=(A^1_1,A^1_2,A^1_3,C^1_1,C^1_2)$,
$\M_2=(A^2_1,A^2_2,A^2_3,C^2_1,C^2_2)$ be two models in $\bK^*$ such that both $M_1,M_2$ have size $\le \chi$.
Use JEP$(<\kappa)$ on $\bkzero^\ell$ to extend the
$\tau_\ell$-structures $C^1_2,C^2_2$ to a common structure
$\check{C}_2$ which has cardinality $\chi$. Use the argument of
Theorem~\ref{Thm:UncountableJEPAP}.1 to find a common embedding of
$(A^1_2,A^1_3,C^1_2)$ and $(A^2_2,A^2_3,C^2_2)$ with domain
$(A^1_2\cup A^2_2,A^1_3\cup A^2_3,\check{C}_2)$ with cardinality
$\chi$. Note that the proof of Theorem~\ref{Thm:UncountableJEPAP}.1
does not add any vertices to the graph. Then use JEP$(<\kappa)$ in
$\bK_0^k$ to find a common extension $\check{C}_1$ of $C^1_1$ and
$C^2_1$ of cardinality $\chi$. Now consider the structures
$(A^1_1,A^1_2,C^1_1)$, $(A^1_1,A^1_3,C^1_1)$,
$(A^1_1,\check{C}_2,C^1_1)$ and $(A^2_1,A^2_2,C^2_1)$,
$(A^2_1,A^2_3,C^2_1)$, $(A^2_1,\check{C}_2,C^2_1)$. Apply the
argument of Theorem~\ref{Thm:UncountableJEPAP}.1 again several times
to find an $\bK^*$ extension of all these models with domain
$(A^1_1\cup A^2_1, A^1_2 \cup A^2_2, A^1_3\cup A^2_3,
\check{C}_1,\check{C}_2)$. Exactly as in
Theorem~\ref{Thm:UncountableJEPAP}.1 we verify this structure is in
$\bK^*$.
(2)The proof for AP follows as in Theorem \ref{Thm:UncountableJEPAP}.
(3) and (4) Assume that $C_1$ has size $\kappa$. By Lemma
\ref{Lem:BasicCombinatorics}, if $A_1$ has size $\kappa^+$, then
$A_2,A_3,C_2$ have size $\leq \kappa^+$ and by Theorem
\ref{Thm:ManyMaxModels} there are $2^{\kappa^+}$ many non-isomorphic
maximal models in $\kappa^+$. If $A_1$ has size $>\kappa^+$, then
$A_2,A_3,C_2$ have size $\leq \kappa$, and notice that the size of $A_1$ can be arbitrarily
large. If $A_1$ has size $\kappa$, then the sizes of
$A_2,A_3,C_2$ can be greater than $\kappa^+$.
Repeating the same argument, assume that $C_1$ and $A_1$ have size
$\kappa$, and $C_2$ has size $\lambda$. If $A_2$ has size
$\lambda^+$, then $A_3$ has size $\leq\lambda^+$ and by Theorem
\ref{Thm:ManyMaxModels} again, there are $2^{\lambda^+}$ many
nonisomorphic maximal models in $\lambda^+$. If $A_2$ (or $A_3$) has
size $\lambda$, then $A_3$ (respectively $A_2$) can have any size and
we get arbitrarily large models.
The failure of JEP in $\lambda\geq \kappa$ now fails as
Theorem~\ref{Thm:UncountableJEPAP}.
(5) The argument is similar to the proof of Lemma \ref{Lem:LS(K)}.
Let $M$ be a model in $\bK^*$ and $X$ be a subset of $M$. Find some
$\check{C}_1\in \bkzero^k$ such that $X\cap C_1^M\subset \check{C}_1$
and $|\check{C}_1|=|X\cap C_1^M|+LS(\bkzero^k)$. Then find some
$\check{C}_2\in\bkzero^\ell$ such that $X\cap C_2^M\subset
\check{C}_2$ and $|\check{C}_2|=|X\cap C_2^M|+LS(\bkzero^\ell)$. Let
$M_0=X\cup \check{C}_1\cup \check{C}_2$. Then $M_0$ belongs to ${\hat
\bK_0}$. Indeed, $M_0$ satisfies $\sigma_1$, since any violations of
$(*)$ in $M_0$ would be violations of $(*)$ in $M$ too.
Contradiction. Considering that
$|M_0|\le|X|+|\check{C}_1|+|\check{C}_2|\le|X|+LS(\bkzero^k)+LS(\bkzero^\ell)=|X|+\max\{LS(\bkzero^k),LS(\bkzero^\ell)\}$,
it follows that $LS(\bK^*)=\max\{LS(\bkzero^k),LS(\bkzero^\ell)\}$.
Finally observe that the conjunction of (a)-(c) is expressible by an $\lomegaone$-sentence if and only if membership in both $\bkzero^k$
and $\bkzero^\ell$ is expressible by an $\lomegaone$-sentence.
\end{proof}
\end{Lem}
We sketch a minor variant in the argument to extend this to
infinitely many cardinals.
\begin{Thm}\label{manymax} Let $\langle \lambda_i: i\le \alpha \rangle$ be a strictly increasing
sequence of cardinals. Assume that for each $i\le\alpha$, there exists an AEC $(\bkzero^i,\prec_{\bkzero^i})$ with models in $\lambda_i$ but no higher.
Then if all $(\bkzero^i,\prec_{\bkzero^i})$ satisfy JEP$(<\lambda_0)$,
there is an AEC $(\bK^*,\prec_{\bK^*})$ which
\begin{enumerate}
\item satisfies JEP$(< \lambda_0)$, while JEP fails for all larger cardinals;
\item fails AP in all infinite cardinals;
\item there exist $2^{\lambda_i^+}$ many nonisomorphic maximal
models in $\lambda_i^+$, for all $i\le\alpha$, but
no maximal models in any other cardinality;
\item has arbitrarily large models; and
\item $LS(\bK^*)=\max\{LS(\bkzero^i)|i\le\alpha\}$.
\end{enumerate}
If all $(\bkzero^i,\prec_{\bkzero^i})$ are pure, then
$(\bK^*,\prec_{\bK^*})$ is pure. Further, if $\alpha < \aleph_1$ and
all $(\bkzero^i,\prec_{\bkzero^i})$ are definable by an
$\lomegaone$-sentence, then the same is true for
$(\bK^*,\prec_{\bK^*})$.
\begin{proof} Let $\bK^*$ be the AEC defined by the following construction.
\begin{enumerate}[a)]
\item The sets $(A_i|i\le\alpha)$ and $(C_i|i<\alpha)$ are non-empty and partition
the universe.
\item For each $i,j$ with $i<j\le\alpha$, the triples
$(A_i,A_j,C_i)$ and $(A_i,C_j,C_i)$ satisfy $\sigma_1$.
\item For each $i\le\alpha$, $C_i$ is a model in $\bkzero^i$, which
implies that $|C_i|\le\lambda_i$.
\end{enumerate}
Define for $\M,\N\in \bK^*$, $\M\prec_{\bK^*}\N$, if $\M\subset \N$
with respect to $\tau_0$ and $C_i^{\M}\prec_{\bkzero^i} C_i^{\N}$,
for all $i\le\alpha$.
The proof is like the proof of Theorem \ref{2max} with some easy modifications.
Observe that if for some $i$, $|C_i|=\lambda_i$ and $|A_i|=|C_i|^+$, then by Lemma
\ref{Lem:BasicCombinatorics} all $A_j$, $C_j$, $j> i$, are ``locked'' to have
size at most $|A_i|$, and by Theorem \ref{Thm:ManyMaxModels} there
are $2^{\lambda_i^+}$ many nonisomorphic maximal models in
$\lambda_i^+$. If $|A_i|=|C_i|$, then the cardinalities of $A_j,C_j$,
$j>i$ can be greater than $\lambda_i^+$. We leave the rest of the details to the reader.
\end{proof}
\end{Thm}
We need some background before getting specific applications
of the previous theorem. The next fact follows from Theorem 4.20 of \cite{BKL}
for $\aleph_r = \kappa$, noting that joint embedding holds in
$\aleph_{r-1}$. Indeed, $2$-AP in $\aleph_{r-2}$ implies $2$-AP of
models with cardinality $\aleph_{r-1}$ over models of cardinality
$<\aleph_{r-1}$ (or the empty set). This yields a complete sentence
$\phi_r$ whose class of models is denoted $At^r$; a similar argument
for the incomplete sentence with models $\hat \bK^r$ is in Theorem
4.3 of that paper.
\begin{Fact}\label{nicecharsm} Every cardinal $\kappa< \aleph_\omega$ is
characterized by a (complete) sentence of $L_{\omega_1,\omega}$ that
satisfies JEP$(< \kappa)$.
\end{Fact}
We describe the next example, based on \cite{Morley65a}, in detail
since the particular formulation is important.
\begin{Ex} Fix some countable ordinal $\alpha$ and let $\{\beta_n|n\in\omega\}$ list the ordinals less than
$\alpha$. Consider the vocabulary $\tau$ that contains a binary
relation $\in$, a unary function $r$ (for `rank') and constants
$(c_{\beta_n})_{n\in\omega}$. Let $\phi_\alpha$ be the conjunction of
the following:
\begin{itemize}
\item $\forall x,\; x\in c_{\beta_n}\leftrightarrow
\bigvee_{\beta_i\in\beta_n} x=c_{\beta_i}$, for each $n$;
\item $\forall x,\; \bigvee_{n\in\omega} r(x)=c_{\beta_n}$;
\item $r(c_{\beta_n})=c_{\beta_n}$, for each $n$;
\item $\forall x,y,\; x\in y \rightarrow r(x)\in r(y)$; and
\item $\forall x,y,\; (\forall z)((z\in x\leftrightarrow z\in
y)\rightarrow x=y)$ (Extensionality).
\end{itemize}
Observe that $\phi_\alpha$ is an $\lomegaone(\tau)$-sentence. Let
$\bK_\alpha$ be the collection of all models of $\phi_\alpha$. If
$M\in \bK_\alpha$, then $M$ can be embedded into $V_\alpha$. In
particular, $|M|\le|V_\alpha|=\beth_\alpha$.
\end{Ex}
\begin{Fact}\label{Cor:MorleyAEC}
For each $\alpha<\omegaone$, $(\bK_\alpha, \subseteq)$ satisfies the following.
\begin{enumerate}[(a)]
\item $\bK_\alpha$ has a unique maximal model in cardinality
$\beth_\alpha$, and no larger models;
\item JEP$(\le\beth_\alpha)$ holds; and
\item $LS(\bK_\alpha)=\aleph_0$.
\end{enumerate}
\end{Fact}
Note that under GCH up to $\aleph_\omega$,
Fact~\ref{Cor:MorleyAEC} is stronger than
Fact~\ref{nicecharsm} since JEP$(<\kappa)$ is replaced by
JEP$(\leq\kappa)$.
\begin{Cor}\label{manymaxap} Here are three applications of
Theorem~\ref{manymax}.
\begin{enumerate}
\item If $<\lambda_i|i\le\alpha\le\omega>$ is any increasing sequence of cardinals below
$\aleph_\omega$, then there exists an $\lomegaone$ sentence $\psi$
\begin{enumerate}
\item whose models satisfy JEP$(<\lambda_0)$;
\item that fails AP in all infinite cardinals;
\item has $2^{\lambda_i^+}$ many nonisomorphic maximal models
in $\lambda_i^+$, for all $i\le\alpha$, but no maximal
models in any other cardinality, while JEP fails for
all larger cardinals; and
\item has arbitrarily large models.
\end{enumerate}
\item If $<\beth_{\alpha_i}|i\le\gamma<\omegaone>$ is a strictly increasing sequence,
then there exists an $\lomegaone$ sentence $\psi'$
\begin{enumerate}
\item whose models satisfy JEP$(\le\beth_{\alpha_0})$;
\item fails AP in all infinite cardinals;
\item has $2^{\beth_{\alpha_i}^+}$ many nonisomorphic maximal
models in $\beth_{\alpha_i}^+$, for all $i\le\gamma$, but
no maximal models in any other cardinality, while JEP
fails for all larger cardinals; and
\item has arbitrarily large models.
\end{enumerate}
\item If $<\lambda_i|i\le\alpha\le\omega>$ is any countable increasing sequence of
cardinals below $\beth_{\omega_1}$ that are characterized by
complete $\lomegaone$ sentences,
then there exists an $\lomegaone$-sentence $\psi''$
\begin{enumerate}
\item whose models satisfy JEP$(\aleph_0)$;
\item fails AP in all infinite cardinals;
\item has $2^{\lambda_i^+}$ many nonisomorphic maximal models
in $\lambda_i^+$, for all $i\le\alpha$, but no maximal models
in any other cardinality; and
\item has arbitrarily large models.
\end{enumerate}
\end{enumerate}
\end{Cor}
\begin{proof} For 1) use Theorem~\ref{manymax} and
Fact~\ref{nicecharsm}. For 2) use Theorem~\ref{manymax} and
Fact~\ref{Cor:MorleyAEC}. 3) is easy by Theorem~\ref{manymax} since
every complete sentence satisfies JEP in $\aleph_0$.
\end{proof}
\begin{question} Is there an $\lomegaone$-sentence that has maximal models in uncountably many
cardinals but arbitrarily large models?
\end{question}
|
\section{Introduction}
After Kibble's investigation \cite{kibble1979geometrization} of the geometric properties of quantum state spaces, geometric formulations of quantum dynamics have been attracting much attention over the last decades \cite{Anandan91,AnAh90,AshAb95,BoCaGra91,BroHu01,ChiMel2012,Chrus94,CleMar08,FaKuMaMa10,Grig92,Montgomery91,SaHuKu2011,KhaBroGla01,uhlmann2009geometry}. In turn, geometric quantum dynamics has opened several modern perspectives: for example, Fubini-Study geodesics have been introduced in Grover's quantum search algorithms \cite{MiWa01,Zhao2012} and in time-optimal quantum control \cite{CarHoKoOku06,CarHoKoOku07,Dalessandro01}, while the holonomy features arising from the quantum geometric phase \cite{AhAn87,BoCaGra91} and its non-Abelian extensions \cite{Anandan88non,Chrus94} have been proposed in quantum computation algorithms \cite{GunWaNa2014,Lucarelli2005,TaNaHa2005}.
Most approaches deal with pure quantum states and involve the geometry of the Hopf fibration
\begin{align*}
S(\mathscr{H})&\to \mathbf{P}_{\!}\mathscr{H}
\\
\psi&\mapsto\psi\psi^\dagger\,,
\end{align*}
where $S(\mathscr{H})$ denotes the unit sphere in a complex Hilbert space $\mathscr{H}$ (so that $\psi\in S(\mathscr{H})$ is a unit vector in $\mathscr{H}$), while $ \mathbf{P}_{\!}\mathscr{H}$ is the corresponding projective space containing the projections $\rho_\psi:=\psi\psi^\dagger$. The geometry of the above Hopf bundle is well known and has been widely studied in the finite dimensional case $\mathscr{H}=\Bbb{C}^n$, although some studies extend to consider infinite-dimensional Hilbert spaces \cite{ChiMel2012}. In the finite dimensional case, one can emphasize the symmetry properties of the Hopf bundle by writing
\begin{equation}\label{bundles}
S(\Bbb{C}^n)=\mathcal{U}(n)/\mathcal{U}(n-1)
\,,\qquad
\mathbf{P}\Bbb{C}^n=S(\Bbb{C}^n)/\mathcal{U}(1)=\mathcal{U}(n)/\big(\mathcal{U}(n-1)\times \mathcal{U}(1)\big)
\,.
\end{equation}
The interplay between the geometry of the Hopf bundle and its symmetries is the basis of geometric quantum dynamics. For example, the emergence of principal bundles leads to the usual horizontal-vertical decomposition in terms of a principal connection that is strictly related to Berry's geometric phase.
This is a beautiful picture, whose symplectic Hamiltonian properties have been widely investigated after Kibble's work \cite{kibble1979geometrization}.
In this paper, we aim to present how this geometric framework emerges naturally from the unitary symmetry properties of quantum variational principles. Time-dependent variational approaches have been most successful in chemical physics (here, we recall the celebrated Car-Parrinello model in molecular dynamics \cite{CarPa85}). The most fundamental quantum variational principle is probably due to Dirac and Frenkel (DF) \cite{dirac1930note,frenkel1934wave}. This action principle produces Schr\"odinger equation $i\hbar\dot\psi=H\psi$ as the Euler-Lagrange equation associated to
\[
\delta\int_{t_1}^{t_2\!}\!\big\langle\psi,i\hbar\dot\psi-{H}\psi\big\rangle\,{\rm d} t=0
\,,
\]
where $H$ is the quantum Hamiltonian operator and we introduce the pairing $\langle A,B\rangle$ and the inner product
\[
\langle A|B\rangle=\operatorname{Tr}(A^\dagger B)
\,,\qquad\text{ so that } \qquad
\langle A,B\rangle:={{Re}}\,\langle A|B\rangle
\]
and ${{Im}}\,\langle A|B\rangle=\langle iA,B\rangle$.
Various properties of the variational principle above have been studied over the decades \cite{LowMuk72,kramer1981geometry,Ohta2000}, after it was first proposed in the context of Hartree-Fock mean field theories. For example, it is known that the DF action principle is simply the quantum correspondent of the classical Hamilton's principle on phase space $\delta\!\int_{t_1}^{t_2}\!\big(\boldsymbol{p}\cdot\dot{\boldsymbol{q}}-{H(\boldsymbol{q,p})}\big)\,{\rm d} t=0$, so that $\hbar\langle\psi,i{\rm d}\psi\rangle$ acquires the meaning of canonical one form on $\mathscr{H}$ (analogously, $\boldsymbol{p}\cdot{\rm d}{\boldsymbol{q}}$ is the canonical one form in classical mechanics). However, to our knowledge, an investigation of the geometric symmetry properties of quantum variational principles has not been carried out. Although an early attempt was proposed in \cite{kramer1981geometry}, the emergence of the Hopf bundle in this context has not been presented so far. For example, the momentum maps associated to quantum variational principles have never been considered in the literature, while they are essential geometric features often associated to fundamental physical quantities. Even in the simplest situation, the phase invariance of quantum Lagrangians produces the momentum map identifying the total quantum probability $\|\psi\|^2$.
The present work applies well known techniques in the theory of geometric mechanics \cite{marsden1999introduction}, which focus on the symmetry properties of the dynamics. For example, momentum map structures are seen to emerge: some are new, while others are related to the principal connections associated to quantum geometric phases. Within geometric mechanics, we shall be using the specific tool of Euler-Poincar\'e theory \cite{holm1998euler} that typically applies to variational principles with symmetry. As we shall see, besides recovering well known relations, these theory allows us to formulate new variational principles for various quantum descriptions, such as the Liouville-Von Neumann equation, Heisenberg dynamics, Moyal-Wigner formulation on phase space and the Ehrenfest theorem for the evolution of expectation values. Some of these descriptions of quantum mechanics have been lacking a variational structure, which is now provided in this paper for the first time. Here, we shall not dwell upon various complications that may emerge in infinite dimensional Hilbert spaces $\mathscr{H}$ and we assume convergence where necessary. When convenient, we shall consider dynamics on finite dimensional spaces and rely on the possibility of extending the results to the infinite dimensional case.
\section{Euler-Poincar\'e variational principles in the Schr\"odinger picture\label{SEC:EPSchr}}
This section presents the Euler-Poincar\'e formulation of quantum dynamics in the \text{Schr\"odinger}\, picture. Two main examples are considered: the Schr\"odinger equation as it arises from the Dirac-Frenkel theory and the Fubini-Study geodesics. Their geometric features will be analyzed in terms of momentum maps.
\subsection{Euler-Poincar\'e reduction for pure quantum states}
Upon denoting by $T\!\mathscr{H}$ the tangent bundle of the Hilbert space $\mathscr{H}$, consider a generic Lagrangian
\begin{equation}\label{genlagr}
L:T\!\mathscr{H}\to\Bbb{R}\,,\qquad
L=L(\psi,\dot\psi)\,,
\end{equation}
so that the assumption of quantum evolution restricts $\psi$ to evolve under the action of that unitary group $\mathcal{U}(\mathscr{H})$, that is
\begin{equation}\label{psievol}
\psi(t)=U(t)\psi_0\,,\qquad
U(t)\in\mathcal{U}(\mathscr{H})
\end{equation}
where $\psi_0$ is some initial condition, whose normalization is ordinarily chosen such that $\|\psi_0\|^2=1$. Then, $\psi_0\in S(\mathscr{H})$ implies $\psi(t)\in S(\mathscr{H})$ at all times.
The relation \eqref{psievol} takes the Lagrangian $L(\psi,\dot\psi)$ to a Lagrangian of the type $L_{\psi_0}(U,\dot{U})$, which then produces Euler-Lagrange equations for the Lagrangian coordinate $U\in\mathcal{U}(\mathscr{H})$. Moreover, by following Euler-Poincar\'e theory \cite{holm1998euler}, one denotes by $\mathfrak{u}(\mathscr{H})$ the Lie algebra of skew Hermitian operators and defines
\[
\xi(t):=\dot{U}(t)U^{-1}(t)\in\mathfrak{u}(\mathscr{H})\,.
\]
Since $\dot\psi=\xi\psi$, one obtains the reduced Lagrangian
\[
\ell: \mathfrak{u}(\mathscr{H})\times\mathscr{H}\to\Bbb{R}
\,,\qquad
\ell(\xi,\psi):=L(\psi,\xi\psi)
\]
and the Euler-Poincar\'e variational principle
\begin{equation}
\delta\int_{t_1}^{t_2\!}\!\ell(\xi,\psi)\,{\rm d} t=0
\,.
\label{varprinc1}
\end{equation}
Then, upon computing
\begin{equation} \label{variations}
\delta\xi=\dot\eta+[\eta,\xi]\,,\qquad
\delta\psi=\eta\psi
\end{equation}
where $\eta:=(\delta U)U^{-1}$, one obtains the following result.
\begin{theorem}
Consider the variational principle \eqref{varprinc1} with the auxiliary equation $\dot\psi=\xi\psi$ and the variations \eqref{variations}, where $\eta$ is arbitrary and vanishes at the endpoints. This variational principle is equivalent to the equations of motion
\begin{equation}
\frac{{\rm d}}{{\rm d} t}\frac{\delta \ell}{\delta\xi} - \left[ \xi,\frac{\delta \ell}{\delta \xi} \right] = \frac{1}{2} \left( \frac{\delta \ell}{\delta \psi}\psi^{\dagger}-\psi \frac{\delta \ell}{\delta \psi}^{\!\dagger} \right) \label{EQ:EPpsi}
,\qquad\qquad
\frac{{\rm d}\psi}{{\rm d} t} = \xi\psi
\,.
\end{equation}
\end{theorem}
Here, we use the ordinary definition of variational derivative
\[
\delta F(q) :=\left\langle\frac{\delta F}{\delta q},\delta q\right\rangle
,
\]
for any function(al) $F\in C^{\infty\!}(M)$ on the manifold $M$.
In typical situations, the reduced Lagrangian is quadratic in $\psi$, so that the $\mathcal{U}(1)$-invariance under phase transformations takes the dynamics to the projective space ${\mathbf{P}_{\!}\mathscr{H}}$. Indeed, as we shall see, the reduced Lagrangian $\ell(\xi,\psi)$ can be written typically in terms of the projection $\rho_\psi=\psi\psi^\dagger$ to produce a new Lagrangian
\[
l: \mathfrak{u}(\mathscr{H})\times{\mathbf{P}_{\!}\mathscr{H}}\to\Bbb{R}
\,,\qquad
l(\xi,\rho_\psi)=\ell(\xi,\psi)
\,.
\]
In this case, a direct calculation shows that
\begin{equation}\label{variations2}
\delta\rho_\psi=[\eta,\rho_\psi]\,,\qquad
\dot{\rho}_\psi=[\xi,\rho_\psi]
\end{equation}
and the previous theorem specializes as follows
\begin{theorem}
Consider the variational principle $
\delta\!\int_{t_1}^{t_2\!}l(\xi,\rho_\psi)\,{\rm d} t=0
$
with the relations \eqref{variations2} and $\delta\xi=\dot\eta+[\eta,\xi]$, where $\eta$ is arbitrary and vanishes at the endpoints. This variational principle is equivalent to the equations of motion
\begin{equation}
\frac{{\rm d}}{{\rm d} t}\frac{\delta l}{\delta\xi} - \left[ \xi,\frac{\delta l}{\delta \xi} \right] =\left[ \frac{\delta l}{\delta\rho_\psi},\rho_\psi \right] \label{EQ:EPrho}
,\qquad\quad
\dot{\rho}_\psi=[\xi,\rho_\psi]\,.
\end{equation}
\end{theorem}
Then, the unitary symmetry properties of the Lagrangian naturally take the evolution to the correct quantum state space (for pure states), that is the projective space ${\mathbf{P}_{\!}\mathscr{H}}$. In the following sections, we shall specialize this construction to two particular examples and we shall present the momentum map properties of the underlying geometry as well as their relation to the usual principal connections appearing in the literature.
\subsection{Dirac-Frenkel variational principle}
It is easy to see that upon following the construction from the previous section, the DF Lagrangian
\begin{equation}\label{DFLAGR}
L(\psi,\dot\psi)=\big\langle\psi,i\hbar\dot\psi-{H}\psi\big\rangle
\end{equation}
produces the Euler-Poincar\'e variational principle
\[
\delta\int_{t_1}^{t_2\!}\!\left\langle\psi,i\hbar\xi\psi-{H}\psi\right\rangle{\rm d} t=0
\]
For simplicity, here we are considering a time-independent Hamiltonian operator $H$. Then, upon computing
\begin{align*}
\frac{\delta l}{\delta\psi} = 2(i\hbar\xi-H) \psi,~~~~~~~~ \frac{\delta l}{\delta\xi} = -i\hbar \psi \psi^{\dagger} \,,
\end{align*}
the first of \eqref{EQ:EPpsi} yields
\begin{equation}
[\left(i\hbar\xi-H\right),\psi\psi^{\dagger}] = 0\,.
\label{EQ:SCHRU_EP1}
\end{equation}
Upon setting $\mathscr{H}=\Bbb{C}^n$ and making use of the anticommutator bracket $\{A,B\}=AB+BA$, the solution of the above equation can be written as
\begin{equation}
\xi+i\hbar^{-1\!}H=\{\boldsymbol{1}-2\psi\psi^{\dagger},\kappa\}
\label{EQ:projsch}
\end{equation}
for an arbitrary time-dependent skew Hermitian matrix $\kappa(t)$. The meaning of this solution will be clear in Section \ref{sec:momaps}. In the \text{Schr\"odinger}\, picture, the relation \eqref{EQ:projsch} recovers the usual phase arbitrariness, as it is shown by simply using the second in \eqref{EQ:EPpsi} to write
\begin{equation}
i\hbar\dot\psi=H\psi+\alpha\psi\,,
\label{EQ:projsch2}
\end{equation}
where $\alpha(t):=2\hbar\langle i\rho_\psi,\kappa\rangle$ (notice that we have chosen a unit initial vector so that $\|\psi_0\|^2=\|\psi\|^2=1$).
In the above equation, the term $\alpha\psi$ generates an arbitrary phase factor. Then, equation \eqref{EQ:projsch2} can be easily written in the form of a projective \text{Schr\"odinger}\, equation \cite{kibble1979geometrization}
\[
(1-\psi\psi^\dagger)(i\hbar\dot\psi-H\psi)=0\,,
\]
which can be recovered from the constrained DF Lagrangian \cite{Ohta2000}
\begin{equation}
L(\psi,\dot\psi,\lambda,\dot{\lambda})=\big\langle\psi,i\hbar\dot\psi-{H}\psi\big\rangle+\lambda(\|\psi\|^2-1)\,.
\label{EQ:constrLagr}
\end{equation}
As we shall see, the right hand side of \eqref{EQ:projsch} modifies the usual Heisenberg picture dynamics.
\subsection{Mixed states dynamics and its Wigner-Moyal formulation}
It is easy to see that all the phase terms in the previous section are consistently projected out by simply defining the Lagrangian $l(\xi,\rho_\psi)=\left\langle\rho_\psi,i\hbar\xi-{H}\right\rangle$ so that the first of \eqref{EQ:EPrho} reads $[(i\hbar\xi-H),\rho_\psi] = 0$ and the second recovers the quantum Liouville equation for pure states.
In the remainder of this section we shall generalise the previous approach to consider a new variational principle for \emph{mixed quantum states}. Let us consider the Lagrangian
\begin{equation}
l(\xi,\rho)=\left\langle\rho,i\hbar\xi-{H}\right\rangle
\label{mixedlagr}
\end{equation}
where $\xi=\dot{U}U^{-1}$ as before, while $\rho$ is a density matrix undergoing unitary evolution $\rho(t)=U\rho_0 U^\dagger$. In the case of mixed states, we have $\rho^2\neq\rho$ although the trace invariants ${\operatorname{Tr}}(\rho^n)$ are still preserved. We notice that a simple computation yields
\begin{equation}
\delta\rho=[\eta,\rho]\,,\qquad\dot{\rho}=[\xi,\rho]\,,
\label{mixedvariations}
\end{equation}
and therefore the application of Euler-Poincar\'e theory is straightforward. Then, one obtains precisely the same equations as in \eqref{EQ:EPrho} (upon replacing $\rho_\psi$ by $\rho$), which in turn give
\begin{equation}
[(i\hbar\xi-H),\rho] = 0\,,
\qquad
\dot\rho=[\xi,\rho]
\,.
\label{mixed}
\end{equation}
At this point the Liouville-Von Neumann equation
\[
i\hbar\dot{\rho} =\left[H,\rho\right]
\]
is obtained by direct substitution. Notice that the solution of the first equation in \eqref{mixed} differs from \eqref{EQ:projsch}, since $\rho^2\neq\rho$. For example, one has particular solutions of the form $i\hbar\xi-H=\sum_n\alpha_n\rho^n$. This reflects the very different geometric structures underlying mixed states and pure states. For a geometric description of mixed states in terms of coadjoint orbits and orthogonal frame bundles, we refer the reader to \cite{Montgomery91}. We emphasize that the action principle associated to the Lagrangian \eqref{mixedlagr} is very different from the one proposed in \cite{Heller76} and to our knowledge it has not appeared before.
Motivated by applications in chemical physics, we show how the above variational principle recovers the celebrated Wigner-Moyal picture of quantum dynamics on phase space. This formulation \cite{moyal1949quantum,wigner1932quantum} is based on the Weyl correspondence between linear operators and phase space functions (see e.g. \cite{Zachos2005}). For simplicity, this section presents the Euler-Poincar\'e formulation on the two-dimensional phase space (one spatial dimension), however this can be easily generalized to higher dimensions.
Consider an arbitrary linear operator $A\in L(\mathscr{H})$: the corresponding phase-space function is given by the Wigner transform $a(x,p)=\mathcal{W}(A)$ and the latter can be inverted by using the Weyl transform, $A=\mathcal{W}^{-1}(a)$. More explicitly, one has
\begin{align*}
\mathcal{W}(A) &\!:= \frac{1}{\pi\hbar} \int\! {\rm d} x^{\prime} \langle x+x^{\prime}|A |x-x^{\prime}\rangle e^{-\frac{2ipx^{\prime}}{\hbar}} ,\label{EQ:WF}
\\
\mathcal{W}^{-1}(a) &= 2\int \!{\rm d} x{\rm d} x^{\prime}\,|x+x^{\prime}\rangle\langle x-x^{\prime}|\int\! {\rm d} p ~ a(x,p) e^{\frac{2ipx^{\prime}}{\hbar}}.
\end{align*}
\rem{
Conversely, given a phase-space distribution, the Hilbert space operator is recovered via the Weyl transform given by
\begin{equation}
A = 2\int {\rm d} x{\rm d} x^{\prime}\int {\rm d} p ~|x+x^{\prime}\rangle a(x,p) e^{2ipx^{\prime}/\hbar} \langle x-x^{\prime}| \,.\label{EQ:WeylTR}
\end{equation}
For notation purposes, one can write equations (\ref{EQ:WF}) and (\ref{EQ:WeylTR}) as
\begin{equation}
a(x,p) = \mathcal{W}(A)
\,, \quad\quad\quad A = \mathcal{W}^{-1}(a)
\,.
\end{equation}
}
\rem
By introducing the star product \cite{Gro46}\cite{Moyal}, one has the phase-space correspondent of the product of two Hilbert space operators, that is (see equation (112) in \cite{Zachos})
\[
a \star b =
\mathcal{W}\left( AB \right)\,.
\]
It is known that the product of Hilbert space operators can be expressed as the Weyl transform of the $\star$-product of the corresponding Wigner distributions via Groenewold's relation \cite{Zachos2005}
\begin{equation*}
AB = \mathcal{W}^{-1}( a\star b )\,,
\end{equation*}
where
\begin{equation*}
(a\star b)(x,p) = \frac{1}{\hbar^2\pi^2} \int dp^{\prime}dp^{\prime\prime}dx^{\prime}dx^{\prime\prime}~a(x^{\prime},p^{\prime}) b(x^{\prime\prime},p^{\prime\prime}) e^{-\frac{2i}{\hbar}\left[ p(x^{\prime}-x^{\prime\prime}) + p^{\prime}(x^{\prime\prime}-x) + p^{\prime\prime}(x-x^{\prime}) \right]}\,,
\label{EQ:star}
\end{equation*}
}
Then, the Moyal bracket (see \cite{moyal1949quantum,Zachos2005} for its explicit definition) is defined in such a way that the commutator between two quantum operators is taken into the Moyal bracket of the corresponding phase space functions, that is \cite{Zachos2005}
\[
\lbrace\!\lbrace a,b \rbrace\!\rbrace = {\frac{1}{i\hbar} \mathcal{W}\left( [A,B]\right)}
\,.
\]
At this point, one can express the Lagrangian \eqref{mixedlagr} in terms of phase space functions. Indeed, upon defining the Wigner distribution $W(x,p)=\mathcal{W}(\rho)$ and by replacing the inverse relation $\rho=\mathcal{W}^{-1}(W)$ in \eqref{mixedlagr} one obtains the equivalent variational principle on phase space
\begin{equation}\label{WMvarprinc}
\delta\int_{t_1}^{t_2}\!\! \iint {\rm d} x {\rm d} p ~W(x,p) \Big( \hbar \Upsilon(x,p) - H(x,p) \Big)\, {\rm d} t=0\,,
\end{equation}
where we have defined $H(x,p)=\mathcal{W}(H)$ and $\Upsilon(x,p):=\mathcal{W}(i\xi)$.
\rem
is defined by
\[
\varUpsilon(x,p) := \frac{1}{\pi\hbar} \int dx^{\prime}~ \langle x+x^{\prime}| i\xi |x-x^{\prime} \rangle e^{-2ipx^{\prime}/\hbar}
\]
}
Then, upon recalling the relations \eqref{mixedvariations} and the first in \eqref{variations},
one computes
\[
{\delta\Upsilon =\frac{\partial \Theta}{\partial t} + \hbar \lbrace\!\lbrace \Theta, \Upsilon \rbrace\!\rbrace \,, \quad\quad \delta W = \hbar\lbrace\!\lbrace \Theta, W \rbrace\!\rbrace}\,, \quad\quad \partial_t W = \hbar\lbrace\!\lbrace \Upsilon, W \rbrace\!\rbrace \,,
\]
where $\Theta :=\mathcal{W}(i\eta)$, and the Euler-Poincar\'e variational principle \eqref{WMvarprinc} gives
\begin{equation*}
\frac{\partial W(x,p,t)}{\partial t} = \big\lbrace\!\big\lbrace H(x,p), W(x,p,t) \big\rbrace\!\big\rbrace
\,.
\end{equation*}
Again, we notice that the variational principle \eqref{WMvarprinc} has never appeared before in the literature and it is very different from other approaches proposed earlier, such as \cite{Poulsen11}. In particular, the variational principle \eqref{WMvarprinc} is entirely derived from the Dirac-Frenkel Lagrangian and no assumption has been made other than unitary evolution.
\subsection{Geodesics on the space of quantum states\label{sec:FS}}
Fubini-Study geodesics on $\boldsymbol{P}\Bbb{C}^n$ are used in various situations of quantum mechanics. Their applications in quantum search algorithms \cite{MiWa01,Zhao2012}, time-optimal control problems \cite{CarHoKoOku06,CarHoKoOku07,Dalessandro01} and holonomic quantum computation \cite{GunWaNa2014,Lucarelli2005,TaNaHa2005} emphasizes their importance and makes this example especially interesting. In the general case, geodesics are optimal curves in the sense that they minimize the distance between two quantum states.
The Fubini-Study geodesics are defined as geodesic equations on ${\mathbf{P}_{\!}\mathscr{H}}$ minimizing the the Fubini-Study distance. These geodesic flows can be written explicitly as Euler-Lagrange equations associated to the action principle $\delta\int_{t_1}^{t_2}{\rm d} t\,L(\psi,\dot\psi)=0$ with Lagrangian $L:T\!\mathscr{H}\to\Bbb{R}$ given by (see e.g. \cite{FaKuMaMa10})
\begin{equation}\label{FSLAGR}
L(\psi,\dot\psi)=\frac{\hbar}{2}\frac{\|\psi\|^2\|\dot\psi\|^2-|\langle\dot\psi|\psi\rangle|^2}{\|\psi\|^4}
\,.
\end{equation}
More explicitly, lengthy computations yield
\[
\big(\|\psi\|^2-\psi\psi^\dagger\big)\big(\|\psi\|^2\ddot\psi-2\langle\psi|\dot\psi\rangle\dot\psi\big)=0
\,.
\]
Notice that this approach does not involve normalized vectors in $\mathscr{H}$. However, a simple way to recover normalization is to use a constrained Lagrangian such as that in \eqref{EQ:constrLagr}.
Application of the Euler-Poincar\'e theory to Fubini-Study geodesics can be performed again by following the procedure outlined in Section \ref{SEC:EPSchr}, without modifications. Then, upon recalling that $\xi^2$ is Hermitian, one obtains the reduced Lagrangian (set $\hbar=1$, for convenience)
\begin{align}
l(\xi,\rho_\psi)
=&-\frac12\left(\left<\rho_\psi,\xi^2\right>+\big<\rho_\psi,i\xi\big>^2\right)
.
\label{FSEPLAGR}
\end{align}
Notice that we recover the well known relation $l(\xi,\rho_\psi)=1/2\,\big(\langle\mathcal{E}^2\rangle- \langle\mathcal{E}\rangle^2\big)$ \cite{AhAn87},
where we have introduced the energy operator $\mathcal{E}=i\xi$ and we have used the standard expectation value notation.
Then, upon replacing the variational derivatives
\begin{equation}\label{varderiv}
\frac{\delta l}{\delta\rho_\psi}=
-\frac12\,\xi^2+\big<\rho_\psi|\xi\big>\xi
\,,\qquad\quad
\frac{\delta l}{\delta\xi}=
\frac12\{\rho_\psi,\xi\}-\big<\rho_\psi|\xi\big>\rho_\psi
\end{equation}
in equations \eqref{EQ:EPrho}, standard matrix computations give
\begin{equation*}
\frac{{\rm d}}{{\rm d} t}\Big(\{\rho_\psi,\xi\}-2\big<\rho_\psi|\xi\big>\rho_\psi\Big)=0
\,,\qquad\quad
\dot\rho_\psi=[\xi,\rho_\psi]
\,,
\end{equation*}
where the first emphasizes the following conservation form of the Fubini-Study geodesic equation
\[
\frac{{\rm d}}{{\rm d} t}\Big((\boldsymbol{1}-2\rho_\psi)\dot{\rho}_\psi\Big)=
\frac{{\rm d}}{{\rm d} t}\!\left(\dot\psi\psi^\dagger-\psi\dot\psi^\dagger-2\langle\psi|\dot\psi\rangle\psi\psi^\dagger\right)=0\,,
\]
as it arises from the left-invariance of the Lagrangian \eqref{FSLAGR} (see e.g. \cite{andersson2013dynamic}). Notice that applying the above conservation law to $\psi$ and writing $\langle\psi|\dot\psi\rangle=i\langle i\psi,\dot\psi\rangle$ yields
\[
\ddot\psi-\big<\ddot\psi|\psi\big>\psi
=
2i\big<i\psi,\ddot\psi\big>\psi+2\big<\psi|\dot\psi\big>\big(\dot\psi+\big<\dot\psi|\psi\big>\psi\big)
\]
and since $\big<\dot\psi|\psi\big>=-\big<\psi|\dot\psi\big>$, expanding $\big<i\psi,\ddot\psi\big>$ leads to
\[
\big(1-\psi\psi^\dagger\big)\big(\ddot\psi-2\langle\psi|\dot\psi\rangle\dot\psi\big)=0
\,.
\]
This geodesic flow can be also recovered as an Euler-Lagrange equation by adding a normalisation constraint $\lambda(\|\psi\|^2-1)$ to the Lagrangian \eqref{FSLAGR} (cf. the constrained DF Lagrangian \eqref{EQ:constrLagr}).
\rem{
\begin{remark}[The isoholonomic problem]
Other variants of the Lagrangian \eqref{FSLAGR} are also studied. For example, in holonomic quantum computation \cite{}, one inserts the extra constraint $r\langle i\psi,\dot\psi\rangle$ to enforce the condition that the connection form $\langle i\psi,{\rm d}\psi\rangle$ vanishes on the Hopf bundle (horizontal lift condition). The so-called \emph{isoholonomic problem} \cite{} is then written as
\[
\delta\int_{t_1}^{t_2\!}\!\left(\frac{\hbar}{2}\frac{\|\psi\|^2\|\dot\psi\|^2-|\langle\dot\psi|\psi\rangle|^2}{\|\psi\|^4}+\lambda(\|\psi\|^2-1)+r\langle i\psi,\dot\psi\rangle\right){\rm d} t=0
\]
The Euler-Poincar\'e approach applies also in this case without modification, thereby leading to
\begin{equation*}
\frac{{\rm d}}{{\rm d} t}\Big(\{\rho_\psi,\xi\}+2ir\rho_\psi\Big)=0
\,,\qquad\quad
\dot\rho_\psi=[\xi,\rho_\psi]
\,.
\end{equation*}
Here, the Lagrange multiplier $r(t)$ has to be determined by using the condition $\langle\rho_\psi|\xi\rangle=0$. For example, one may take the trace of the first equation to conclude $\dot{r}=0$. Then, combining the two equations above leads to
\[
\frac{{\rm d}}{{\rm d} t}\Big((\boldsymbol{1}-2\rho_\psi)\dot{\rho}_\psi+2ir\rho_\psi\Big)=0
\,,\qquad\quad
\langle\rho_\psi|\dot{\rho}_\psi\rangle=0
\,.
\]
\end{remark}
}
Geodesic flows on the quantum state space have always raised questions concerning their underlying geometric properties \cite{Anandan91,uhlmann2009geometry}. For example, in holonomic quantum computing, a fundamental role is played by the connection form $\langle\psi|\dot\psi\rangle$, whose loop integral defines the celebrated geometric phase. In addition, connection forms also allow the usual horizontal/vertical decomposition on the Hopf bundle. In geometric mechanics, this decomposition can be performed by using a more sophisticated theory than Euler-Poincar\'e reduction. This is called \emph{Lagrange-Poincar\'e reduction} \cite{cendra2001lagrangian} and it was recently formulated in the context of homogeneous spaces (arising from symmetry breaking) in \cite{gay2010reduction}. Without entering the technicalities of Lagrange-Poincar\'e reduction, we shall only mention that this theory often takes advantage of a particular connection form that appears to have a precise physical meaning in many different cases: this is called the \emph{mechanical connection} and it is defined in terms of a \emph{momentum map}, another fundamental object in geometric mechanics. This is the topic of the next section.
\subsection{Momentum maps of quantum variational principles\label{sec:momaps}}
The first momentum map one encounters in quantum mechanics is probably the density matrix for pure states \cite{CleMar08}. More precisely, the action of the unitary group $\mathcal{U}(\mathscr{H})$ on the quantum Hilbert space $\mathscr{H}$ (endowed with the symplectic form $\Omega(\psi_1,\psi_2)=2\hbar\left<i\psi_1,\psi_2\right>$) produces the momentum map
\[
J(\psi)=-i\hbar\psi\psi^\dagger\in\mathfrak{u}(\mathscr{H})^*
\,,
\]
as it can be easy obtained by the general formula $\langle J(\psi),\xi\rangle=1/2\,\Omega(\xi\psi,\psi)$ \cite{marsden1999introduction,holm2009geometric}, holding for an arbitrary skew-hermitian operator $\xi\in\mathfrak{u}(\mathscr{H})$. Also, restricting to consider phase transformations yields the total probability or, more precisely, the quantity $J(\psi)=\hbar\|\psi\|^2$.
Other than those above, other momentum map structures appear in geometric quantum dynamics and each correspond to different group actions and different reduction processes. It turns out that in quantum variational principles, the most important momentum map is associated to the action of the isotropy subgroup of the initial state. In order to explain this statement, let us replace the relation \eqref{psievol} in a Lagrangian of the type \eqref{genlagr} and observe that this produces a Lagrangian $\mathcal{L}:T\mathcal{U}(\mathscr{H})\to\Bbb{R}$ by
\begin{equation}\label{callagr-def}
\mathcal{L}(U,\dot{U}):=L(U\psi_0,\dot{U}\psi_0)
\,.
\end{equation}
Although, this Lagrangian is not symmetric under right multiplication, i.e.
\[
\mathcal{L}(U,\dot{U})\neq\mathcal{L}(UU',\dot{U}U')\,,\qquad\quad U'\in\mathcal{U}(\mathscr{H}),
\]
the invariance property is recovered by restricting to the isotropy group of $\psi_0$, that is
\[
\mathcal{U}_{\psi_0}(\mathscr{H})=\{U\in\mathcal{U}(\mathscr{H})\,|\,U\psi_0=\psi_0\}\,.
\]
Indeed, one evidently has
\begin{equation}\label{callagr}
\mathcal{L}(U,\dot{U})=\mathcal{L}(UU_0,\dot{U}U_0)\,,\qquad\quad \forall\,U_0\in\mathcal{U}_{\psi_0}(\mathscr{H}),
\end{equation}
and one may choose the initial vector $\psi_0$ to coincide with the basis vector $\psi_0=(0\,\dots\,0\,1)^\dagger$.
\begin{remark}[Analogy with the heavy top dynamics]
We observe that the argument above holds in a wide range of situations, including, for example, the Lagrangian reduction for the heavy top dynamics \cite{gay2010reduction}. In that context, the unitary group is replaced by the rotation group $SO(3)$ and the isotropy symmetry is defined to preserve the gravity vector, thereby leading to planar rotations in $SO(2)$. The Noether's conserved quantity (i.e. the momentum map) is then the vertical angular momentum.
\end{remark}
At this point, it is natural to ask what the momentum map is for the reduction of quantum variational principles. More particularly, we look for the momentum map associated to (the cotangent lift of) the right action of $\mathcal{U}_{\psi_0}(\mathscr{H})$ on the cotangent bundle $T^*\mathcal{U}(\mathscr{H})$. In the general case, it has recently been shown \cite{gay2010reduction} that the momentum map for the right action of a subgroup $G_0\subset G$ on (the trivialisation of) the cotangent bundle $T^*G\simeq G\times\mathfrak{g}^*$ reads
\begin{equation}\label{genmomap}
J\!\left(g,\mu\right)=\iota^*\!\left(\operatorname{Ad}^*_g\mu\right)
\end{equation}
where $(g,\mu)\in G\times\mathfrak{g}^*$, $\operatorname{Ad}^*_g\mu=g^\dagger\mu g^{-\dagger}$ is the standard matrix coadjoint representation and $\iota^*$ is the dual of the Lie algebra inclusion $\iota:\mathfrak{g}_0\hookrightarrow\mathfrak{g}$. To simplify the treatment, set $\mathscr{H}=\Bbb{C}^n$ and choose $\psi_0=(0\,\dots\,0\,1)^\dagger$ without loss of generality. Then, $\mathcal{U}_{\psi_0}(\mathscr{H})=\mathcal{U}(n-1)\subset\mathcal{U}(n)=\mathcal{U}(\mathscr{H})$ and the group inclusion $\mathcal{U}(n-1)\hookrightarrow\mathcal{U}(n)$ is
\begin{equation}\label{inclusion}
\mathcal{U}(n-1)\ni{U}_0\mapsto\left(\begin{array}{cc}
{U} _0& 0\\
0&1
\end{array}\right)\in \mathcal{U}(n)
\,.
\end{equation}
The corresponding Lie algebra inclusion $\iota:\mathfrak{u}(n-1)\hookrightarrow\mathfrak{u}(n)$ reads
\[
\iota({\xi}_0)=\left(\begin{array}{cc}
{\xi}_0 & 0\\
0&0
\end{array}\right),
\]
while its dual $\iota^*$ is given by $\iota^*(\mu)=(\boldsymbol{1}-\rho_{\psi_0})\mu(\boldsymbol{1}-\rho_{\psi_0})$, that is the standard projection on the upper left block. This result is independent of the number of dimensions and it leads to the following momentum map:
\begin{align}\nonumber
J_1(U,\mu)=&\,\frac12\big\{(\boldsymbol{1}-2\rho_{\psi_0}),\operatorname{Ad}^*_U\mu\big\}+\langle\rho_{\psi_0}|\operatorname{Ad}^*_U\mu\rangle\rho_{\psi_0}
\\
&\,
=\operatorname{Ad}^*_U\!\left(\frac{\delta \ell}{\delta\xi}-\bigg\{\!\rho_{\psi},\frac{\delta \ell}{\delta\xi}\bigg\}+\bigg\langle\rho_{\psi}\bigg|\frac{\delta \ell}{\delta\xi}\bigg\rangle\rho_{\psi}\!\right)
\label{momap1}
\,,
\end{align}
where we have simply rewritten \eqref{genmomap} by replacing the formula for $\iota^*$. Here, we recall $i\rho_{\psi_0}=\operatorname{Ad}^*_{U}(i\rho_{\psi})$, from the definition $\rho_{\psi}:=\psi\psi^\dagger=U\rho_{\psi_0} U^{-1}$. Therefore, because of the symmetry property \eqref{callagr} possessed by any Lagrangian of the type \eqref{genlagr}, the corresponding Euler-Poincar\'e equations \eqref{EQ:EPpsi} conserve $J_{1\!}\!\left(U,{\delta \ell}/{\delta\xi}\right)$. More particularly, one shows that any quantum system with an arbitrary Lagrangian of the type \eqref{callagr-def} produces dynamics on the zero-level set of $J_{1\!}$. This is easily shown by using the relation $L(\psi,\dot{\psi}) = L(\psi,\xi\psi) =: \ell (\xi,\psi)$, so that
\begin{equation}\label{mark}
\frac{\delta\ell}{\delta\xi} = \frac{1}{2}\left( \frac{\delta L}{\delta\dot{\psi}}\psi^{\dagger} - \psi\frac{\delta L}{\delta\dot{\psi}}^{\dagger} \right)\,,
\end{equation}
thereby verifying $J_{1\!}\!\left(U,{\delta \ell}/{\delta\xi}\right)\equiv0$.
Now that we have characterised the momentum map associated to the action of $\mathcal{U}(n-1)$, we recall that all physically relevant Lagrangians must be also phase invariant, so that they can be eventually written in terms of the projection $\rho_\psi=\psi\psi^\dagger\in\mathbf{P}\!\mathscr{H}$. Therefore, the most general symmetry group of the Lagrangian \eqref{callagr-def} has to include phase transformations and this leads us to consider the direct product $\mathcal{U}(n-1)\times\mathcal{U}(1)$. The latter can be embedded in $\mathcal{U}(n)$ by the inclusion
\begin{equation}\label{inclusion2}
\mathcal{U}(n-1)\times\mathcal{U}(1)\ni({U}_0,\varphi)\mapsto\left(\begin{array}{cc}
{U} _0& 0\\
0&e^{-i\varphi}
\end{array}\right)\in \mathcal{U}(n)
\,,
\end{equation}
where the minus sign in the exponent is of purely conventional nature.
Since the momentum map associated to $\mathcal{U}(n-1)$ has already been presented in \eqref{momap1}, we need to compute only the momentum map associated to the group $\mathcal{U}(1)\subset\mathcal{U}(n)$, endowed with the group inclusion
\begin{equation}\label{inclusion3}
\mathcal{U}(1)\ni\varphi\mapsto\left(\begin{array}{cc}
\boldsymbol{1}& 0\\
0&e^{-i\varphi}
\end{array}\right)\in \mathcal{U}(n)
\,.
\end{equation}
Upon computing the dual of the corresponding Lie algebra inclusion $\iota(\alpha)=-i\alpha\rho_{\psi_0}\in\mathfrak{u}(n)$ and by identifying $\mathfrak{u}(1)\simeq\Bbb{R}$, one has the momentum map formula
\begin{equation}\label{momap2}
J_2(U,\mu)=i\left\langle\rho_{\psi_0}\,|\,\operatorname{Ad}^*_U\mu\right\rangle
\,.
\end{equation}
Any quantum system with an arbitrary Lagrangian of the type \eqref{callagr-def} takes the above momentum map to the form
\[
J_2(U,\delta\ell/\delta\xi)=\left\langle\psi,i\,\frac{\delta L}{\delta\dot\psi}\right\rangle
=:\mathcal{J}\!\left(\!\psi,\frac{\delta L}{\delta\dot\psi}\right)
,
\]
as it is easily shown by using the relation \eqref{mark}. Here, the momentum map $\mathcal{J}:T^*S(\Bbb{C}^n)\to\Bbb{R}$ arises from the action of $\mathcal{U}(1)$ (notice that we idenitified $\mathfrak{u}(1)\simeq\Bbb{R}$) on the cotangent bundle $T^{*\!}S(\Bbb{C}^n)$. For example, the DF Lagrangian yields $\mathcal{J}(\psi,\delta L/\delta\dot\psi)=\hbar\|\psi\|^2$, while one verifies that the FS Lagrangian \eqref{FSLAGR} makes $\mathcal{J}$ vanish identically.
In more generality, the momentum map corresponding to the action of the full symmetry group $\mathcal{U}(n-1)\times\mathcal{U}(1)$ is given by
\begin{equation}\label{totalmomap}
J(U,\mu)=\big(J_1(U,\mu), J_2(U,\mu)\big)=\left(\frac12\big\{(\boldsymbol{1}-2\rho_{\psi_0}),\operatorname{Ad}^*_U\mu\big\}+\langle\rho_{\psi_0}|\operatorname{Ad}^*_U\mu\rangle\rho_{\psi_0},\,i\left\langle\rho_{\psi_0}\,|\,\operatorname{Ad}^*_U\mu\right\rangle\right)
\end{equation}
Therefore, because of the symmetry property \eqref{callagr} possessed by any phase-invariant Lagrangian of the type \eqref{genlagr}, the corresponding Euler-Poincar\'e quantum dynamics \eqref{EQ:EPrho} conserves the momentum map $J_2$ and lies on the kernel of $J_1$.
\begin{remark}[Phases and $\mathcal{U}(1)-$actions] Notice that this section has used a $\mathcal{U}(1)-$action that is different from usual phase transformations. Indeed, while the latter act on vectors by the diagonal action $\psi\mapsto e^{-i\vartheta}\psi$, the $\mathcal{U}(1)-$action used in this section reads $\psi\mapsto (\boldsymbol{1}-\rho_{\psi_0}+e^{-i\varphi}\rho_{\psi_0})\psi$. However, usual phase transformations (denoted by $\mathcal{U}_d(1)$ to emphasise the diagonal action) are a subgroup of $\mathcal{U}(n-1)\times\mathcal{U}(1)$, as it is given by the inclusion
\[
\mathcal{U}_d(1)\ni(\vartheta)\mapsto\left(\begin{array}{cc}
e^{-i\vartheta}\boldsymbol{1}& 0\\
0&e^{-i\vartheta}
\end{array}\right)\in \mathcal{U}(n-1)\times\mathcal{U}(1)\subset\mathcal{U}(n)
\,.
\]
Therefore, our treatment naturally includes the ordinary phase transformations, whose corresponding momentum map is given by $\iota_d^*(J_1(U,\mu),J_2(U,\mu))$. Here, $\iota_d^*$ is the dual of the inclusion $\iota_d:{\mathfrak{u}_d(1)\hookrightarrow\mathfrak{u}(n-1)\times\mathfrak{u}(1)}$ given by $\iota_d(\alpha)=({-i\alpha}\boldsymbol{1},{\alpha})$ (again, we identify $\mathfrak{u}_d(1)\simeq\Bbb{R}$). Upon using the pairing $\langle(\mu,\omega),(\eta,\alpha)\rangle=\langle\mu,\eta\rangle+\omega\alpha$, a direct calculation shows {$\iota_d^*(\mu,\omega)=\omega+\operatorname{Tr}(i\mu)$}, so that $\iota_d^*\big(J_1(U,\mu),J_2(U,\mu)\big)=i\operatorname{Tr}(\mu)$.
\end{remark}
The momentum maps provided in this Section are of paramount importance in geometric quantum dynamics, as they incorporate essential geometric properties. For example, it is interesting to notice that the momentum maps \eqref{momap1} and \eqref{totalmomap} can be used to define connection forms, respectively on the bundles $\mathcal{U}(n)\to S(\Bbb{C}^n)$ and $\mathcal{U}(n)\to \mathbf{P}\Bbb{C}^n$, as they are given in \eqref{bundles}. These connection forms are obtained by applying Lagrange-Poincar\'e reduction for symmetry breaking and we refer the reader to \cite{gay2010reduction} for more details on this topic. In the particular case under consideration, one identifies $\mathfrak{u}(k)^*\simeq\mathfrak{u}(k)$ so that the dual of the inclusion $\iota^*:{\mathfrak{u}(n)^*\to\mathfrak{u}(n-1)^*}$ determines a projection $\Bbb{P}:\mathfrak{u}(n)\to\mathfrak{u}(n-1)$, which in turn can be used to define the \emph{mechanical connection} $\mathcal{A}(\dot{U})={\Bbb{P}(\operatorname{Ad}_{U^{-1}}\xi)}={(\boldsymbol{1}-\rho_{\psi_0})}{(\operatorname{Ad}_{U^{-1}}\xi)}{(\boldsymbol{1}-\rho_{\psi_0})}$ on the bundle $\mathcal{U}(n)\to S(\Bbb{C}^n)$. An analogous construction yields a connection on the bundle $\mathcal{U}(n)\to \mathbf{P}\Bbb{C}^n$, given by $\mathcal{A}(\dot{U})={(\boldsymbol{1}-\rho_{\psi_0})}{(\operatorname{Ad}_{U^{-1}}\xi)}{(\boldsymbol{1}-\rho_{\psi_0})}+i\langle{\operatorname{Ad}_{U^{-1}}\xi\,|\,\rho_{\psi_0}\rangle\rho_{\psi_0}}$.
Also, it is well known that $\mathcal{A}(\dot\psi)=\langle\psi, i\dot\psi\rangle$ is a principal connection on the Hopf bundle $S(\Bbb{C}^n)\to\mathbf{P}\Bbb{C}^n$.
Then, the momentum maps presented in this section generate a connection form on each bundle of the diagram below. The study of these connection forms and their curvatures is left for future work.
\begin{diagram}
& & \mathcal{U}(n) & &
\\
& \ldTo & & \rdTo &
\\
S(\Bbb{C}^n) & & \rTo & & \mathbf{P}\Bbb{C}^n
\\
\end{diagram}
The arguments in this Section clarify the meaning of equation \eqref{EQ:projsch}. Indeed, the latter can be interpreted as simply saying that the infinitesimal generator $\xi\in\mathfrak{u}(n)$ (or, equivalently, the Hamiltonian operator $H$) is defined only up to an element of the isotropy subalgebra $\mathfrak{u}(n-1)\times\mathfrak{u}(1)$. This can be made explicit upon introducing $({\kappa}_0,\alpha_0)\in\mathfrak{u}(n-1)\times\mathfrak{u}(1)$ such that $\{\boldsymbol{1}-2\rho_\psi,\kappa\}={(\boldsymbol{1}-\rho_{\psi})}{\kappa_0}{(\boldsymbol{1}-\rho_{\psi})}+i\alpha_0\rho_\psi$. Although computing $\dot\psi=\xi\psi$ returns the usual phase arbitrariness, as shown in equation \eqref{EQ:projsch2}, the relation \eqref{EQ:projsch} discloses a rich geometry content underlying quantum evolution. Indeed, it reminds that the propagator of quantum dynamics is defined up to elements of $\mathcal{U}(n-1)\times\mathcal{U}(1)$ (not only up to phases in $\mathcal{U}(1)$), which is actually a non Abelian symmetry group (unlike phase transformations). As we shall see in the next Section, this argument accounts for propagators that depend on the initial quantum state $\psi_0$.
\section{The Heisenberg and Dirac pictures}
While the previous sections mainly dealt with the Schr\"odinger picture of quantum mechanics, the Heisenberg picture is rather unexplored in the geometry of quantum evolution. For example, one is interested in the role of projection operators, as they emerge from the projective geometry of the quantum state space in the Schr\"odinger picture. We shall consider the Heisenberg picture for the DF Lagrangian and the Fubini-Study geodesics.
\subsection{Euler-Poincar\'e reduction in the Heisenberg picture\label{sec:HeisDF}}
Most of this Section is devoted to the Heisenberg picture for the DF Lagrangian. It is easy to see that \eqref{DFLAGR} can be written in the Heisenberg picture by introducing
\[
\xi_H:=U^{-1}\dot{U}=\operatorname{Ad}_{U^{-1}} \xi
\,,\qquad\
H_H:=U^\dagger H U\,.
\]
Indeed, with these definitions, the Lagrangian \eqref{DFLAGR} becomes
\begin{equation}\label{HeisDF}
l(\xi_H,H_H)=\big\langle\rho_{\psi_0}, i\hbar\xi_H-{H}_H\big\rangle\,.
\end{equation}
Then, upon computing
\begin{equation} \label{variations3}
\delta\xi_H=\dot\eta_H-[\eta_H,\xi_H]\,,\qquad
\delta H_H=[H_H,\eta_H]
\,, \qquad\quad
\dot{H}_H=[H_H,\xi_H]
\end{equation}
(with $\eta_H:=U^{-1}\delta U$), inserting the Lagrangian \eqref{HeisDF} in the variational principle $\int_{t_1}^{t_2} \!l(\xi_H,H_H)\,{\rm d} t=0$ yields the following Euler-Poincar\'e equations:
\begin{equation}\label{generator}
\left[i\hbar\xi_H- H_H,\rho_{\psi_0}\right]=0
\,.
\end{equation}
At this point, we observe that although the above relation is satisfied by $i\hbar\xi_H- H_H=\alpha\boldsymbol{1}$ (so that the Heisenberg Hamiltonian $H_H$ is defined up to a phase factor $\alpha$), more general solutions are present such as
\begin{equation}\label{generator2}
\xi_H=-i\hbar^{-1\!}H_H+\{\boldsymbol{1}-2\rho_{\psi_0},\kappa\}
\end{equation}
($\kappa$ being arbitrary and skew Hermitian, see equation \eqref{EQ:projsch}). These solutions have the property of depending on the initial state $\psi_0$. Upon setting $\mathscr{H}=\Bbb{C}^n$ for simplicity, one may choose $\psi_0=(0\,\dots 0\,1)^\dagger$ without loss of generality. Interestingly enough, these more general solutions lead to the unfamiliar equation
\begin{equation}\label{EQ:projsch3}
\dot{H}_H=\left[H_H,\{\boldsymbol{1}-2\rho_{\psi_0},\kappa\}\right],
\end{equation}
so that the Hamiltonian operator $H_H$ is \emph{not} conserved in the general case. Although this may seem surprising, we observe that the above dynamics does not change the physics of the system under consideration. For example, we observe that the total energy is preserved:
\[
\langle\dot{H}_H\rangle=\langle\rho_{\psi_0}| \dot{H}_H\rangle=0\,,
\]
as shown by a direct verification. Moreover, one realizes that the above dynamics of quantum Hamiltonians returns exactly the \text{Schr\"odinger}\, equation \eqref{EQ:projsch2}: indeed, one has
\[
\xi_H\psi_0=-i\hbar^{-1\!}H_H\psi_0-2\langle\rho_{\psi_0}|\kappa\rangle\psi_0
\,
\]
so that, recalling $\psi=U\psi_0$ and applying $U$ on both sides returns \eqref{EQ:projsch2}. Notice that it is indeed essential that $\rho_{\psi_0}$ identifies the initial quantum state. We conclude that the physical content is unaltered by the {Heisenberg equation} \eqref{EQ:projsch3}, which in turn generalizes the standard Heisenberg dynamics (recovered by $\kappa=0$) to incorporate the geometry of quantum dynamics.
As a practical example, we consider spin dynamics in the Heisenberg picture. In this case, the Hamiltonian reads $H_H=\mathbf{n}\cdot{\boldsymbol{S}}_H$, where ${\boldsymbol{S}}_H(t)=U(t)^{-1}\boldsymbol{S} U(t)$ in standard spin operator notation. The DF Lagrangian \eqref{DFLAGR} is written in the Heisenberg picture as $l(\xi_H,{\boldsymbol{S}}_H)=\big\langle\rho_{\psi_0}, i\hbar\xi_H-\mathbf{n}\cdot{\boldsymbol{S}}_H\big\rangle$, so that the Euler-Poincar\'e equations
\begin{align*}
\left[i\hbar\xi_H- \mathbf{n}\cdot{\boldsymbol{S}}_H,\,\rho_{\psi_0}\right]=0
\,,
\qquad\quad
\dot{\boldsymbol\sigma}_H=[{\boldsymbol{S}}_H,\xi_H]
\end{align*}
specialize to yield
\[
\dot{\boldsymbol{S}}_H=i\hbar^{-1\!}\big[\mathbf{n}\cdot{\boldsymbol{S}}_H+i\{2\rho_{\psi_0}-\boldsymbol{1},\kappa\}, {\boldsymbol{S}}_H\big]
=\mathbf{n}\times{\boldsymbol{S}}_H-
\big[\{\boldsymbol{1}-2\rho_{\psi_0},\kappa\}, {\boldsymbol{S}}_H\big]
\]
Notice that this approach can be applied in the general case. For example, one can study linear oscillator dynamics by recalling the Hamiltonian $H_H=\hbar\omega a^\dagger_H a_H$ and following precisely the same steps as above. The present approach leads to the following Heisenberg equation:
\[
\dot{A}_H=i\hbar^{-1}\big[H_H,A_H\big]-\big[\{\boldsymbol{1}-2\rho_{\psi_0},\kappa\}, A_H\big],
\]
where $\kappa$ is an arbitrary skew-symmetric operator and $H_H$ undergoes its own evolution \eqref{EQ:projsch3}. In addition, equation \eqref{generator2} yields a new form of the propagator equation
\begin{equation}\label{propagator}
\dot{U}=i\hbar^{-1}HU+U\{\boldsymbol{1}-2\rho_{\psi_0} ,\kappa\}\,.
\end{equation}
It is necessary to point out that, since $\kappa$ is an arbitrary skew-Hermitian matrix parameter, one can simply choose it in such a way that $\{\boldsymbol{1}-2\rho_{\psi_0} ,\kappa\}=0$, thereby eliminating the dependence of the propagator on the initial conditions. A similar argument leads to eliminating phase terms in the Schr\"{o}dinger equation \eqref{EQ:projsch2} \cite{kibble1979geometrization}.
In all this Section, we assumed the initial state is a pure state ${\psi_0}$. If this is not the case, then different solutions of the type $\xi_H=-i\hbar^{-1\!}H_H+\kappa
$ (with $[\kappa,\rho_0]=0$) are allowed by equation \eqref{generator}, because in this case $\rho_{\psi_0}$ is replaced by a density matrix $\rho_0\neq\rho_0^2$.
The Heisenberg picture is particularly natural for the description of FS geodesics. Indeed, while in the \text{Schr\"odinger}\, picture the right unitary symmetry is broken by $\rho_\psi$, in the Heisenberg picture one may use the full left symmetry of the Lagrangian \eqref{FSLAGR}, that is $L(\psi,\dot\psi)=L(U^{-1}\psi,U^{-1}\dot\psi)$. Upon recalling \eqref{psievol} and by setting $\hbar=1$ for convenience, one obtains the Euler-Poincar\'e Lagrangian
\begin{equation}\label{HeisFS}
l(\xi_H)=-{\frac{1}{2}}\Big(\left<\rho_{\psi_0}|\xi_H^2\right>+\left<\rho_{\psi_0},i\xi_H\right>^2\Big)
\,.
\end{equation}
Then, upon using the first of \eqref{variations2}, the variational principle $\int_{t_1}^{t_2} l(\xi_H)\,{\rm d} t=0$ yields
\[
\{\dot{\xi}_H,\rho_{\psi_0}\}-2\big<\rho_{\psi_0}|\dot{\xi}_H\big>\rho_{\psi_0}+\Big[{\xi}_H^2-2\big<\rho_{\psi_0}|\xi_H\big>{\xi}_H,\rho_{\psi_0}\Big]=0\,,
\]
which reflects all the properties already discussed in Section \ref{sec:FS}
\subsection{Dirac-Frenkel Lagrangian in the Dirac picture}
This section extends the arguments from the previous Sections to formulate a new variational principle for quantum dynamics in the Dirac (interaction) picture. As we shall see, the Euler-Poincar\'e construction involves the semidirect product of the unitary group with itself.
In the Dirac picture, the \text{Schr\"odinger}\, Hamiltonian operator is split in two parts as $H = H_0 + H_1$, where $H_0$ is typically a simple linear Hamiltonian, while $H_1$ usually contains nonlinear potential terms. This replacement can then be inserted in the DF Lagrangian \eqref{DFLAGR}. However, it is convenient to keep track of the quantum state $\psi_s$ that is propagated by $H_0$, such that $i\hbar\dot\psi_s=H_0\psi_s$ (up to phase terms). Then, one is led to consider the following DF Lagrangian
\begin{equation}\label{DFLAGR2}
L(\psi,\dot\psi,\psi_s,\dot\psi_s)=\big\langle\psi,i\hbar\dot\psi-({H}_0+H_1)\psi\big\rangle+\big\langle\psi_s,i\hbar\dot\psi_s-{H}_0\psi_s\big\rangle
.
\end{equation}
Performing Euler-Poincar\'e reduction by replacing the evolution relation \eqref{psievol} on the first part yields the Lagrangian
\[
\bar{L}(\xi,\rho_\psi,\psi_s,\dot\psi_s)=\left\langle\rho_\psi,i\hbar\xi-{H}_0-H_1\right\rangle+\big\langle\psi_s,i\hbar\dot\psi_s-{H}_0\psi_s\big\rangle
\,,
\]
with $\rho_\psi=U\rho_{\psi_0}U^{-1}$ and $\xi:=\dot{U}U^{-1}$.
At this point, the propagator associated to $H_0$ can be used to replace the evolution relation $\psi_s(t)=U_0(t)\bar{\psi}_0$ in the second term, thereby leading to the Euler-Poincar\'e Lagrangian
\begin{equation}\label{DFLagrInt}
l(\xi_0,\xi_I,\rho_{\psi_I},H_{0,I},H_{1,I})=\left\langle\rho_{\psi_I},i\hbar\xi_I-H_{0,I}-H_{1,I}\right\rangle+\big\langle\rho_{\bar\psi_0},i\hbar\xi_0-{H}_{0,I}\big\rangle
\end{equation}
where $\rho_{\bar\psi_0}=\bar{\psi}_0\bar{\psi}_0^\dagger$ and we have introduced the following definitions
\begin{equation}\label{defs}
\psi_I=U_0^{-1}\psi
\,,\qquad
H_{j,I}=U_0^{-1}H_{j}U_0
\,,\qquad
\xi_0={U}_0^{-1}\dot{U}_0
\,,\qquad
\xi_I=U_0^{-1}\xi U_0=\operatorname{Ad}_{U_0^{-1}\!}\xi
\,.
\end{equation}
In order to write the resulting equations of motion, we start by using the last two definitions in \eqref{defs} to compute the variations
\begin{equation}\label{semidvars}
\delta(\xi_0,\xi_I)=\big(\dot{\eta}_0+[\xi_0,\eta_0], \dot{\eta}_I+[\xi_0,\eta_I]-[\eta_0,\xi_I]+[\eta_I,\xi_I]\big)
\end{equation}
where $\eta_0=U_0^{-1}\delta U_0$ and $\eta_I=\operatorname{Ad}_{U_0^{-1}}((\delta U) U^{-1})$. One recognizes that the variations \eqref{semidvars} are Euler-Poincar\'e variations of the type $\delta\nu=\dot{\zeta}+[\zeta,\nu]_\mathfrak{g}$, where $[\cdot\,,\cdot]_\mathfrak{g}$ is the Lie bracket on $\mathfrak{g}=\mathfrak{u}_{0}(\mathscr{H})\,\circledS\,\mathfrak{u}(\mathscr{H})$, that is the Lie algebra of the semidirect product group $\mathcal{U}_{0}(\mathscr{H})\,\circledS\,\mathcal{U}(\mathscr{H})$. Here, the group $\mathcal{U}_{0}(\mathscr{H})$ is a copy of the unitary group $\mathcal{U}(\mathscr{H})$ and one thinks of $\mathcal{U}_{0}(\mathscr{H})$ as accounting for the propagators $U_0\in\mathcal{U}_{0}(\mathscr{H})$.
Computation of the other variations by using \eqref{defs} yields
\[
\delta\rho_{\psi_I}=[\eta_I-\eta_0,\rho_{\psi_I}]
\,,\qquad
\delta H_{j,I}=[H_{j,I},\eta_0]
\,,
\]
so that the variational principle $\int_{t_1}^{t_2\!}l(\xi_0,\xi_I,\rho_I,H_{0,I},H_{1,I})\,{\rm d} t=0$ produces the following equations of motion for an arbitrary Lagrangian $l$:
\begin{align}
&\frac{{\rm d}}{{\rm d} t}\frac{\delta l}{\delta\xi_{I}} - \left[ \xi_{I},\frac{\delta l}{\delta\xi_{I}} \right] + \left[ \xi_{0},\frac{\delta l}{\delta \xi_{I}}\right] + \left[\rho_{\psi_I},\frac{\delta l}{\delta\rho_{\psi_I}} \right]
= 0\,, \label{EQ:DiracEP1}
\\
&\frac{{\rm d}}{{\rm d} t}\frac{\delta l}{\delta\xi_{0}} + \left[ \xi_{0},\frac{\delta l}{\delta\xi_{0}} \right] + \left[ \xi_{I},\frac{\delta l}{\delta\xi_{I}} \right]
-\left[\rho_{\psi_I},\frac{\delta l}{\delta\rho_{\psi_I}} \right]
- \left[ H_{0,I},\frac{\delta l}{\delta H_{0,I}} \right] - \left[ H_{1,I},\frac{\delta l}{\delta H_{1,I}} \right] = 0 \,,
\label{EQ:DiracEP2}
\\
&
\dot{\rho}_{I} = [\xi_{I}- \xi_{0},\rho_{I}] \,, \qquad
\dot{H}_{j,I} = [H_{j,I},\xi_{0}]\,.
\label{EQ:DiracEP3}
\end{align}
Then, computing the variational derivatives of the Lagrangian \eqref{DFLagrInt} and replacing them into \eqref{EQ:DiracEP1} and \eqref{EQ:DiracEP2} gives
\[
i\hbar\big[ \xi_{I}, \rho_{\psi_I} \big]= \big[H_{0,I}+H_{1,I}, \rho_{\psi_I} \big]
\,,\qquad\
\big[ i\hbar\xi_{0}-H_{0,I},\rho_{\bar\psi_0} \big]=0
\]
As seen in Section \ref{sec:HeisDF}, the second relation above is solved by
\[
\xi_0=-i\hbar^{-1\!}H_{0,I}+\{\boldsymbol{1}-2\rho_{\bar\psi_0},\kappa\}
\,,
\]
so the second in \eqref{EQ:DiracEP3} gives
\[
\dot{H}_{0,I} = [H_{0,I},\{\boldsymbol{1}-2\rho_{\bar\psi_0},\kappa\}]
\,,\qquad
\dot{H}_{1,I} = i\hbar^{-1\!}[H_{0,I},H_{1,I}]+[H_{1,I},\{\boldsymbol{1}-2\rho_{\bar\psi_0},\kappa\}]\,.
\]
On the other hand, the first in \eqref{EQ:DiracEP3} becomes
\[
\dot{\rho}_{\psi_I}=i\hbar^{-1}\big[{\rho}_{\psi_I},H_{1,I}\big]+\big[{\rho}_{\psi_I},\{\boldsymbol{1}-2\rho_{\bar\psi_0},\kappa\}\big]
\,.
\]
Then, we notice that the choice $\kappa=0$ returns the usual quantum dynamics in the Dirac picture. As we know from Section \ref{sec:HeisDF}, the $\kappa$-terms do not change the overall physical content of the dynamics. For example, a direct calculation verifies the following energy conservations:
\[
\frac{{\rm d}}{{\rm d} t}\,\big\langle\rho_{\psi_I}|H_{0,I}+H_{1,I}\big\rangle=0
\,,\qquad\
\frac{{\rm d}}{{\rm d} t}\,\big\langle\rho_{\bar\psi_0}|H_{0,I}\big\rangle=0
\,.
\]
As we have seen, the application of Euler-Poincar\'e reduction theory reveals the geometric features emerging in the Heisenberg picture of quantum dynamics. These geometric features reveal the form \eqref{propagator} of the propagator equation, without affecting the physical content of quantum dynamics. When this form of the propagator equation is considered in the Dirac picture, this introduces extra terms in the dynamics, which still preserve the total energy of the system.
In the next Section, we shall formulate a new variational principle for the interaction of classical and quantum degrees of freedom.
\section{Classical-quantum variational principles}
The interplay of quantum and classical degrees of freedom has always attracted much attention in quantum mechanics. For example, the consistent formulation of hybrid quantum-classical models in molecular dynamics remains an outstanding issue \cite{Salcedo2012}. In this section, we present a geometric formulation of the most elementary system coupling classical and quantum dynamics. This is given by combining the Ehrenfest equations for the expectation of the canonical variables with the Schr\"odinger/Liouville equation for the quantum degrees of freedom. More particularly, we shall present a novel variational principle for the Ehrenfest mean field model and in more generality for expectation value dynamics.
\subsection{The classical-quantum mean field model}
In order to approach the dynamics of quantum expectations, we observe that the mean field closure of any classical-quantum system can be derived in first instance by the following Lagrangian
\begin{equation}\label{miguel}
L({\mathbf{z}},\dot{\mathbf{z}},\psi,\dot\psi)=\frac12\,\dot{\mathbf{z}}\cdot\Bbb{J}{\mathbf{z}}+\langle\psi, i\hbar\dot\psi - H({\mathbf{z}})\psi\rangle
\,,
\end{equation}
where $-\Bbb{J}_{ij\,}{\rm d} z^i\wedge{\rm d} z^j$ is the canonical symplectic form and $H({\mathbf{z}})$ is a Hermitian operator depending on the classical degrees of freedom ${\mathbf{z}}=({\mathbf{x}},{\mathbf{p}})$. Indeed, the corresponding Euler-Lagrange equations yield
\begin{equation}
\dot{\mathbf{z}}=\Bbb{J}\left\langle\psi|\nabla_{\mathbf{z}} H({\mathbf{z}})\psi\right\rangle
\,,\qquad\qquad
i\hbar\dot\psi=H({\mathbf{z}})\psi\,,
\label{HybridEqs1}
\end{equation}
thereby recovering the ordinary mean field model of classical-quantum dynamics (see e.g. equations (12.2)-(12.4) in \cite{Wyatt2006}). Here, purely classical dynamics is recovered by the phase type Hamiltonian $H({\mathbf{z}})={\sf h}({\mathbf{z}})I$ (here, $I$ denotes the identity operator on $\mathscr{H}$), while purely quantum dynamics is recovered when ${\nabla_{\mathbf{z}} H({\mathbf{z}})=0}$.
An Euler-Poincar\'e formulation of the above equations can again be obtained by letting the quantum state evolve under unitary transformations. This leads to a coupled Euler-Lagrange equation for $({\mathbf{z}},\dot{\mathbf{z}})$ and the Euler-Poincar\'e equations for the quantum dynamics (expressed in terms of either $\psi$ or its density matrix). However, in order to find a full set of Euler-Poincar\'e equations that includes the classical evolution, we may choose to evolve the phase-space vector ${\mathbf{z}}$ under the action of the Heisenberg group (i.e., phase-space translations), which is prominent in the theory of quantum coherent states. To this purpose, consider a curve $h(t)=(\mathbf{h}(t),\varphi(t))$ in the Heisenberg group $\mathcal{H}(\Bbb{R}^{2n})\simeq \Bbb{R}^{2n+1}$ and let the phase space vector ${\mathbf{z}}$ evolve as
\begin{equation}\label{DirectProdAct}
{\mathbf{z}}(t)={\mathbf{z}}_0+\mathbf{h}(t)
\,.
\end{equation}
Also, we recall $\psi(t)=U(t)\psi_0$. Then, upon inserting the auxiliary phase factor $\varphi$ in the Lagrangian \eqref{miguel}, the latter becomes
\[
L_{{\mathbf{z}}_0,\psi_0}({h},\dot{{h}},U,\dot{U})=\frac12\,\dot{\mathbf{h}}\cdot\Bbb{J}({\mathbf{z}}_0+\mathbf{h})+\big\langle U\psi_0, \big(i\hbar\dot{U}+\dot\varphi U - H({\mathbf{z}}_0+\mathbf{h})U\big)\psi_0\big\rangle
\,.
\]
The above Lagrangian is of the type
\[
L_{{\mathbf{z}}_0,\psi_0}:T\mathcal{H}(\Bbb{R}^{2n})\!\times T\mathcal{U}(\mathscr{H})\to\Bbb{R}
\]
and its dynamics can be approached by Euler-Poincar\'e reduction. Therefore, in order to find an expression for the reduced variable $\zeta:=\dot{h}h^{-1}$, we define the Lie algebra element
\[
(\boldsymbol{\zeta},\phi):=\left.\frac{{\rm d}}{{\rm d} s}\right|_{s=0\!}\left(g(s)h^{-1}\right)
\in\mathfrak{h}(\Bbb{R}^{2n})
\]
where $g(s)=(\mathbf{g}(s),\vartheta(s))\in\mathcal{H}(\Bbb{R}^{2n})$ is a curve such that $g(0)=h$ and ${g}'(0)=\dot{h}$ (for some fixed time). Here, we recall the Heisenberg group operation
\begin{equation}\label{Heismult}
gh
=\left(\mathbf{g}+\mathbf{h},\, \vartheta+\varphi+\frac12\,\mathbf{g}\cdot\Bbb{J}\mathbf{h}\right)
\,,\qquad
\qquad
\forall
g,h\in\mathcal{H}(\Bbb{R}^{2n})
\,,
\end{equation}
which gives $h^{-1}=(-\mathbf{h},-\varphi)$. Eventually, one finds
\[
\zeta=(\boldsymbol{\zeta},\phi)=\left(\dot{\mathbf{h}},\,\dot{\varphi}-\frac12\,\dot{\mathbf{h}}\cdot\Bbb{J}\mathbf{h}\right)
,
\]
so that the Euler-Poincar\'e Lagrangian is written as
\begin{equation}
\label{HybridLagrangian1}
\ell(\boldsymbol\zeta,\phi,\xi,{\mathbf{z}},\rho_\psi)=\boldsymbol{\zeta}\cdot\Bbb{J}{\mathbf{z}} +\big\langle\rho_\psi,i\hbar\xi+\phi-H({\mathbf{z}})\big\rangle
\,,
\end{equation}
where we have used the convenient initial condition ${\mathbf{z}}_0=0$ in \eqref{DirectProdAct}, without loss of generality.
Notice, this Lagrangian is of the type
\[
\ell:\big(\mathfrak{h}(\Bbb{R}^{2n})\oplus\mathfrak{u}(\mathscr{H})\big)\times\big(\Bbb{R}^{2n}\times\mathbf{P}_{\!}\mathscr{H}\big)\to\Bbb{R}
\,.
\]
Then, the Euler-Poincar\'e equations follow in the theorem below, upon recalling the infinitesimal adjoint representation
\begin{equation}\label{InfAct}
\operatorname{ad}_{(\boldsymbol\zeta_1,\phi_1)}(\boldsymbol\zeta_2,\phi_2)=(0,- \boldsymbol{\zeta}_1\cdot{\Bbb{J}\boldsymbol{\zeta}_2})
\end{equation}
in the Heisenberg Lie algebra $\mathfrak{h}(\Bbb{R}^{2n})$.
The present treatment is now extended to the case of mixed quantum states.
\begin{theorem}
Consider the variational principle
\[
\delta\int_{t_1}^{t_2}\!\Big(\boldsymbol{\zeta}\cdot\Bbb{J}{\mathbf{z}} +\big\langle\rho,i\hbar\xi+\phi-H({\mathbf{z}})\big\rangle\Big)\,{\rm d} t = 0
\]
and the variations
\[
(\delta\boldsymbol\zeta,\delta\phi) =
\big(\dot{\boldsymbol\gamma},\,\dot{\theta}+\boldsymbol\zeta\cdot\Bbb{J}\boldsymbol\gamma\big)
\,,\qquad
\delta\xi=\dot{\eta}-[\xi,\eta]
\,
\,,
\qquad \delta{\mathbf{z}} = \boldsymbol{\gamma}\,,
\qquad
\delta\rho = [\eta,\rho]
\,,
\]
where $({\boldsymbol\gamma},{\theta})$ and $\eta$ are arbitrary and vanish at the endpoints. Together with the auxiliary equations
\[
\dot{\mathbf{z}}=\boldsymbol{\zeta}\,,\qquad
\dot\rho=[\xi,\rho]
\,,
\]
this variational principle is equivalent to the equations of motion
\begin{equation*}
\dot{{\mathbf{z}}} = \Bbb{J}\langle \rho | \nabla_{{\mathbf{z}}}H({\mathbf{z}}) \rangle
\,,\qquad\qquad i\hbar\dot{\rho} = \left[ H({\mathbf{z}}),\rho \right]\,.
\end{equation*}
\end{theorem}
\paragraph{Proof.}
Consider the general Lagrangian of the form $\ell(\boldsymbol\zeta,\phi,\xi,{\mathbf{z}},\rho_\psi)$. By direct substitution of the variations into the variational principle
\begin{align*}
\delta\int_{t_1}^{t_2}\!\Big( \left\langle \frac{\delta\ell}{\delta\boldsymbol{\zeta}}, \, \dot{\boldsymbol\gamma} \right\rangle
+ \left\langle \frac{\delta\ell}{\delta\phi}, \, \dot{\theta}+\boldsymbol\zeta\cdot\Bbb{J}\boldsymbol\gamma \right\rangle
+ \left\langle \frac{\delta\ell}{\delta\xi}, \, \dot{\eta}-[\xi,\eta] \right\rangle
+ \left\langle \frac{\delta\ell}{\delta{\mathbf{z}}}, \, \boldsymbol{\gamma} \right\rangle
+ \left\langle \frac{\delta\ell}{\delta\rho}, \, [\eta,\rho] \right\rangle
\Big) {\rm d} t = 0\,,
\end{align*}
one writes the Euler-Poincar\'e equations as
\[
-\frac{{\rm d}}{{\rm d} t}\frac{\delta\ell}{\delta\boldsymbol{\zeta}} + \frac{\delta\ell}{\delta\phi}\Bbb{J}\boldsymbol{\zeta} + \frac{\delta\ell}{\delta{\mathbf{z}}} = 0\,, \qquad
\frac{{\rm d}}{{\rm d} t} \frac{\delta\ell}{\delta\phi} = 0\,, \qquad
-\frac{{\rm d}}{{\rm d} t} \frac{\delta\ell}{\delta\xi} + \left[ \xi,\frac{\delta\ell}{\delta\xi} \right] + \left[ \rho, \frac{\delta\ell}{\delta\rho} \right] =0\,.
\]
In particular, for the Lagrangian \eqref{HybridLagrangian1}, we have
\[
\frac{\delta\ell}{\delta\boldsymbol{\zeta}} = \Bbb{J}{\mathbf{z}}\,, \qquad
\frac{\delta\ell}{\delta\phi} = \langle \rho |1 \rangle\,, \qquad
\frac{\delta\ell}{\delta\xi} = -i\hbar \rho\,, \qquad
\frac{\delta\ell}{\delta\rho} = i\hbar\xi + \phi - H({\mathbf{z}})\,, \qquad
\frac{\delta\ell}{\delta{\mathbf{z}}} = -\Bbb{J}\boldsymbol{\zeta} - \langle \rho | \nabla_{{\mathbf{z}}}H({\mathbf{z}}) \rangle\,,
\]
such that the Euler-Poincar\'e equations yield
\begin{align*}
-\Bbb{J}\dot{{\mathbf{z}}} + \Bbb{J}\boldsymbol{\zeta} - \Bbb{J}\boldsymbol{\zeta} - \langle \rho | \nabla_{{\mathbf{z}}}H({\mathbf{z}}) \rangle& = 0
\\
-i\hbar\dot{\rho} + \left[ \xi,-i\hbar\rho \right] - \left[ \rho, i\hbar\xi - H({\mathbf{z}}) \right] &=0 \,.
\end{align*}
thereby completeing the proof. \quad$\blacksquare$
\noindent
The observation that hybrid classical-quantum dynamics can be expressed by using the Heisenberg and unitary groups motivates us to investigate further the interplay between these two symmetry structures. The next Section shows that combining the two groups into a semidirect product yields the variational formulation of quantum expectation dynamics.
\subsection{The semidirect product $\mathcal{H}(\Bbb{R}^{2n})\,\circledS\; \mathcal{U}(\mathscr{H})$}
While the previous Section used the direct product $\mathcal{H}(\Bbb{R}^{2n})\!\times\mathcal{U}(\mathscr{H})$ group structure to obtain hybrid classical-quantum dynamics, we shall now illustrate how constructing the semidirect product $\mathcal{H}(\Bbb{R}^{2n})\,\circledS\,\mathcal{U}(\mathscr{H})$ allows to shed new light on the dynamics of expectation values, thereby extending Ehrenfest theorem to more general situations.
The semidirect product $\mathcal{H}(\Bbb{R}^{2n})\,\circledS\,\mathcal{U}(\mathscr{H})$ can be constructed upon using the celebrated displacement operator from the theory of coherent quantum states. This is defined as follows
\[
U_h\psi(\boldsymbol{x})=e^{-\frac{i}{\hbar}\big(\varphi+\frac{\mathbf{h}_p\cdot\mathbf{h}_q}{2}-\mathbf{h}_p\cdot\boldsymbol{x}\big)}\psi(\boldsymbol{x}-\mathbf{h}_q)
\,,\qquad
\qquad
\forall
h=(\mathbf{h},\varphi)\in\mathcal{H}(\Bbb{R}^{2n})
\,,
\]
where the phase space vector $\mathbf{h}\in\Bbb{R}^{2n}$ is expressed as $\mathbf{h}=(\mathbf{h}_q,\mathbf{h}_p)$. This operator defines a group homomorphism that can be used to construct the following product rule in $\mathcal{H}(\Bbb{R}^{2n})\,\circledS\,\mathcal{U}(\mathscr{H})$:
\begin{equation} \label{productrule}
(h_1,U_1)(h_2,U_2)=\big(h_1h_2,U_1({U_{h_1}}U_2U_{h_1}^\dagger)\big)
\,,\qquad
\qquad
\forall
h_1,h_2\in\mathcal{H}(\Bbb{R}^{2n})
\,,
\quad
\forall U_1,U_2\in\mathcal{U}(\mathscr{H})
\,,
\end{equation}
where $h_1h_2$ is the product rule in the Heisenberg group, already defined in \eqref{Heismult}. Notice, upon denoting ${Z}=({Q},{P})$ (quantum canonical operators), the displacement operator $U_h$ leads to the following Lie algebra homomorphism $\iota:\mathfrak{h}(\Bbb{R}^{2n})\to\mathfrak{u}(\mathscr{H})$
\[
\iota(\zeta)=-i\hbar^{-1}(\phi+\boldsymbol{\zeta}\cdot\Bbb{J}{Z})
\,,\qquad
\qquad
\forall
\zeta=(\boldsymbol{\zeta},\phi)\in\mathfrak{h}(\Bbb{R}^{2n})\,,
\]
which occurs in the Lie bracket structure on $\mathfrak{h}(\Bbb{R}^{2n})\,\circledS\,\mathfrak{u}(\mathscr{H})$, given by
\[
\operatorname{ad}_{(\zeta_1,\xi_1)}(\zeta_2,\xi_2)=\Big(\operatorname{ad}_{\zeta_1\!}\zeta_2,\, [\xi_1,\iota(\zeta_2)]-[\xi_2,\iota(\zeta_1)]+[\xi_1,\xi_2]\Big)
\,,
\]
where the operator `$\operatorname{ad}$' appearing in the first slot on the RHS is the infinitesimal adjoint action on $\mathfrak{h}(\Bbb{R}^{2n})$, as it was defined in \eqref{InfAct}. No confusion should arise from this notation.
In order to construct a dynamical theory by using the group structure above in the Lagrangian \eqref{miguel}, we need to find an action of $\mathcal{H}(\Bbb{R}^{2n})\,\circledS\,\mathcal{U}(\mathscr{H})$ on the space $\Bbb{R}^{2n\!}\times S(\mathscr{H})$. This task can be achieved by computing the coadjoint representation on the semidirect product. This computation can benefit from the following property
\begin{lemma}[Equivariance] \label{equivariance}
With the notation above, the following relations hold
\[
U_h Z U_h^\dagger= Z-\mathbf{h}{I}
\,,
\qquad\qquad
\iota(\operatorname{Ad}_{h}\zeta)={U_h\,}\iota(\zeta)\,U_h^\dagger\,,
\]
where ${I}$ is the identity operator on $\mathscr{H}$ and $\operatorname{Ad}_h\zeta=\left( \boldsymbol{\zeta}, \,\phi + \mathbf{h} \cdot\Bbb{J}\boldsymbol{\zeta} \right)$ is the adjoint representation on $\mathcal{H}(\Bbb{R}^{2n})$.
\end{lemma}
\paragraph{Proof.}
The first relation is easily proved by a direct verification. The first component reads as follows:
\begin{align*}
\big(U_h{X}U_h^\dagger\big)\psi(\boldsymbol{x}) &= \big( U_h \boldsymbol{x}\big)\left[ e^{\frac{i}{\hbar}\varphi} e^{-i\frac{\mathbf{h}_p\cdot\mathbf{h}_q}{2\hbar}} e^{-i\frac{\mathbf{h}_p\cdot\boldsymbol{x}}{\hbar}} \psi(\boldsymbol{x}+\mathbf{h}_q) \right]
\\
&=e^{-i\frac{\mathbf{h}_p\cdot\mathbf{h}_q}{\hbar}} e^{i\frac{\mathbf{h}_p\cdot\boldsymbol{x}}{\hbar}} ( \boldsymbol{x}-\mathbf{h}_q ) e^{-i\frac{\mathbf{h}_p\cdot (\boldsymbol{x}-\mathbf{h}_q)}{\hbar}} \psi(\boldsymbol{x})
\\
&= ( \boldsymbol{x}-\mathbf{h}_q ) \psi(\boldsymbol{x})
\,.
\end{align*}
Similarly, the second component reads
\begin{align*}
\big(U_h{P}U_h^\dagger\big)\psi(\boldsymbol{x}) &= -i\hbar\,\big( U_{h} \nabla \big)\left[ e^{\frac{i}{\hbar}\varphi} e^{-i\frac{\mathbf{h}_p\cdot\mathbf{h}_q}{2\hbar}} e^{-i\frac{\mathbf{h}_p\cdot\boldsymbol{x}}{\hbar}} \psi(\boldsymbol{x}+\mathbf{h}_q) \right]
\\
&= - U_h \Big[ e^{\frac{i}{\hbar}\varphi} e^{-i\frac{\mathbf{h}_p\cdot\mathbf{h}_q}{2\hbar}} e^{-i\frac{\mathbf{h}_p\cdot\boldsymbol{x}}{\hbar}} \big({\mathbf{h}_p}\, \psi(\boldsymbol{x}+\mathbf{h}_q) + i\hbar \nabla\psi(\boldsymbol{x}+\mathbf{h}_q) \big)\Big]
\\
&=
(-{\mathbf{h}_p}+P) \psi(\boldsymbol{x})
\,.
\end{align*}
Combining both components proves the first relation in the lemma.
The second relation follows by direct substitution
\begin{multline*}
\iota(\operatorname{Ad}_{h}\zeta) = \iota\Big( \boldsymbol{\zeta}, \,\phi + \mathbf{h} \cdot\Bbb{J}\boldsymbol{\zeta} \Big) = -i\hbar^{-1}\Big( \phi + \mathbf{h}\cdot\Bbb{J}\boldsymbol{\zeta} - Z\cdot\Bbb{J}\boldsymbol{\zeta} \Big) = -i\hbar \left( \phi - \left( Z - \mathbf{h}{I} \right)\cdot \Bbb{J}\boldsymbol{\zeta} \right)
=
\\
= -i\hbar \big( \phi - \big( U_h Z U_h^\dagger\big)\cdot \Bbb{J}\boldsymbol{\zeta} \big)
= U_h\big(-i\hbar \big( \phi +\Bbb{J} Z \cdot \boldsymbol{\zeta} \big) \big) U_h^\dagger
={U_h\,}\iota(\zeta)\,U_h^\dagger
\,,
\end{multline*}
thereby completing the proof.
$\quad\blacksquare$
\noindent
Eventually, by making use of the previous relations in the definition of coadjoint representation, one finds the following expression:
\[
\operatorname{Ad}^*_{(h,U)}(\nu,\mu)=\Big(\boldsymbol\nu-\alpha\Bbb{J}\mathbf{h}+\big\langle\mu-U^\dagger\mu U,i\hbar^{-1}\Bbb{J}Z\big\rangle, \, \alpha,\,U_h^\dagger U^\dagger\mu UU_h\Big)
\]
where we have used the notation $\nu=(\boldsymbol\nu,\alpha)\in \mathfrak{h}(\Bbb{R}^{2n})^*\simeq\Bbb{R}^{2n+1}$. This coadjoint representation is computed explicilty in the Appendix \ref{Appendix1}.
Then, upon fixing the invariant set $\alpha=1$ and by introducing the variables ${\mathbf{z}}=-\Bbb{J}\boldsymbol\nu$ and $\rho_\psi=i\hbar^{-1}\mu$, we obtain the following action of $\mathcal{H}(\Bbb{R}^{2n})\,\circledS\,\mathcal{U}(\mathscr{H})$ on the space $\Bbb{R}^{2n\!}\times \mathbf{P}_{\!}\mathscr{H}$:
\[
\Phi_{(h,U)}({\mathbf{z}},\rho_\psi)=\Big({\mathbf{z}}-\mathbf{h}+\langle UZU^\dagger-Z \rangle,\,U_h^\dagger U^\dagger\rho_\psi U U_h\Big)
\,,
\]
where we have used the expectation value notation $\langle A\rangle=\langle A | \rho_\psi\rangle$.
\subsection{Geometry of quantum expectation dynamics}
At this point, the semidirect product $\mathcal{H}(\Bbb{R}^{2n})\,\circledS\; \mathcal{U}(\mathscr{H})$ has been characterized and it has been showed to possess an action on the classical-quantum phase space $\Bbb{R}^{2n\!}\times \mathbf{P}_{\!}\mathscr{H}$. Then, we consider the evolution of the classical-quantum variables $({\mathbf{z}},\rho_\psi)$ under the action of $(h^{-1},U^{-1})$, which then gives
\begin{equation}\label{SemiDEvol}
{\mathbf{z}}(t)={\mathbf{z}}_0+\mathbf{h}(t)+\big\langle U(t)^\dagger ZU(t)-Z\,\big|\,\rho_{\psi_0} \big\rangle
\,,\qquad\quad
\rho_\psi(t)=U_h(t) U(t)\rho_{\psi_0}U(t)^\dagger U_h(t)^\dagger
\,.
\end{equation}
The evolution above has the following crucial feature:
\[
{\mathbf{z}}(t) -\big\langle Z\big|\rho_{\psi}(t) \big\rangle={\mathbf{z}}_0
-\langle Z|\rho_{\psi_0} \rangle\,,
\]
as it is verified upon computing
\[
\big\langle Z\,\big|\,U(t)\rho_{\psi_0}U(t)^\dagger \big\rangle=
\big\langle Z\,\big|\,U_h(t)^\dagger\rho_{\psi}(t)U_h(t) \big\rangle
=
\big\langle U_h(t) ZU_h(t)^\dagger\,\big|\,\rho_{\psi}(t) \big\rangle=
\big\langle Z-\mathbf{h}(t)I\,\big|\,\rho_{\psi}(t) \big\rangle\,.
\]
Therefore, in order to study expectation value dynamics, one can simply initiate the evolution under the initial condition ${\mathbf{z}}_0=\langle Z\,|\,\rho_{\psi_0} \rangle$, which is then replaced in \eqref{SemiDEvol}. Moreover, the evolution above, produces the equations of motion
\[
\dot{{\mathbf{z}}} = \boldsymbol{\zeta} - \big\langle [\rho,{Z}],\xi \big\rangle\,,
\qquad\quad
\dot\rho = \left[ i\hbar^{-1}{Z}\cdot\Bbb{J}\boldsymbol{\zeta}+ \xi,\rho \right]
\]
where $\zeta=\dot{\mathbf{h}}$ and $\xi=U_{h}\dot{U}U^{\dagger}U_{h}^{\dagger}$. Analogous expressions hold for the variations $(\delta{\mathbf{z}},\delta\rho)$.
At this point, we consider the Euler-Poincar\'e Lagrangian of the classical-quantum mean field model \eqref{HybridLagrangian1}. Although that was written previously on the space $(\mathfrak{h}(\Bbb{R}^{2n})\oplus\mathfrak{u}(\mathscr{H}))\times(\Bbb{R}^{2n}\times\mathbf{P}_{\!}\mathscr{H})$, we now change perspective and we interpret the same expression \eqref{HybridLagrangian1} for $\ell(\boldsymbol\zeta,\phi,\xi,{\mathbf{z}},\rho_\psi)$ as a Lagrangian of the type
\[
\ell:\big(\mathfrak{h}(\Bbb{R}^{2n})\,\circledS\,\mathfrak{u}(\mathscr{H})\big)\times\big(\Bbb{R}^{2n}\times\mathbf{P}_{\!}\mathscr{H}\big)\to\Bbb{R}
\,.
\]
Notice that the Hamiltonian operator $H({\mathbf{z}})$ depends on the classical variable ${\mathbf{z}}$, which has to be interpreted as the expectation value $\langle Z|\rho_{\psi}\rangle$. This amounts to consider quantum systems for which the total energy can be written in terms of both the quantum state $\rho_\psi$ and its corresponding expectation values ${\mathbf{z}}=\langle Z|\rho_{\psi}\rangle$. (Notice that this is a very general case, as it shown by considering the kinetic energy expression $\langle P^2\rangle_\psi/2=\langle p\rangle^2/2+\langle P-\langle p\rangle\rangle_{\!\psi}^{\,2}/2$).
\begin{theorem}
Consider the Lagrangian \eqref{HybridLagrangian1} and its associated variational principle for mixed quantum states
\[
\delta\int_{t_1}^{t_2}\bigg(\boldsymbol{\zeta}(t)\cdot\Bbb{J}{\mathbf{z}}(t) +\Big\langle\rho(t),i\hbar\xi(t)+\phi(t)-H({\mathbf{z}}(t))\Big\rangle\bigg){\rm d} t=0
\,,
\]
with variations
\begin{align*}
\delta\boldsymbol{\zeta} &= \dot{\boldsymbol{\gamma}}\,,
\qquad\ \delta\phi = \dot{\theta} - \boldsymbol{\zeta}\cdot\Bbb{J}\boldsymbol{\gamma}\,,
\\
\delta\xi &= \dot{\eta} - i\hbar^{-1}\big(\left[\xi,{Z}\cdot\Bbb{J}\boldsymbol{\gamma}\right] - \left[ \eta,{Z}\cdot\Bbb{J}\boldsymbol{\zeta} \right]\big) + [\eta,{\xi}]\,,
\\
\delta{\mathbf{z}} &= \boldsymbol{\gamma} - \big\langle [\rho,Z],\eta \big\rangle\,,
\\
\delta\rho &= \left[ i\hbar^{-1}{Z}\cdot\Bbb{J}\boldsymbol{\gamma}+ \eta,\rho \right]\,,
\end{align*}
where $\boldsymbol{\gamma}$, $\theta$ and $\eta$ are arbitrary and vanish at the endpoints. Then, this is equivalent to the following equations of motion
\begin{align}\label{michael}
&\dot{{\mathbf{z}}} = \Bbb{J}\nabla_{{\mathbf{z}}\!}\left\langle\rho | H({\mathbf{z}})\right\rangle -i\hbar^{-1} \big\langle {Z}\big|\big[ H({\mathbf{z}}),\rho \big] \big\rangle
\,,
\\
&i\hbar\dot{\rho} = \left[ H({\mathbf{z}}),\rho \right]+\nabla_{{\mathbf{z}}\!} \left\langle\rho|H({\mathbf{z}})\right\rangle\cdot\left[{Z},\rho \right].
\label{frank}
\end{align}
\end{theorem}
\paragraph{Proof.}
This follows by a direct subsistution of the variations in the action principle. We have
\begin{align*}
\delta& \int \bigg( \boldsymbol{\zeta}(t)\cdot\Bbb{J}{\mathbf{z}}(t) +\Big\langle\rho(t),i\hbar\xi(t)+\phi(t)-H({\mathbf{z}}(t))\Big\rangle \bigg) \ {\rm d} t=
\\
=&\int \bigg(
\Bbb{J}{\mathbf{z}}\cdot\delta\boldsymbol{\zeta}
+\left\langle \rho, \delta\phi \right\rangle
- \left\langle i\hbar\rho,\delta\xi \right\rangle
- \left( \Bbb{J}\boldsymbol{\zeta} + \left\langle\rho | \nabla_{{\mathbf{z}}}H({\mathbf{z}}) \right\rangle \right)\cdot \delta{\mathbf{z}}
+\left\langle i\hbar\xi +\phi-H({\mathbf{z}}), \delta\rho \right\rangle
\bigg) \ {\rm d} t
\\
=&\int \bigg(
\Bbb{J}{\mathbf{z}}\cdot\dot{\boldsymbol{\gamma}}
+\left\langle \rho, \dot{\theta} - \boldsymbol{\zeta}\cdot\Bbb{J}\boldsymbol{\gamma} \right\rangle
- \left\langle i\hbar\rho,\dot{\eta} - i\hbar^{-1}\big(\left[\xi,{Z}\cdot\Bbb{J}\boldsymbol{\gamma}\right] - \left[ \eta,{Z}\cdot\Bbb{J}\boldsymbol{\zeta} \right]\big) + [\eta,{\xi}] \right\rangle +
\\
&\ - \left( \Bbb{J}\boldsymbol{\zeta} + \left\langle\rho | \nabla_{{\mathbf{z}}}H({\mathbf{z}}) \right\rangle \right)\cdot \left(\boldsymbol{\gamma} - \big\langle [\rho,Z],\eta \big\rangle\right)
+\left\langle i\hbar\xi +\phi-H({\mathbf{z}}), \left[ i\hbar^{-1}{Z}\cdot\Bbb{J}\boldsymbol{\gamma}+ \eta,\rho \right] \right\rangle
\bigg) \ {\rm d} t
\\
=&\int\bigg( \Big\langle
- \Bbb{J}\dot{{\mathbf{z}}}
- \left\langle \rho | \nabla_{{\mathbf{z}}}H({\mathbf{z}}) \right\rangle
+ \left\langle \left[ i\hbar^{-1}\rho,H({\mathbf{z}}) \right], \Bbb{J}Z \right\rangle
, \boldsymbol{\gamma} \Big\rangle
\\
&\ + \left\langle
i\hbar\dot{\rho}
+ \left[ \rho, \hat{Z}\cdot \left\langle \rho | \nabla_{{\mathbf{z}}}H({\mathbf{z}}) \right\rangle \right]
+\left[ \rho,H({\mathbf{z}}) \right]
, \boldsymbol{\eta} \right\rangle\bigg) \ {\rm d} t
\end{align*}
Then, since $\boldsymbol{\gamma}$, $\theta$ and $\eta$ are arbitrary and vanish at the endpoints, the proof follows. $\quad\blacksquare$
\medskip
\noindent
In order to understand how the above result is related to the usual Ehrenfest equations for quantum expectation dynamics, we immediately observe how these equations \eqref{michael}-\eqref{frank} are recovered (along with the evolution of $\rho$) in the case when $\nabla_{\mathbf{z}} H({\mathbf{z}})=0$. As it was pointed out previously, the new feature of equations \eqref{michael}-\eqref{frank} lies in the fact that the expectation values have been considered as independent variables already occurring in the expression of the conserved total energy $\langle H({\mathbf{z}})\rangle$. This confers the system \eqref{michael}-\eqref{frank} a hybrid classical-quantum structure. Indeed, one observes that new coupled classical-quantum terms appear in Ehrenfest dynamics: these are the first term on the RHS of \eqref{michael} and the second term on the RHS of \eqref{frank}.
Notice, the first term on the RHS of \eqref{michael} does not involve the quantum scales given by $\hbar$. For example, a purely classical system is given by a quantum phase-type Hamiltonian operator of the form $H({\mathbf{z}})=h({\mathbf{z}})\boldsymbol{1}$, where $h({\mathbf{z}})$ is the classical expression of the Hamiltonian. In this case, while equation \eqref{michael} recovers classical Hamilton's equations, the quantum evolution \eqref{frank} specializes to coherent state dynamics of the type
\[
i\hbar\dot{\rho} = \nabla_{\mathbf{z}} h\cdot\left[ {Z},\rho \right].
\]
This establishes how quantum states evolve under the action of purely classical degrees of freedom, thereby enlightening the interplay between classical and quantum dynamics. The same equation can also be obtained by linearizing the quantum Hamiltonian operator $H(Z)$ around the expectation values (i.e. in the limit $Z\to{\mathbf{z}} I$), as prescribed by Littlejohn's nearby orbit approximation for semiclassical mechanics \cite{Littlejohn86}.
\section{Conclusions}
This paper investigated the geometric symmetry properties of quantum and classical-quantum variational principles. Upon departing from the Dirac-Frenkel Lagrangian, different quantum mechanics pictures were recovered from the same variational principle, upon making extensive use of Euler-Poincar\'e theory. This reduction by symmetry naturally leads to consider the Hopf bundle as the natural setting for the Schr\"odinger picture of pure state dynamics, as already proposed by Kibble \cite{kibble1979geometrization}. In addition, new variational principles were presented for mixed state dynamics in both the density matrix and the Wigner-Moyal formulation. Later, new quantum variational principles were also presented for the Heisenberg and Dirac's interaction pictures of quantum dynamics, where the isotropies of the initial state was shown to possess the same role as phases in the Schr\"odinger picture. In particular, Dirac's interaction picture involves the geometric semidirect product structure of two different unitary groups associated to the different quantum propagators arising from the splitting of the Hamiltonian operator.
In the last part of the paper, the Dirac-Frenkel Lagrangian was augmented to account for classical degrees of freedom, incorporating the dynamics of classical motion in hybrid classical-quantum dynamics. As a first step, the mean field model of classical-quantum dynamics was described by using the direct product of the Heisenberg group $\mathcal{H}(\Bbb{R}^{2n})$ (governing classical evolution) and the unitary group $\mathcal{U}(\mathscr{H})$ (governing quantum evolution). Later, Ehrenfest's expectation value dynamics was shown to arise from a novel set of equations, whose dynamics evolves both expectation values and the quantum density matrix under the action of the semidirect product $\mathcal{H}(\Bbb{R}^{2n})\,\circledS\; \mathcal{U}(\mathscr{H})$, whose group and Lie algebra structures were presented. In this context, purely classical dynamics was shown to arise by using Littlejohn's nearby orbit approximation \cite{Littlejohn86}.
\medskip
\subsection*{Acknowledgments} We are indebted with Fran\c{c}ois Gay-Balmaz, Dorje Brody, Jos{\'e} Cari{\~n}ena, Darryl Holm, Alberto Ibort, Henry Jacobs, David Meier, Tudor Ratiu, Paul Skerritt and Alessandro Torrielli for interesting remarks during the development of this work. Financial support by the Leverhulme Trust Research Project Grant 2014-112, the London Mathematical Society Grant No. 31320 (Applied Geometric Mechanics Network), and the EPSRC Grant No. EP/K503186/1 is greatly acknowledged.
\bigskip
|
\section{Introduction}
\label{sec:intro}
\subsection{History}
\label{subsec:history}
\input{history.tex}
\subsection{An appetizing example}
\label{subsec:an_example}
\input{an_example.tex}
\subsection{Summary}
\input{summary.tex}
\subsection{Conventions}
\input{conventions.tex}
\subsection{Acknowledgments}
\input{acknowledgements.tex}
\section{The slice genus lower bounds from separable potentials}
\label{sec:lowerbounds}
\subsection{Reduced cohomology and slice-torus invariants}
\label{sec:red_slice_torus}
\input{red_slice_torus.tex}
\subsection{Unreduced cohomology}
\label{subsec:basic_unred}
\input{basic_unred.tex}
\subsection{Appetizing example revisited}
\label{subsec:example_redux}
\input{example_redux.tex}
\section{Comparing different potentials}
\label{subsec:comparison}
\subsection{The KR-equivalence classes}
\input{generic}
\subsection{A lower bound on the number of KR-equivalence classes}
\label{subsec:bunch}
\input{bunch.tex}
\section{Further illuminating examples}
\label{subsec:further_egs}
\input{more_egs.tex}
\section{Computer calculations}
\label{sec:calc}
\subsection{Bipartite links}
\label{sec:bipartite}
\input{bipartite.tex}
\subsection{A computer program}
\label{subsec:compu}
\label{sec:examples}
\input{compu.tex}
\section{Outlook}
\input{open.tex}
\bibliographystyle{myamsalpha}
|
\section{Introduction}
}%\textcolor{red}{For a positive integer $s\in\mathbb{N}$, the $s$-th (higher or sequential) topological complexity of a path connected space $X$, $\protect\operatorname{TC}_{s}(X)$, is defined in~\cite{Ru10} as the reduced Schwarz genus of the fibration $$e_s=e^X_{s}:X^{J_{s}}\to X^{s}$$ given by $e_{s}(f)=\left(}%\textcolor{red}{f_{1}}(1),\dots,}%\textcolor{red}{f_{s}(1})\right)$. Here $J_{s}$ denotes the wedge of $s$ copies of the closed interval $[0,1]$, in all of which $0\in[0,1]$ is the base point, and we think of an element $f$ in the function space $X^{J_{s}}$ as an $s$-tuple $f=(f_1,\ldots,f_s)$ of paths in $X$ all of which start at a common point. Thus, $\protect\operatorname{TC}_{s}(X)+1$ is the smallest cardinality of open covers $\{U_i\}_i$ of $X^s$ so that, on each $U_i$, $e_s$ admits a section $\sigma_i$. In such a cover, $U_i$ is called a {\it local domain}, the corresponding section $\sigma_i$ is called a {\it local rule}, and the resulting family of pairs $\{(U_i,\sigma_i)\}$ is called a {\it motion planner}. The latter is said to be {\it optimal} if it has $\protect\operatorname{TC}_{s}(X)+1$ local domains.}
\medskip}%\textcolor{red}{For practical purposes, the openness condition on local domains can be replaced (without altering the resulting numeric value of $\protect\operatorname{TC}_{}%\textcolor{red}{s}}(X)$) by the requirement that local domains }%\textcolor{red}{are} pairwise disjoint Euclidean neighborhood retracts (ENR).}
\medskip
}%\textcolor{red}{Since} $e_s$ is the }%\textcolor{red}{standard} fibrational substitute of the diagonal }%\textcolor{red}{inclusion} $$d_s=d^{X}_{s}: X\hookrightarrow X^{s},$$ $\protect\operatorname{TC}_{s}(X)$ coincides with the reduced Schwarz genus of }%\textcolor{red}{$d_{s}$. This suggests part (a) in the following definition, where we allow cohomology with local coefficients:}
\begin{definition} }%\textcolor{red}{Let $X$ be a connected space and $R$ be a commutative ring.}
\begin{itemize}\item[(a)] Given a positive integer $s$, we denote by $\protect\operatorname{zcl}_s\, (H^*(X;}%\textcolor{red}{R}))$ the cup-length of elements in the kernel of the map induced by $d_s$ in cohomology. }%\textcolor{red}{Explicitly}, $\protect\operatorname{zcl}_{s}\, (H^*(X;}%\textcolor{red}{R}))$ is the largest integer $m$ for which there exist cohomology classes $u_i \in H^*(X^s, A_i)$, where $X^s$ is the $s$-th Cartesian }%\textcolor{red}{power} of $X$ and }%\textcolor{red}{each} $A_i$ is a system of local coefficients, such that $}%\textcolor{red}{d_s^*(u_i)} =0$ for $i=1, \ldots, m\hspace{.3mm}$ and $\hspace{.5mm}0 \neq u_1 \otimes \cdots \otimes u_m \in H^*(X^s, \, A_1 \otimes \cdots \otimes A_m)$.
\item[(b)] The homotopy dimension of $X$, $\protect\operatorname{hdim} (X)$, is the smallest dimension of $\protect\operatorname{CW}$ complexes having the homotopy type of $X$. The connectivity of $X$, $\protect\operatorname{conn} (X)$, }%\textcolor{red}{is the largest integer $c$ such that $X$ has trivial homotopy groups in dimensions at most $c$. We set $\protect\operatorname{conn}(X)=\infty$ when no such $c$ exists.}
\end{itemize}
\end{definition}
\begin{proposition}\label{ulbTCn}
For a path connected space $X$, $$ \protect\operatorname{zcl}_{s}\,(H^*(X;}%\textcolor{red}{R})) \leq \protect\operatorname{TC}_s (X) \leq \frac{s \, \protect\operatorname{hdim} (X) }{\protect\operatorname{conn} (X) +1}.$$
}%\textcolor{red}{In particular for every path connected $X$,} $$}%\textcolor{red}{\protect\operatorname{TC}_s(X)\leq s\protect\operatorname{hdim}(X).}$$
\end{proposition}
For a proof see~\cite[Theorem~3.9]{bgrt} }%\textcolor{red}{or, more generally,} \cite[Theorems 4 and 5]{Schwarz66}.
\medskip
}%\textcolor{red}{The spaces we work with arise as follows. For} a positive integer }%\textcolor{red}{$k_i$} consider the }%\textcolor{red}{minimal} cellular structure }%\textcolor{red}{on} the }%\textcolor{red}{$k_i$}-dimensional sphere $S^{}%\textcolor{red}{k_i}}=e^0\cup e^{}%\textcolor{red}{k_i}}$. Here $e^0$ }%\textcolor{red}{is} the base point, which }%\textcolor{red}{is simply denoted} by $e$. Take the product }%\textcolor{red}{(}%\textcolor{red}{therefore} minimal)} }%\textcolor{red}{cell} decomposition }%\textcolor{red}{in}
$$
}%\textcolor{red}{\mathbb{S}{(k_1,\ldots, k_n)}:={S}^{k_1}\times\dots\times{S}^{k_n}}=\hspace{.3mm}}%\textcolor{red}{\bigsqcup_Je_J}
$$
whose cells $e_J$, indexed by subsets $J\subseteq[n]=\{1,\dots,n\}$, are defined as $e_J=\prod_{i=1}^ne^{d_i}$ where $d_i=0$ if $i \notin }%\textcolor{blue}{J}$ and $d_i=k_i$ if $i \in }%\textcolor{blue}{J}$. Explicitly, $$}%\textcolor{red}{e_J} =\left\{}%\textcolor{red}{(x_1,\ldots,x_n)}\in }%\textcolor{red}{\mathbb{S}(k_1,\ldots,k_n)}\;| \textrm{ $x_i=}%\textcolor{red}{e^0}$ if and only if $i\notin J$}\right\}.$$
}%\textcolor{red}{It is well known that the lower bound in Proposition~\ref{ulbTCn} is optimal for $\mathbb{S}(k_1,\ldots,k_n)$; Theorem~\ref{TC_s(X)} below asserts that the same phenomenon }%\textcolor{red}{holds} for subcomplexes. Note that, while $\mathbb{S}(k_1,\ldots,k_n)$ can be thought of as the configuration space of a mechanical robot arm whose $i$-th node moves freely in $k_i$ dimensions, a subcomplex $X$ of $\mathbb{S}(k_1,\ldots,k_n)$ encodes the information of the configuration space that results by imposing restrictions on the possible combinations of simultaneously moving nodes of the robot arm.}
\begin{theo}\label{TC_s(X)}
}%\textcolor{red}{A subcomplex $X\hspace{-.2mm}$ of $\hspace{.7mm}\mathbb{S}(k_1,\ldots,k_n)$ has} $\protect\operatorname{TC}_s(X)=\protect\operatorname{zcl}_{s}(H^*(X;}%\textcolor{red}{\mathbb{Q}}))$.
\end{theo}
Our methods imply that Theorem~\ref{TC_s(X)} could equally be stated using cohomology with coefficients in any ring of characteristic $0$.
\medskip
}%\textcolor{red}{We provide an explicit description of $\protect\operatorname{zcl}_{s}(H^*(X;}%\textcolor{red}{\mathbb{Q}}))$. The answer turns out to depend exclusively on the parity of the sphere dimensions $k_i$ (and on the combinatorics of the abstract simplicial complex underlying $X$). In order to better appreciate the phenomenon, it is convenient to focus first on the case where all the $k_i$ have the same parity\footnote{}%\textcolor{red}{An earlier version of the paper, signed by the current three authors, dealt only with the case when all the $k_i$ have the same parity. The unrestricted case was worked out later by the second named author using a mild variation of the original methods. Her results are included in the current updated version of the paper.}}. The corresponding descriptions,} in Theorems~\ref{tcsX odd} and~\ref{TC_s(X) even} as well as Corollary~\ref{sergey} }%\textcolor{blue}{in the next section,} }%\textcolor{red}{generalize} those in~\cite{CohenPruid08,sergeypreprint}. }%\textcolor{red}{The unrestricted description is given in Subsection~\ref{stn4.1} (see~Theorem~\ref{TC_s(X), gene}). In either case,} the optimality of }%\textcolor{red}{the} cohomological lower bound will be a direct consequence of the fact that we actually construct an optimal motion planner. Our construction generalizes, in a high}%\textcolor{red}{l}y non-trivial way, the one given }%\textcolor{red}{first by }%\textcolor{red}{the third author} (\cite{Yu07}) for $s=2$ when $X$ is an arrangement complement, and then independently }%\textcolor{blue}{by} Cohen-Pruidze (\cite{CohenPruid08}, as corrected in~\cite{morfismos}) in a more general case.}
\medskip
}%\textcolor{red}{By Hattori's work~\cite{hattori}, complements of generic complex hyperplane arrangements are up-to-homotopy examples of the spaces dealt with in Theorem~\ref{TC_s(X)} (with $k_i=1$ for all $i$). Those spaces are known to be formal, so their {\it rational} higher topological complexity has been shown in~\cite{carrasquel} to agree with the cohomological lower bound. Of course, such an observation can be recovered from Theorem~\ref{TC_s(X)} in view of the general fact that the rational topological complexity bounds from below the regular one. In any case it is to be noted that the rational higher TC agrees with the regular one for complements of generic complex hyperplane arrangements. Furthermore, these observations apply also for complements of the ``redundant'' arrangements considered in~\cite{CCX}, as well as for Eilenberg-Mac Lane spaces of }%\textcolor{red}{all Artin type groups for finite groups generated by reflections, see \cite{sergeypreprint}}. In this direction, it is interesting to highlight that the agreement noted above between the rational higher TC and the usual one does not hold for other formal spaces. For instance, Lucile Vandembroucq has brought to the author's attention the fact that the rational $\protect\operatorname{TC}_2$ of the symplectic group Sp$(2)$ is 2, one lower than its regular topological complexity.}
\medskip
}%\textcolor{red}{The bounds }%\textcolor{red}{in Proposition~\ref{ulbTCn}} for the higher topological complexity of a space easily yield Theorem~\ref{TC_s(X)} when all the $k_i$ agree with a fixed even number. If all the $k_i$ are even (but not necessarily equal), the result can still be proved with relative ease using the fact that the sectional category of a fibration is bounded from above by the cone-length of its base~(c.f.~\cite{LuMa}). }%\textcolor{red}{This} idea will be used in Section~\ref{othersec} in order to analyze the higher topological complexity of other polyhedral product spaces. But insisting on obtaining the required upper bound from the construction of explicit optimal motion planners (as we do) imposes a mayor task which, ironically, is much more elaborate when all the $k_i$'s are even. Yet, it seems to be extremely hard to give a proof of Theorem~\ref{TC_s(X)} that does not depend on the construction of an optimal motion planner if }%\textcolor{red}{at least one of the} $k_i$'s }%\textcolor{red}{is} odd.}
\section{}%\textcolor{red}{Optimal motion planners}}\label{secciondosmeramera}
}%\textcolor{red}{In this section we construct optimal motion planners for a subcomplex $X$ of $\mathbb{S}(k_1,\ldots,k_n)$} when all the }%\textcolor{red}{$k_i$'s} have the same parity. }%\textcolor{red}{We start by setting up some basic notation.}
\medskip
We }%\textcolor{red}{think of an element} $(b_1, b_2, \ldots, b_s) \in X^s$, with $b_j=(b_{1j}, \ldots, b_{nj}) \in X \subseteq }%\textcolor{red}{\mathbb{S}(k_1,\ldots,k_n)}$, as a matrix of size $n \times s$ whose }%\textcolor{red}{entry} $b_{ij}$ belongs to $}%\textcolor{red}{S^{k_i}}$ for all }%\textcolor{red}{$(i,j)\in [n]\times[s]$. (Here and below, for a positive integer $m$, $[m]$ stands for the initial integer interval $\{1,2,\ldots,m\}$, while $[m]_0$ stands for $[m]\cup\{0\}$).} Let $$\mathcal{P}=\{(P_1, \ldots, P_n)\,\, |\,\, P_i \,\, \mbox{is a partition of}\,\, }%\textcolor{red}{[s]} \,\, \mbox{for each}\,\, }%\textcolor{red}{i\in[n]}\,\}$$ be the set of $n$-}%\textcolor{blue}{tuples} of partitions of the interval }%\textcolor{red}{$[s]$. We assume that} elements $(P_1, \ldots, P_n) \in \mathcal{P}$ are ``ordered'' }%\textcolor{red}{in the sense that, if} $P_i=\{\alpha_1^i, \ldots, \alpha_{n(P_i)}^i\}$, }%\textcolor{red}{then} $L(\alpha_k^i) < L(\alpha_{k+1}^i)$ for $k \in }%\textcolor{red}{[n(P_i)-1]}$ where $L(\alpha_k^i)$ }%\textcolor{red}{is defined as the smallest element of the set} $\alpha_{k}^i$. }%\textcolor{red}{In particular $1\in\alpha^i_1$.} }%\textcolor{red}{The norm of} each such $P=(P_1, \ldots, P_n) \in \mathcal{P}$ }%\textcolor{red}{is defined as}
\begin{equation}\label{lanormadeP}
|P|:=\sum_{i=1}^{n}(n(P_i)-1)=\sum_{i=1}^{n} |P_i|-n,
\end{equation}
the sum of all cardinalities of the partitions $P_i$ minus $n$. }%\textcolor{red}{We let}
$$
X^s_{P}=\Big\{(b_1, b_2, \ldots, b_s) \in X^s\,\, \Big| \!\!\!\!\!\begin{array}{ll} & \mbox{for each } i \in }%\textcolor{red}{[n]}, \;b_{ik}=\pm b_{i\ell} \mbox{ if and only }%\textcolor{blue}{if}} \\ & \mbox{}%\textcolor{blue}{both} $k$ }%\textcolor{blue}{and} $\ell$ belong to the same part of $P_i$}
\end{array} \Big\},
$$
}%\textcolor{red}{and say that an element $(b_1,b_2,\ldots,b_n)\in X^s_P$ has type $P$. Note that, if} $G:= \mathbb{Z}_2= \{1, -1\}$ }%\textcolor{red}{acts antipodally} on each sphere }%\textcolor{red}{$S^k$ and, for $x \in S^k$, $G\cdot x$ stands for the $G$-orbit of $x$, then}
\begin{equation}\label{Pcardinality}
|P_i|= \left|\{G\cdot b_{ij} \,\,|\,\, }%\textcolor{red}{j\in[s]}\}\right|
\end{equation}
}%\textcolor{red}{for $(b_1, \ldots, b_s) \in X^s_{P}$ and $i \in[n]$. In addition, we consider $n$-tuples} $\beta=(\beta^1, \ldots, \beta^n)$ }%\textcolor{red}{of (possibly empty) subsets} $\beta^i \subseteq \alpha_1^i- \{1\}$ for $i \in}%\textcolor{red}{[n]}$, }%\textcolor{red}{and set}
$$X^s_{P, \beta}= X^s_{P} \cap \left\{(b_1, }%\textcolor{red}{b_2,}\ldots, b_s) \in X^s \,\,|\,\, b_{i1}= b_{ik} }%\textcolor{red}{{}\Leftrightarrow{}} k \in \beta^i, \, }%\textcolor{red}{\forall\, (i,k)\in[n]\times \left([s]-\{1\}\right)}\right\}.$$
}%\textcolor{red}{Note that the disjoint union decomposition}
\begin{equation}\label{X^s_P}
X^s_{P}=\bigsqcup_{\beta} X^s_{P, \beta},
\end{equation}
}%\textcolor{red}{running over all $n$-tuples $\beta=(\beta^1, \ldots, \beta^n)$ as above, is }%\textcolor{red}{topological,} }%\textcolor{blue}{that is, the subspace topology in $X^s_P$ agrees with the so called \emph{disjoint union topology} determined by the subspaces $X^s_{P,\beta}$. In other words, a subset $U\subseteq X^s_P$ is open if and only if each of its pieces $U\cap X^s_{P,\beta}$ (for $\beta$ as above) is open in $X^s_{P,\beta}$. Needless to say, the relevance of this property comes from the fact that the continuity of a local rule on $X^s_P$ is equivalent to the continuity of the restriction of the local rule to each $X^s_{P,\beta}$.}}
\subsection{Odd case}\label{casoimp}
}%\textcolor{red}{Throughout this subsection we assume that all $k_i$ are odd. We start by recalling an optimal motion planner for the sphere $\mathbb{S}(2d+1)=S^{2d+1}\hspace{.3mm}$---for which $\protect\operatorname{TC}_s(\mathbb{S}(2d+1))=s-1$ }%\textcolor{red}{a}s well known.}
\begin{ejem}\label{TCs S}{\em
}%\textcolor{red}{Local domains for $\mathbb{S}(2d+1)$ in the case $s=2$ are given by}
$$
A_0=\left\{(x, }%\textcolor{red}{-x}) \in }%\textcolor{red}{\mathbb{S}(2d+1)\times\mathbb{S}(2d+1)}\right\} \mbox{ \ \ and \ \ }A_1=\left\{(x, y) \in}%\textcolor{red}{\mathbb{S}(2d+1)\times\mathbb{S}(2d+1)} \,|\, x \neq -y\right\}
$$
}%\textcolor{red}{with corresponding} local rules $\phi_i$ $(i=0,1)$ described as follows: }%\textcolor{red}{For} $(x, -x)\in A_{0}$, $\phi_0 (x, }%\textcolor{red}{-x})$ }%\textcolor{red}{is} the }%\textcolor{red}{path} at constant speed from $x$ to $-x$ along the semicircle determined by $\nu(x)$, }%\textcolor{red}{where $\nu$ is some fixed }%\textcolor{red}{non-zero} tangent vector field of $\mathbb{S}(2d+1)$.} For $(x,y)\in A_1$, $\phi_1(x,y)$ }%\textcolor{red}{is} the path at constant speed along the geodesic arc connecting $x$ with $y$. }%\textcolor{red}{To deal with the case $s>2$, we consider the domains $B_j$, $j \in[s-1]_0$, consisting of $s$-tuples} $(x_1, \ldots, x_s) \in}%\textcolor{red}{\mathbb{S}(2d+1)^s}$ }%\textcolor{red}{for which} $$\left\{\,k \in \{2, \ldots, s\}\,\,|\,\,x_1\neq -x_k\,\right\}$$ has cardinality $j$, }%\textcolor{red}{with local rules $\psi_j: B_j \rightarrow \mathbb{S}(2d+1)^{J_s}$ given by} $$\psi_j((x_1, \ldots, x_s))=\left(\psi_{j1}(x_1, x_1), \ldots, \psi_{js}(x_1, x_{s})\right)$$ }%\textcolor{red}{where} $\psi_{ji}(x_1, x_{i})= \phi_{}%\textcolor{red}{r}}(x_1, x_{i})$ }%\textcolor{red}{if} $(x_1, x_i)\in}%\textcolor{red}{A_r}$. }%\textcolor{red}{As shown in~\cite[Section~4]{Ru10}, the family $\{(B_j,\psi_j)\}$ is an optimal (higher) motion planner for $\mathbb{S}(2d+1)$.}
}\end{ejem}
A well known chess-board combination of the domains $B_j$ in Example~\ref{TCs S} yield domains for an optimal motion planner for the product $\mathbb{S}(k_1,\ldots,k_n)$ (see for instance the proof of Proposition~22 in page~84 of~\cite{Schwarz66}). }%\textcolor{red}{But the situation for an arbitrary subcomplex $X}%\textcolor{blue}{{}\subseteq{}}\mathbb{S}(k_1,\ldots,k_n)$ is much more subtle.}
}%\textcolor{red}{Actually, as} it will be clear from }%\textcolor{blue}{the discussion} below, $\protect\operatorname{TC}_s(X)$ is determined by }%\textcolor{blue}{the} combinatorics of $X$ which we define }%\textcolor{blue}{next}.
\medskip
First, for a given integer $s>1$, }%\textcolor{blue}{the $s$-norm of a finite (abstract) simplicial complex $\mathcal{K}$ is the integer invariant}
$$
\protect\operatorname{N}^{}%\textcolor{red}{s}}(\mathcal{K}):=\displaystyle{\protect\operatorname{max} \left\{ \, }%\textcolor{red}{\protect\operatorname{N}_{}%\textcolor{blue}{\mathcal{K}}}(J_1,J_2,\ldots,J_s)}
\,\,|\,\, J_j \, \mbox{is a
simplex of } \mathcal{K}\,\, \text{for all}\,\, j \in }%\textcolor{red}{[s]} \right\}},
$$
}%\textcolor{red}{where}
\begin{equation}\label{nx}
\displaystyle{\protect\operatorname{N}_{\mathcal{K}}(J_1, J_2, \ldots, J_s): = \sum_{\ell=2}^s \left(\Big| \bigcap_{m=1}^{\ell-1} J_m- J_{\ell} \Big|+ \Big| J_{\ell} \Big| \right)}.
\end{equation}
}%\textcolor{red}{Now we notice some properties of the above formulas and give a simpler more symmetric definition of $\displaystyle{\protect\operatorname{N}_{\mathcal{K}}}$. }%\textcolor{blue}{Start by observing that $\protect\operatorname{N}_{\mathcal{K}}(J_1, J_2, \ldots, J_s)\leq\protect\operatorname{N}_{\mathcal{K}}(J'_1, J'_2, \ldots, J'_s)$ provided $J_i\subseteq J'_i$ for $i\in[s]$. Consequently}
$$
}%\textcolor{blue}{\protect\operatorname{N}^{}%\textcolor{red}{s}}(\mathcal{K})=\displaystyle{\protect\operatorname{max} \left\{ \, }%\textcolor{red}{\protect\operatorname{N}_{}%\textcolor{blue}{\mathcal{K}}}(J_1,J_2,\ldots,J_s)}
\,\, |\,\, J_j \, \mbox{is a {\it maximal}
\hspace{.2mm}simplex of } \mathcal{K}\,\, \text{for all}\,\, j \in }%\textcolor{red}{[s]} \right\}},}
$$
}%\textcolor{blue}{a formula that is well suited for the computation of $\protect\operatorname{N}^s(\mathcal{K})$ in concrete cases.} Also let us put $I_{\ell}=\bigcap_{m=1}^{\ell-1} J_m- J_{\ell}$ for $\ell=2,3,\ldots,s$.} }%\textcolor{blue}{Since $\bigcup_{\ell=2}^s I_{\ell}\subseteq J_1$ with $I_m\cap I_{m'}=\emptyset$ for every $m\not=m'$, we have:}
\begin{lema}\label{properties}
}%\textcolor{blue}{For (not necessarily maximal) simplexes $J_1,J_2,\ldots,J_s$ of $\mathcal{K}$,}
$$}%\textcolor{red}{\displaystyle{\protect\operatorname{N}_{\mathcal{K}}}(J_1, J_2, \ldots, J_s) = \sum_{\ell=2}^s\Big| I_{\ell}\Big|+ \sum_{\ell=2}^s\Big| J_{\ell}\Big|\leq \sum_{\ell=1}^s \Big| J_{\ell}\Big|.}$$
\end{lema}
\begin{proposition}\label{difdef}
}%\textcolor{blue}{For $J_1,J_2,\ldots,J_s$ as above}
\begin{equation}\label{sy}
}%\textcolor{red}{\displaystyle{\protect\operatorname{N}_{\mathcal{K}}}(J_1, J_2, \ldots, J_s)= \sum_{\ell=1}^s| J_{\ell}|-\Big|\bigcap_{\ell=1}^s J_{\ell}
\Big|.}
\end{equation}
\end{proposition}
\begin{proof}
}%\textcolor{red}{Due to Lemma~\ref{properties} it suffices to }%\textcolor{blue}{prove} the equality
$$\bigcup_{\ell=2}^sI_{\ell}=J_1-\bigcap_{\ell=1}^sJ_{\ell}.$$
}%\textcolor{blue}{An element $x$ on} the left hand side (LHS) }%\textcolor{blue}{satisfies} $x\in I_{\ell}$ for some $\ell\geq 2$ whence $x\not\in J_{\ell}$. Thus $x$ }%\textcolor{blue}{lies on} the right hand side (RHS). Conversely, }%\textcolor{blue}{for an element $x$ on} the RHS chose the smallest $\ell\geq 2$ such that $x\not\in J_{\ell}$. By the choice of $\ell$
and definition of $I_{\ell}$ we have $x\in I_{\ell}$ whence $x$ }%\textcolor{blue}{lies on} LHS.}
\end{proof}
\begin{corollary}
}%\textcolor{red}{$\displaystyle{\protect\operatorname{N}_{\mathcal{K}}}(J_1, J_2, \ldots, J_s)$ does not depend on the ordering of the set of simplexes.}
\end{corollary}
}%\textcolor{red}{Now we apply the combinatorics we have developed to }%\textcolor{blue}{a} CW subcomplex $X}%\textcolor{blue}{{}\subseteq{}} \mathbb{S}(k_1,\ldots,k_n)$.}
\begin{definition}
}%\textcolor{blue}{The index of $X$ is the (abstract)}
}%\textcolor{red}{simplicial complex}
$$}%\textcolor{red}{\mathcal{K}_X=\{J\subseteq[n] \;\;|\;\; e_J \mbox{ is a cell of $X$}\}.}$$
}%\textcolor{blue}{For $d\in [n]$, we say that $X$ is $d$-pure (or simply pure, if $d$ is implicit) if its index is $d$-pure in the sense that all maximal simplexes of $\mathcal{K}_X$ have cardinality $d$.}
\end{definition}
\begin{remark}\label{gmac} }%\textcolor{red}{\em Using the terminology from \cite{BahBende10}, $X$ is the }%\textcolor{red}{polyhed}%\textcolor{blue}{ral} product space} determined by the set of pairs $}%\textcolor{blue}{\{}(S^{k_1},e),\ldots,(S^{k_n},e)}%\textcolor{blue}{\}}$ and $\mathcal{K}_X$.}
\end{remark}
We use the notation $\displaystyle{\protect\operatorname{N}_{X}}(J_1, J_2, \ldots, J_s)$ and $\protect\operatorname{N}^{}%\textcolor{red}{s}}(X)$ for $\displaystyle{\protect\operatorname{N}_{\mathcal{K}_{}%\textcolor{blue}{X}}}}(J_1, J_2, \ldots, J_s)$ and $\displaystyle{\protect\operatorname{N}^{}%\textcolor{red}{s}}(\mathcal{K}_{}%\textcolor{blue}{X}}})$ respectively.
\smallskip}%\textcolor{red}{Now we state one of }%\textcolor{blue}{the} main results of the paper.}
\begin{theo}\label{tcsX odd}
}%\textcolor{red}{Assume all of the $k_i$ are odd. A subcomplex $X$ of the minimal CW cell structure on $\mathbb{S}(k_1,\cdots,k_n)$ has} $$\protect\operatorname{TC}_s(X)=\protect\operatorname{N}^{}%\textcolor{red}{s}}(X).$$
\end{theo}
}%\textcolor{red}{The proof of Theorem~\ref{tcsX odd} is deferred to the next sections; here we analyze its consequences and} interesting special instances, }%\textcolor{red}{starting with the case} when }%\textcolor{blue}{$X$} }%\textcolor{red}{is pure.}
\begin{corollary}\label{purebis}
}%\textcolor{red}{Suppose all of the $k_i$ are odd and $}%\textcolor{blue}{X}$ is $}%\textcolor{blue}{d}$-pure. Then}
}%\textcolor{red}{$$\protect\operatorname{TC}_s(X)=sd-\min\Big|\bigcap_{i=1}^sJ_i\Big|$$
where the minimum is taken over all sets $\{J_1,\ldots,J_s\}$ of maximal simplexes of ${\mathcal{K}_{}%\textcolor{blue}{X}}}$.
In particular $\protect\operatorname{TC}_s(X)\leq sd$ }%\textcolor{blue}{with} equality if and only if $\hspace{.4mm}\bigcap_{i=1}^sJ_i$ }%\textcolor{blue}{is empty}
for some choice of }%\textcolor{blue}{maximal simplexes} $J_i$}%\textcolor{blue}{'s}.}
\end{corollary}
}%\textcolor{red}{Corollary~\ref{purebis} implies that, for $X$ }%\textcolor{magenta}{$d$-pure,} }%\textcolor{blue}{$\protect\operatorname{TC}_s(X)$ growths linearly on $s$ provided $s$ is large enough. More }%\textcolor{magenta}{precisely}, if} }%\textcolor{red}{$w=w({\mathcal{K}_{}%\textcolor{blue}{X}}})$ denotes the number of maximal simplexes }%\textcolor{blue}{in $\mathcal{K}_{X}$, then}
\begin{equation}\label{asintotico}
\protect\operatorname{TC}_s(X)=d(s-w)+\protect\operatorname{TC}_w(X)
\end{equation}
for $s\geq w$. More generally we have:}}
\begin{proposition}\label{crecilineagene}
}%\textcolor{red}{Let $w$ be as above, }%\textcolor{red}{and set $d=1+\dim(\mathcal{K}_X)$.} }%\textcolor{magenta}{Equation~$(\ref{asintotico})$ holds for any (pure or not) subcomplex $X$ of $\,\mathbb{S}(k_1,\ldots,k_n)$ as long as $s\geq w$.}}
\end{proposition}
}%\textcolor{magenta}{The proof of Proposition~\ref{crecilineagene} uses the following auxiliary result:}
\begin{lema}\label{auxcrecilineagene}
}%\textcolor{magenta}{In the setting of Proposition~$\ref{crecilineagene}$, if $J_1,\ldots,J_{w}$ are simplexes of $\mathcal{K}_X$ such that $\protect\operatorname{TC}_{w}(X)=\sum_{i=1}^{w}|J_i|-\Big|\bigcap_{i=1}^{w}J_i\Big|$, then
\protect\operatorname{max}\{\,|J_i|\;|\;i\in[w]\}=}%\textcolor{red}{d}.
}\end{lema}
\begin{proof}
}%\textcolor{magenta}{}%\textcolor{red}{Assume} for a contradiction that $J_1,\ldots,J_{w}$ are simplexes of $\mathcal{K}_X$ such that $\protect\operatorname{TC}_{w}(X)=\sum_{i=1}^{w}|J_i|-\left|\bigcap_{i=1}^{w}J_i\right|$ with $|J_i|<d$ for all $i\in[w]$. Choose a simplex $J_0$ of $\mathcal{K}_X$ with $|J_0|=d$, and indexes $i_1,i_2\in[w]$, $i_1<i_2$, with $J_{i_1}=J_{i_2}$. Set
$$
(J'_1,\ldots,J'_w):=(J_0,J_1,\ldots,J_{i_1-1},J_{i_1+1},\ldots,J_w).
$$
The contradiction comes from
$$
\protect\operatorname{N}_X(J'_1,\ldots,J'_w)=\sum_{i=1}^w
|J'_i|-\Big|\bigcap_{i=1}^wJ'_i\Big|>\sum_{i=1}^w
|J_i|-\Big|\bigcap_{i=1}^wJ'_i\Big|
\geq\sum_{i=1}^w
|J_i|-\Big|\bigcap_{i=1}^wJ_i\Big|=\protect\operatorname{TC}_w(X)
$$
where the last inequality holds because $\bigcap_{i=1}^wJ'_i\subseteq\bigcap_{i=2}^wJ'_i=\bigcap_{i=1}^wJ_i$.}
\end{proof}
\begin{proof}[}%\textcolor{red}{Proof of Proposition~$\ref{crecilineagene}$}] }%\textcolor{red}{}%\textcolor{magenta}{Let $s\geq w$.} Choose }%\textcolor{red}{maximal} simplexes $J'_1,\ldots,J'_s$ }%\textcolor{magenta}{and $J_1,\ldots,J_w$} of $\mathcal{K}_X$ with $$\protect\operatorname{N}^s(X)=\sum_{i=1}^s|J'_i|-\Big|\bigcap_{i=1}^sJ'_i\Big| \;\;\;\;}%\textcolor{magenta}{\mbox{and}\;\;\;\,\protect\operatorname{N}^w(X)=\sum_{i=1}^w|J_i|-\Big|\bigcap_{i=1}^wJ_i\Big|.}$$ Assume without loss of generality (since $s\geq w$) that $\{J'_1,\ldots,J'_s\}=\{J'_1,\ldots,J'_w\}$. Then
\begin{eqnarray*}
\protect\operatorname{TC}_s(X)&=&\sum_{i=1}^{s}|J'_i|\,-\,\Big|\bigcap_{i=1}^{s}J'_i\Big|\;\;=\;\;\sum_{i=1}^{w}|J'_i|\,+\sum_{i=w+1}^{s}|J'_i|\,-\,\Big|\bigcap_{i=1}^{w}J'_i\Big|\\
&\leq&\protect\operatorname{TC}_w(X)\,+\,\sum_{i=w+1}^{s}|J'_i|\;\;\leq\;\;\protect\operatorname{TC}_w(X)\,+\,(s-w)d
\end{eqnarray*}
where, }%\textcolor{red}{as before,} $d=1+\dim(\mathcal{K}_{}%\textcolor{red}{X}})$. On the other hand, }%\textcolor{magenta}{Lemma~\ref{auxcrecilineagene}} yields an integer $i_0\in[w]$ with $|J_{i_0}|=d$. Set $J_j:=J_{i_0}$ for $w+1\leq j\leq s$. Then
$$
\protect\operatorname{TC}_w(X)+(s-w)d=\sum_{i=1}^{w}|J_i|\,-\,\Big|\bigcap_{i=1}^{w}J_i\Big|\,+\sum_{i=w+1}^{s}|J_i|=\sum_{i=1}^{s}|J_i|\,-\,\Big|\bigcap_{i=1}^{s}J_i\Big|
\leq\protect\operatorname{TC}_s(X),
$$
completing the proof.}
\end{proof}
}%\textcolor{magenta}{A more precise description of $\protect\operatorname{TC}_s(X)$ can be obtained by imposing condition}%\textcolor{blue}{s} on $X$ which are stronger than purity.} }%\textcolor{red}{}%\textcolor{blue}{For instance, let $\mathbb{S}(k_1, \ldots, k_n)^{(d)}$ stand for the $d$-pure subcomplex of $\mathbb{S}(k_1, \ldots, k_n)$ with index} $\Delta[n-1]^{d-1}$, the $(d-1)$-skeleton of the full simplicial complex on $n$ vertices.} For instance, when $k_i=1$ for all $i\in[n]$, $\mathbb{S}(k_1, \ldots, k_n)^{(d)}$ is the $d$-dimensional skeleton in the minimal CW structure of the $n$-torus---the $n$-fold Cartesian product of $S^1$ with itself.
\begin{corollary}\label{sergey}
}%\textcolor{red}{If all of the $k_i$ are odd,} then $\protect\operatorname{TC}_s\left(\mathbb{S}(k_1, \ldots, k_n)^{}%\textcolor{red}{(d)}}\right)=\min \{sd, (s-1)n\}$.
\end{corollary}
In view of Hattori's theorem (\cite{hattori}, see also \cite[Theorem 5.21]{OT}), Corollary~\ref{sergey} specializes, with $k_i=1$ for all $i\in[n]$, to the assertion in~\cite[page~8]{sergeypreprint} describing the higher topological complexity of complements of complex hyperplane arrangements that are either linear generic, or affine in general position (cf.~\cite[Section~3]{Yu07}). It is also interesting to highlight that the ``min'' part in Corollary~\ref{sergey} (with $d=1$) can be thought of as a manifestation of the fact that, while the $s$-th topological complexity of an odd sphere is $s-1$, wedges of at least two spheres have $\protect\operatorname{TC}_s=s$ ---just as any other nilpotent suspension space which is neither contractible nor homotopy equivalent to an odd sphere (\cite{minTC}). In addition, the ``min'' part in Corollary~\ref{sergey} detects a phenomenon not seen in terms of the Lusternik-Schnirelmann category since, as }%\textcolor{red}{indicated in Remark~\ref{downtocat} at the end of the paper, $\protect\operatorname{cat}(\mathbb{S}(k_1, \ldots, k_n)^{(d)})=d$.}
\begin{proof}[Proof of Corollary~$\ref{sergey}$]
}%\textcolor{blue}{Let $X$ stand for $\mathbb{S}(k_1,\ldots,k_n)^{}%\textcolor{red}{(d)}}$. For simplexes $J_1,\ldots,J_s$ of $\Delta[n-1]^{d-1}$, the} }%\textcolor{red}{inequality $\protect\operatorname{N}_X(J_1, \ldots, J_s)\leq \min\{sd,(s-1)n\}$ follows from
Corollary~\ref{purebis} and Lemma \ref{properties} since $| I_{\ell}|+| J_{\ell}|\leq n$. }%\textcolor{blue}{Thus $\protect\operatorname{TC}_s}%\textcolor{blue}{(X)})\leq\min\{sd,(s-1)n\}$} }%\textcolor{red}{(notice }%\textcolor{blue}{this holds for any $d$-pure} $X$).} }%\textcolor{red}{To prove the opposite inequality suppose }%\textcolor{blue}{first} that $sd\leq (s-1)n$, equivalently $n\leq s(n-d)$. Then
there exist a covering $\{C_1\ldots,C_s\}$ }%\textcolor{blue}{of $[n]$} with $| C_k|=n-d$ for every $k\in[s]$. Put $J_k=[n]-C_k$ and
notice that $J_k$ is a maximal simplex of $}%\textcolor{blue}{\Delta[n-1]^{d-1}}$ for every $k$. }%\textcolor{blue}{Further} $\bigcap_{k=1}^sJ_k=\emptyset$, }%\textcolor{blue}{so that Corollary~\ref{purebis} yields} $$\protect\operatorname{TC}_s}%\textcolor{blue}{(X)}=sd=}%\textcolor{blue}{\min\{sd,(s-1)n}\}.$$
Finally assume that $(s-1)n\leq sd$, i.e., $s(n-d)\leq n$. Then there exists a collection $\{C_1\ldots,C_s\}$ of mutually disjoint }%\textcolor{blue}{subsets of $[n]$} with $| C_k|=n-d$ for every $k$. Put again $J_k=[n]-C_k$. We have
$$
\protect\operatorname{TC}_s}%\textcolor{blue}{(X)}\geq\sum_{}%\textcolor{blue}{k=1}}^{}%\textcolor{blue}{s}}|J_k|-\Big|\bigcap_{}%\textcolor{blue}{k=1}}^{}%\textcolor{blue}{s}}J_k\Big| =
sn-\sum_{}%\textcolor{blue}{k=1}}^{}%\textcolor{blue}{s}}|C_k|-\Big|\bigcap_{}%\textcolor{blue}{k=1}}^{}%\textcolor{blue}{s}}J_k\Big|=sn-\sum_{}%\textcolor{blue}{k=1}}^{}%\textcolor{blue}{s}}|C_k|-n+\Big|\bigcup_{}%\textcolor{blue}{k=1}}^{}%\textcolor{blue}{s}}C_k\Big|.
$$
}%\textcolor{blue}{The result follows since the latter term simplifies to $(s-1)n=\min\{sd,(s-1)n\}$.}}}
\end{proof}
}%\textcolor{magenta}{}%\textcolor{red}{The higher topological complexity of a subcomplex $X$ of $\mathbb{S}(k_1,\ldots,k_n)$ whose index is }%\textcolor{red}{pure but} not a skeleton depends heavily on the combinatorics of $\mathcal{K}_X$ ---and not just on its dimension. To illustrate the situation, we offer the following example.}}
\begin{ejem}{\em
}%\textcolor{red}{Suppose the parameters are $n=4,\ d=2, \ s=3$; $\mathcal{K}_1$ has the set of maximal simplexes $\{\{1,2\},\{2,3\},\{3,4\}\}$ while $\mathcal{K}_2$ the set $\{\{1,2\},\{1,}%\textcolor{blue}{3}\},\{1,4\}\}$. }%\textcolor{blue}{Fix positive odd integers $k_1,k_2,k_3,k_4$, and let} $X_i$ ($i=1,2$) be }%\textcolor{blue}{the} CW }%\textcolor{blue}{sub}complex of $}%\textcolor{blue}{\mathbb{S}}(k_1,k_2,k_3,k_4)$ having $\mathcal{K}_i$ as its index. Then }%\textcolor{blue}{Corollary~\ref{purebis}} gives $\protect\operatorname{TC}_3(X_1)=6$
while $\protect\operatorname{TC}_3(X_2)=5$.}}
\end{ejem}
}%\textcolor{magenta}{Interesting phenomena can arise if $X$ is not pure.} }%\textcolor{red}{This can be demonstrated by the following example}%\textcolor{red}{s}:}
\begin{ejem}{\em }%\textcolor{red}{
Take $s=n$. For $i\in [n]$, let $K_i=[n]-\{i\}$, and for $I \subseteq [n]$ , let $$W_I=\mathbb{S}(k_1, \ldots, k_n)^{(n-1)}-\bigcup_{i\in I} e_{K_i},$$ the subcomplex obtained from the fat wedge after removing the facets corresponding to vertices $i\in I$. As before, we assume that all of the $k_i$ are odd. Note that $W_I$ is $(n-1)$-pure if $|I|\leq1$, in which case Corollary~\ref{purebis} gives
\begin{equation}\label{purebutsamedim}
\protect\operatorname{TC}_{}%\textcolor{red}{n}}(W_I)=n(n-1)-|I|.
\end{equation}
But the situation is slightly subtler when $2\leq|I|<n$ because, although the corresponding $W_I$ all have the same dimension, they fail to be pure, in fact:\begin{equation}\label{nonpurebutsamedim}
\protect\operatorname{TC}_{n}(W_I)= \begin{cases}
n(n-1)-(\delta+1), & \text{if} \,\, |I|=2\delta+1 ; \\
n(n-1)-\delta, & \text{if} \,\, |I|=2\delta .
\end{cases}
\end{equation}
Note however that, by Corollary~\ref{sergey}, once all maximal simplexes have been removed from the fat wedge, we find the rather smaller value $\protect\operatorname{TC}_n(W_{[n]})=n(n-2)$, back in accordance to~(\ref{purebutsamedim}). The straightforward counting argument verifying~(\ref{nonpurebutsamedim}) is left as an exercise for the interested reader; we just hint to the fact that the set of maximal simplexes of $\mathcal{K}_{W_I}$ is} $$}%\textcolor{red}{\left\{\,K_i \;|\; i \notin I\,\right\} \cup
\left\{\,J \;|\; [n]-J \subseteq I \hspace{1.6mm}\mbox{and}\hspace{1.3mm} |J|= n-2 \,\right\}.}$$}
\end{ejem}
\begin{ejem}\label{allpossibilities}{\em
}%\textcolor{red}{Let $c_1>c_2$ be positive integers and $n=c_1+c_2$. Consider the simplicial complex $\mathcal{K}=\mathcal{K}^{c_1,c_2}$ }%\textcolor{blue}{with vertices $[n]$} determined by two disjoint maximal simplexes $K_1$ and $K_2$ with $|K_1|=c_1$ and $|K_2|=c_2$. }%\textcolor{blue}{Then, for any collection $J_1,\ldots,J_s$ of maximal simplexes of $\mathcal{K}$, where} precisely $s_1$ sets among $J_1,\ldots,J_s$ are equal to $K_1$ with }%\textcolor{blue}{$0\leq s_1\leq s$, Proposition~\ref{difdef} yields}
$$
\protect\operatorname{N}_{}%\textcolor{blue}{\mathcal{K}}}(J_1,\ldots,J_s)=
\begin{cases}
(s-1)c_2, & s_1=0;\\
}%\textcolor{red}{s_1c_1+}%\textcolor{blue}{(s-s_1)}c_2}, & }%\textcolor{red}{0<s_1<s;}\\
(s-1)c_1, & s_1=s.
\end{cases}
$$
This function of $s_1$ reaches its largest value when $s_1=s-1$ whence $\protect\operatorname{N}^{}%\textcolor{blue}{s}}({\mathcal{K}})=(s-1)c_1+c_2=sc_1-(c_1-c_2)$. }%\textcolor{blue}{The latter} formula shows that, }%\textcolor{red}{as} $c_1-c_2$ runs through the integers $1,2,\ldots,c_1-1$, $\protect\operatorname{N}^{}%\textcolor{blue}{s}}({\mathcal{K}})$ runs through $sc_1-1,sc_1-2,\dots, (s-1)c_1+1$. }%\textcolor{red}{Whence,} due to Theorem~\ref{tcsX odd}, the same is true for $\protect\operatorname{TC}_s(X)$ where $X=X_{}%\textcolor{blue}{c_1,c_2}}$ }%\textcolor{blue}{is the} subcomplex of }%\textcolor{blue}{some} $}%\textcolor{blue}{\mathbb{S}}(k_1,\ldots,k_n)$ (with all $k_i$ odd) whose index equal}%\textcolor{blue}{s} ${\mathcal{K}}$.}}
\end{ejem}
\begin{remark}{\em }%\textcolor{red}{The previous example} should be compared with the fact (proved in~\cite[Corollary~3.3]{bgrt}) that the $s$-th topological complexity of a given path connected space $X$ is bounded by $\protect\operatorname{cat}(X^{s-1})$ from below, and by $\protect\operatorname{cat}(X^s)$ from above.
}%\textcolor{red}{Example~\ref{allpossibilities} implies} that not only can both bounds be attained (with Hopf spaces in the former case, and with closed simply connected symplectic manifold in the latter) but any possibility in between }%\textcolor{red}{can occur}. Indeed, }%\textcolor{red}{as indicated in Remark~\ref{downtocat} at the end of the paper,}
$\protect\operatorname{cat}(X_{}%\textcolor{blue}{c_1,c_2}}^p)=pc_1$ for every }%\textcolor{blue}{positive integer $p$.}}
\end{remark}
\subsection{Proof of Theorem~\ref{tcsX odd}: the upper bound}\label{suboddmotion}
}%\textcolor{red}{The inequality $\protect\operatorname{N}^{}%\textcolor{red}{s}}(X)\leq\protect\operatorname{TC}_s(X)$ will be dealt with in Section~\ref{sectres} using cohomological methods; this subsection is devoted to establishing} the inequality $\protect\operatorname{TC}_s(X)\leq \protect\operatorname{N}^{}%\textcolor{red}{s}}(X)$ }%\textcolor{red}{by proving that the domains
}%\textcolor{blue}{\begin{equation}\label{unionxp}
D_j :=\bigcup X^s_{P},\quad j\in[\protect\operatorname{N}^{}%\textcolor{red}{s}}(X)]_0,
\end{equation}
}%\textcolor{red}{where the union runs over those
$P \in \mathcal{P}$ with $|P|=j$ as
defined in~(\ref{lanormadeP}),}}
give a cover of $X^s$ by pairwise disjoint ENR subspaces each of which admits a local rule---a section for $e_s$.}
\medskip
It is }%\textcolor{red}{easy} to see that }%\textcolor{red}{the $D_j$'s are pairwise disjoint. On the other hand, it follows from Proposition~\ref{interxp} below that~(\ref{unionxp}) is a topological disjoint union, so that~\cite[Proposition~IV.8.10]{Dold} and the obvious fact that each $X^s_P$ is an ENR imply the corresponding assertion for each $D_j$.}
\begin{lema}\label{coverodd}
$$
X^s = \bigcup_{j=0}^{\protect\operatorname{N}^{}%\textcolor{red}{s}}(X)} D_j.
$$
\end{lema}
\begin{proof}
}%\textcolor{red}{Let $b\in X^s$,} say $b=(b_1, \ldots, b_s)\in e_{J_1} \times e_{J_2} \times \cdots \times e_{J_s} }%\textcolor{blue}{{}\subseteq{}} X^s$, where $J_j }%\textcolor{blue}{{}\subseteq{}} }%\textcolor{red}{[n]}$ for all }%\textcolor{red}{$j\in[s]$.} }%\textcolor{red}{Recall $G=\mathbb{Z}_2$ which acts antipodally on each sphere $S^{k_i}$.} }%\textcolor{red}{Note that} $$\sum_{i=1}^n \left|\left\{G\cdot b_{ij}\,\, |\,\, }%\textcolor{red}{j\in[2]}\right\}\right|-n= \left|\left\{}%\textcolor{red}{i\in[n]} \,\,|\,\, b_{i1} \neq \pm b_{i2}\right\}\right| \leq |J_1-J_2|+ |J_2|$$ }%\textcolor{red}{where the last inequality holds} since $\{}%\textcolor{red}{i\in[n]} \,\,|\,\, b_{i1} \neq \pm b_{i2} \} \subseteq J_1 \cup J_2$. }%\textcolor{red}{More generally,}
\begin{equation}\label{interpretacion1}
\sum_{i=1}^n |\{G\cdot b_{ij}\,\, |\,\, }%\textcolor{red}{j\in[s]}\}|-n= \sum_{\ell=2}^s |\{}%\textcolor{red}{i\in[n]} \,\,|\,\, b_{i}%\textcolor{red}{t}} \neq \pm b_{i}%\textcolor{red}{\ell}}\, \text{ for all } \, 1\leq t < \ell \}|
\end{equation}
where, for each $2 \leq \ell \leq s$,
\begin{equation}\label{interpretacion2}
|\{}%\textcolor{red}{i\in[n]}\,\, |\,\, b_{i}%\textcolor{red}{t}} \neq \pm b_{i}%\textcolor{red}{\ell}}\, \,\, \text{for all}\,\, 1\leq t < \ell \}| \leq | \bigcap_{}%\textcolor{red}{t}=1}^{\ell-1} J_{}%\textcolor{red}{t}} - J_{\ell}|+ |J_{\ell}|
\end{equation}
since }%\textcolor{red}{in fact} $$\{}%\textcolor{red}{i\in[n]}\,\, |\,\, b_{i}%\textcolor{red}{t}} \neq \pm b_{i}%\textcolor{red}{\ell}} \, \text{ for all }\, 1\leq t < \ell \} \subseteq \Big(\bigcap_{}%\textcolor{red}{t}=1}^{\ell-1} J_{}%\textcolor{red}{t}} \Big) \cup J_{\ell}.$$ Therefore, }%\textcolor{red}{if $P=(P_1,\ldots,P_n)\in\mathcal{P}$ is the type of $b$, and we set $j=|P|$, then $b\in X^s_P\subseteq D_j$ where the inequality $j\leq\protect\operatorname{N}^{s}(X)$ holds in view of~(\ref{Pcardinality}), (\ref{interpretacion1}), and~(\ref{interpretacion2}).}
\end{proof}
}%\textcolor{red}{Next,} in order to }%\textcolor{red}{construct a (well defined and continuous)} local section of $e_s$ over each $D_j$, }%\textcolor{red}{$j\in[\protect\operatorname{N}^s(X)]$, we prove that~(\ref{unionxp}) is a topological disjoint union.}
\begin{proposition}\label{interxp}
}%\textcolor{red}{For any pair of elements} $P, P' \in \mathcal{P}$ with $|P|=|P'|$ and $P \neq P'$ }%\textcolor{red}{we have}
\begin{equation}\label{inter}
\overline{X^s_P} \cap X^s_{P'}= \emptyset = X^s_{P} \cap \overline{X^s_{P'}}.
\end{equation}
\end{proposition}
\begin{proof}
}%\textcolor{red}{Write} $P=(P_1, \ldots, P_n)$ and $P'=(P'_1, \ldots, P'_n)$ so that $$ \displaystyle{\sum_{i=1}^{n} |P_i|}=\displaystyle{\sum_{i=1}^{n} |P'_i|}.$$ If there exists }%\textcolor{red}{an integer $j_1\in[n]$ with} $|P_{j_1}| > |P'_{j_1}|$ }%\textcolor{red}{(or $|P_{j_1}| < |P'_{j_1}|$), then the hypothesis forces the existence of another integer $j_2\in[n]$ with} $|P_{j_2}| < |P'_{j_2}|$ }%\textcolor{red}{($|P_{j_2}| > |P'_{j_2}|$, respectively) and, in such a case~(\ref{inter}) obviously holds. Thus, without loss of generality we can assume} $|P_i|=|P'_i|$ for all }%\textcolor{red}{$i\in[n]$.} Since $P \neq P'$, there exists $k\in }%\textcolor{red}{[n]}$ such that $P_k \neq P'_k$. }%\textcolor{red}{Write} $P_k= \{\alpha_1, \ldots, \alpha_{\ell_0}\}$ and $P'_k= \{\alpha'_1, \ldots, \alpha'_{\ell_0}\}$, both ordered in the sense }%\textcolor{red}{indicated at the beginning of the section}.
\medskip}%\textcolor{red}{Assume there are integers $t\in[\ell_0]$} with $L(\alpha_t) < L(\alpha'_t)$, }%\textcolor{red}{and let $t_0$ be the first such $t$ (necessarily $t_0>1$). Then any} $(b_1, \ldots, b_s) \in X^s_{P'}$ }%\textcolor{red}{must satisfy}
$$
b_{kL(\alpha_{}%\textcolor{red}{t_0}})} = \pm b_{kj_0}
$$
for some }%\textcolor{red}{$1 \leq j_0 \leq L(\alpha'_{t_0-1})\leq L(\alpha_{t_0-1})<L(\alpha_{t_0})$, condition that is then inherited by elements in $\overline{X^s_{P'}}$. However, by definition, any} $(b_1, \ldots, b_s) \in X^s_{P}$ }%\textcolor{red}{satisfies}
$$
b_{kL(\alpha_{}%\textcolor{red}{t_0}})}\neq \pm b_{kj}
$$
for all $1 \leq j < L(\alpha_{}%\textcolor{red}{t_0}}).$ Therefore $X^s_P \cap \overline{X^s_{P'}}= \emptyset$. }%\textcolor{red}{A symmetric argument shows} $\overline{X^s_P} \cap X^s_{P'}= \emptyset$ }%\textcolor{red}{whenever there are integers $t\in[\ell_0]$ with $L(\alpha'_t) < L(\alpha_t)$. As a consequence, we can assume, without loss of generality, that} $L(\alpha_{j}) \leq L(\alpha'_{j})$ for all $j \in }%\textcolor{red}{[\ell_0]}\;$}%\textcolor{red}{{}---{this looses the symmetry, so we now have to make sure we show \emph{both} equations in~(\ref{inter}).}}
\medskip\noindent}%\textcolor{red}{{\bf Case $1$.} Assume there are integers $t\in[\ell_0]$} such that $L(\alpha_t) < L(\alpha'_t)$, and }%\textcolor{red}{let $t_0$ be the largest} }%\textcolor{red}{such $t$. We have already noticed that} $X^s_P \cap \overline{X^s_{P'}}= \emptyset$ }%\textcolor{red}{is forced}. Moreover, note that }%\textcolor{red}{either $t_0=\ell_0$ or, else,} $L(\alpha_{t_0}) < L(\alpha'_{t_0}) <L(\alpha'_{t_0+1})=L(\alpha_{t_0+1})$, }%\textcolor{red}{but in any case we have}
\begin{itemize}
\vspace{-.5mm}
\item if $(b_1, \ldots, b_s) \in X^s_{P}$, }%\textcolor{red}{then} $b_{kL(\alpha'_{t_0})}= \pm b_{kj_0}$ for some $1 \leq j_0 < L(\alpha'_{t_0})$, and
\vspace{-2mm}
\item if $(b_1, \ldots, b_s) \in X^s_{P'}$, then $b_{kL(\alpha'_{t_0})}\neq \pm b_{kj}$ for all $1 \leq j < L(\alpha'_{t_0})$.
\vspace{-2mm}
\end{itemize}
\noindent }%\textcolor{red}{Since the former condition is inherited on elements of $\overline {X^s_{P}}$, we see} $\overline{X^s_{P}}\cap X^s_{P'}= \emptyset$.
\medskip\noindent }%\textcolor{red}{{\bf Case $2$}. Assume} $L(\alpha_j) = L(\alpha'_j)$ for all }%\textcolor{red}{$j\in[\ell_0]$. (Note }%\textcolor{red}{that} the symmetry is now restored.)} }%\textcolor{red}{Since $P_k\neq P'_k$, there is an integer $j_0\in[\ell_0]$ with} $\alpha_{j_0}\neq \alpha'_{j_0}$. Without }%\textcolor{red}{loss} of generality }%\textcolor{red}{we can further assume there is an integer $m_0 \in \alpha_{j_0}-\alpha'_{j_0}$ (note $m_0 \neq L(\alpha_{j_0})$, but once again the symmetry has been destroyed). Under these conditions we have}
\begin{itemize}
\vspace{-.5mm}
\item if $(b_1, \ldots, b_s) \in X^s_P$, then $b_{kL(\alpha_{j_0})}= \pm b_{km_0}$, and
\vspace{-2mm}
\item if $(b_1, \ldots, b_s) \in X^s_{P'}$ then $b_{kL(\alpha_{j_0})}=b_{kL(\alpha'_{j_0})}\neq \pm b_{km_0} $.
\vspace{-2mm}
\end{itemize}
\noindent }%\textcolor{red}{Since the former condition is inherited on elements of $\overline {X^s_{P}}$, we see} $\overline{X^s_P} \cap X^s_{P'}= \emptyset$. Moreover, since $m_0 \notin \alpha'_{j_0}$, there }%\textcolor{red}{is $d_0\in[\ell_0]$} with $m_0 \in \alpha'_{d_0}$. }%\textcolor{red}{Necessarily $d_0\neq j_0$ and $m_0\notin\alpha_{d_0}$, so we now have}
\begin{itemize}
\vspace{-.5mm}
\item if $(b_1, \ldots, b_s) \in X^s_{P'}$, then $b_{kL(\alpha'_{d_0})}=\pm b_{km_0}$, and
\vspace{-2mm}
\item if $(b_1, \ldots, b_s)\in X_P^s$, then $b_{kL(\alpha'_{d_0})}=b_{kL(\alpha_{d_0})} \neq \pm b_{km_0}$,
\vspace{-2mm}
\end{itemize}
\noindent }%\textcolor{red}{implying} $X^s_P\cap \overline{X^s_{P'}}=\emptyset$.
\end{proof}
}%\textcolor{red}{Our only remaining task in this subsection is the construction of a local rule over $D_j$ for each $j\in[\protect\operatorname{N}^s(X)]_0$. Actually, by~(\ref{X^s_P}), (\ref{unionxp}), and Proposition~\ref{interxp}, the task can be simplified to the construction of a local rule over each $X^{s}_{P,\beta}$. To fulfill such a goal, it will be convenient to normalize each sphere $S^{k_i}$} so to have great semicircles of length $1/2$. }%\textcolor{red}{Then, for} $x, y \in }%\textcolor{red}{S^{k_i}}$, we let $d(x, y)$ stand for the length of the shortest geodesic in }%\textcolor{red}{$S^{k_i}$} between $x$ and $y$ (e.g.~$d(x,-x)=1/2$). Likewise, the local rules $\phi_0$ and $\phi_1$ for }%\textcolor{red}{each $S^{k_i}$} defined at Example~\ref{TCs S} need to be adjusted---}%\textcolor{red}{but the domains $A_i$, $i=0,1$, remain unchanged---as follows:}
}%\textcolor{red}{For} $i= 0,1$ }%\textcolor{red}{and} $(x, y) \in A_i$ we set
$$\tau_i(x, y)(t)=\begin{cases}
\phi_i(x, y)\left(\frac{1}{d(x, y)}t\right), &
0\leq }%\textcolor{green}{t < d(x, y)}; \\ y, & d(x, y)\leq t \leq 1.
\end{cases}$$
Thus, $\tau_i$ reparametrizes $\phi_i$ so to perform the motion at speed $1$, keeping still at the final position once it is reached---which happens at most at time $1/2$.
\medskip
}%\textcolor{red}{In what follows it is helpful to keep in mind that, as before, elements} $(}%\textcolor{red}{b}_1, \ldots, }%\textcolor{red}{b}_s) \in X^s$, with $}%\textcolor{red}{b}_j=(}%\textcolor{red}{b}_{1j}, \ldots, }%\textcolor{red}{b}_{nj})$ for $j\in }%\textcolor{red}{[s]}$, }%\textcolor{red}{can be thought of as matrices $(b_{i,j})$ whose columns represent the various stages in $X$ through which motion is to be planned (necessarily along rows). Actually, we follow a ``pivotal'' strategy: starting at the first column, motion spreads to all other columns---keeping still in the direction of the first column. In detail, in terms of the notation set at the beginning of the introduction for elements in the function space $X^{J_s}$,} consider the map
\begin{equation}\label{elplanmot}
}%\textcolor{red}{}%\textcolor{green}{\varphi}\colon X^s} \to \mathbb{S}}%\textcolor{red}{(k_1,\ldots, k_n)}^{J_s}
\end{equation}
}%\textcolor{red}{given by} $}%\textcolor{green}{\varphi}\left((}%\textcolor{red}{b}_1, \ldots, }%\textcolor{red}{b}_s)\right)=\left(}%\textcolor{green}{\varphi_1}(}%\textcolor{red}{b}_1, }%\textcolor{red}{b}_1), \ldots , }%\textcolor{green}{\varphi_s}(}%\textcolor{red}{b}_1, }%\textcolor{red}{b}_s)\right)$ where, }%\textcolor{red}{for $j\in[s]$,} $$}%\textcolor{green}{\varphi_j}(}%\textcolor{red}{b}_1, }%\textcolor{red}{b}_j)= \left(}%\textcolor{green}{\varphi_{1j}} (}%\textcolor{red}{b}_{11}, }%\textcolor{red}{b}_{1j}), \ldots, }%\textcolor{green}{\varphi_{nj}}(}%\textcolor{red}{b}_{n1}, }%\textcolor{red}{b}_{nj})\right)$$ }%\textcolor{red}{is the path in $\mathbb{S}(k_1,\ldots,k_n)$, from $b_1$ to $b_j$, whose $i$-th coordinate} $}%\textcolor{green}{\varphi_{ij}}(}%\textcolor{red}{b}_{i1}, }%\textcolor{red}{b}_{ij})$, }%\textcolor{red}{}%\textcolor{red}{$i\in[n]$,} is the path in $S^{k_i}$, from $b_{i1}$ to $b_{ij}$,} defined by $$}%\textcolor{green}{\varphi_{i,j}} (}%\textcolor{red}{b}_{i1},}%\textcolor{red}{b}_{ij})(t)=\begin{cases} }%\textcolor{red}{b}_{i1}, & 0 \leq t \leq t_{ }%\textcolor{red}{b}_{i1}}, \\ }%\textcolor{red}{\sigma}(}%\textcolor{red}{b}_{i1}, }%\textcolor{red}{b}_{ij})(t-t_{}%\textcolor{red}{b}_{i1}}), & t_{}%\textcolor{red}{b}_{i1}} \leq t \leq 1. \end{cases}$$ Here $t_{}%\textcolor{red}{b}_{i1}}=\frac{1}{2} -d(}%\textcolor{red}{b}_{i1}, }%\textcolor{red}{e^0})$ and
\begin{equation}\label{lasigma}
}%\textcolor{red}{\sigma(b_{i1}, b_{ij})}=\begin{cases}\tau_1}%\textcolor{red}{(b_{i1}, b_{ij})}, & (}%\textcolor{red}{b}_{i1}, }%\textcolor{red}{b}_{ij}) \in A_1; \\ \tau_0}%\textcolor{red}{(b_{i1}, b_{ij})}, & (}%\textcolor{red}{b}_{i1}, }%\textcolor{red}{b}_{ij}) \in A_0. \end{cases}
\end{equation}
}%\textcolor{red}{Fix $n$-tuples $P=(P_1, \ldots, P_{n}) \in \mathcal{P}$ and $\beta=(\beta^1, \ldots, \beta^n)$, with $P_i=\{\alpha_{1}^i, \ldots, \alpha_{n(P_i)}^i\}$ and $\beta^i \subseteq \alpha_{1}^i-\{1\}$ for all $i\in[n]$. Although }%\textcolor{green}{$\varphi$} is not continuous, its restriction $}%\textcolor{green}{\varphi}_{P, \beta}$ to $X^s_{P, \beta}$ is, for then~(\ref{lasigma}) takes the form}
$$
\sigma=\begin{cases}\tau_1, & \mbox{$j \notin \alpha^i_{1}$ or $j \in }%\textcolor{red}{\beta^i\cup\{1\}}$}; \\ \tau_0, & \mbox{$j \in \alpha^i_{1}$ and $j \notin }%\textcolor{red}{\beta^i\cup\{1\}}$}. \end{cases}
$$
}%\textcolor{red}{Since} $}%\textcolor{green}{\varphi}_{P, \beta}$ is }%\textcolor{red}{clearly} a section for the end-points evaluation map $e_s^{}%\textcolor{red}{\mathbb{S}(k_1,\ldots,k_n)}}$, we only }%\textcolor{red}{need to} check that $}%\textcolor{green}{\varphi}_{P, \beta}$ }%\textcolor{red}{actually} takes values in $X^{J_s}$, }%\textcolor{red}{i.e.~that our proposed motion planner does not leave $X$.}
\begin{remark}\label{intento}{\em
}%\textcolor{red}{An attempt to verify the analogous assertion in~\cite[proof of Proposition~3.5]{CohenPruid08} (where $s=2$), and the eventual realizing and fixing of the problems with that assertion, led to the work in~\cite{morfismos}. The verification in the current more general setting }%\textcolor{red}{(i.e.~proof of Proposition~\ref{noleave} below)} is inspired by the one carefully explained in~\cite[page~7]{morfismos}, and here we include full details for completeness.}
}\end{remark}
\begin{proposition}\label{noleave}
}%\textcolor{red}{The image of $}%\textcolor{green}{\varphi}$ is contained in $X^{}%\textcolor{green}{J_s}}$.}
\end{proposition}
\begin{proof}
}%\textcolor{magenta}{}%\textcolor{red}{Choose $(b_1,b_2,\ldots,b_s)\in X^s$ where, as above, $b_j=(b_{1j},b_{2j},\ldots,b_{nj})\in X$. We need to check that, for all $j\in[s]$, the }%\textcolor{red}{image of} $}%\textcolor{green}{\varphi}_j(b_1,b_j)\colon[0,1]\to \mathbb{S}(k_1,\ldots,k_n)$ lies inside $X$. By construction, the path} $}%\textcolor{green}{\varphi}_{j}(b_1,b_j)$ }%\textcolor{red}{runs coordinate-wise,} from $b_1$ to $b_j$, according to the }%\textcolor{red}{instructions} $\tau_k(b_{i1}, b_{ij})$ ($k=0, 1, \,\, i \in [n] $), except that, }%\textcolor{red}{in the $i$-th coordinate,} the movement is delayed a time $t_{b_{i1}}\leq1/2$. The closer $b_{i1}$ gets to $}%\textcolor{red}{e^0}$, the closer the delaying time $t_{b_{i1}}$ gets to $1/2$. It is then convenient to think of the path $}%\textcolor{green}{\varphi}_j(b_1, b_j)$ as }%\textcolor{red}{running} in two }%\textcolor{red}{sections}. In the first }%\textcolor{red}{section} ($t\leq1/2$) all initial coordinates $b_{i1}=e^0$ keep still, while the rest of the coordinates (eventually) start traveling to their corresponding final position $b_{ij}$. Further, }%\textcolor{red}{when} the second section starts ($t=1/2$), any final coordinate $b_{ij}=}%\textcolor{red}{e^0}$ will already have been reached, }%\textcolor{red}{and will keep still throughout the rest of the motion.} As a result, }%\textcolor{red}{the image of} $}%\textcolor{green}{\varphi_j(b_1,b_j)}$ }%\textcolor{red}{is forced to be contained in} $X$. In more detail, let $e(J_1,\ldots,J_s):=e_{J_1} \times e_{J_2} \times \cdots \times e_{J_s} }%\textcolor{blue}{{}\subseteq{}} X^{}%\textcolor{green}{s}}$ be }%\textcolor{red}{the} product of cells of $X$ }%\textcolor{red}{containing} $(b_1, b_2, \ldots, b_s )$. Then, coordinates corresponding to indexes }%\textcolor{red}{$i \in [n]-J_1$} keep their initial position $b_{i1}=e^0$ through time $t\leq1/2$. Therefore $}%\textcolor{green}{\varphi}_j(b_1, b_j)[0,1/2]$ stays within $}%\textcolor{red}{\overline{e_{J_1}}}\subseteq X$. On the other hand, by construction, $}%\textcolor{green}{\varphi}_{ij}(b_{i1}, b_{ij})(t)=b_{ij}=e^0$ whenever $t\geq1/2$ and }%\textcolor{red}{$i \in[n]- J_j$.} Thus, $}%\textcolor{green}{\varphi}_j(b_1, b_j)[1/2,1]$ stays within $}%\textcolor{red}{\overline{e_{J_j}}}\subseteq X$.}
\end{proof}
\subsection{Even case}\label{laspares}
We }%\textcolor{red}{now turn} our attention to the case when $X$ }%\textcolor{red}{is} a subcomplex of }%\textcolor{red}{$\mathbb{S}(k_1,\ldots,k_n)$ with all the $k_i$ even---assumption that will be in force throughout this subsection. As above,} the goal }%\textcolor{red}{is the construction of an optimal motion planner for the $s$-th topological complexity of $X$.} We }%\textcolor{red}{start with the following analogue of Example~\ref{TCs S}:}
\begin{ejem}\label{TCs S even}{\em
}%\textcolor{red}{Local domains for the sphere $\mathbb{S}(2d)=S^{2d}$ in the case $s=2$ are given by}
\begin{eqnarray*}
}%\textcolor{red}{B_0}&=&\{(e^{}%\textcolor{red}{0}}, -e^{}%\textcolor{red}{0}}), (-e^{}%\textcolor{red}{0}}, e^{}%\textcolor{red}{0}})\} }%\textcolor{red}{{}\subseteq{}} \mathbb{S}}%\textcolor{red}{(2d)} \times \mathbb{S}}%\textcolor{red}{(2d)},\\
}%\textcolor{red}{B_1}&=&\{(x, -x) \in \mathbb{S}}%\textcolor{red}{(2d)} \times \mathbb{S}}%\textcolor{red}{(2d)} \,|\, x \neq \pm e^{}%\textcolor{red}{0}}\}, \mbox{ and}\\
}%\textcolor{red}{B_2}&=&\{(x, y) \in \mathbb{S}}%\textcolor{red}{(2d)} \times \mathbb{S}}%\textcolor{red}{(2d)} \,\,|\,\, x \neq -y\}\;\;=\;\;\mathbb{S}}%\textcolor{red}{(2d)} \times \mathbb{S}}%\textcolor{red}{(2d)} - (B_0 \cup B_1),
\end{eqnarray*}
}%\textcolor{red}{with corresponding local rules $\lambda_i\colon }%\textcolor{red}{B_i}\to \mathbb{S}(2d)^{[0,1]}$ ($i=0,1,2$) described as follows:}
\begin{itemize}
\item $\lambda_0 (e^{}%\textcolor{red}{0}}, -e^{}%\textcolor{red}{0}})$ }%\textcolor{red}{and} $\lambda_0 (-e^{0}, e^{0})$ }%\textcolor{red}{are the} paths, at constant speed, from $e^{}%\textcolor{red}{0}}$ to $-e^{}%\textcolor{red}{0}}$ }%\textcolor{red}{and from $-e^{0}$ to $e^{0}$, respectively, along some fixed meridian---thinking of $e^0$ and $-e^0$ as the poles of $\mathbb{S}(2d)$.}
\vspace{-1.5mm}
\item }%\textcolor{red}{For} a }%\textcolor{red}{fixed nowhere zero} tangent vector field }%\textcolor{red}{$\upsilon$ on} $\mathbb{S}(2d)- \{}%\textcolor{red}{\pm}e^{}%\textcolor{red}{0}}\}$, $\lambda_1(x, -x)$ (with $x \neq}%\textcolor{red}{\pm}e^{}%\textcolor{red}{0}}$) }%\textcolor{red}{is} the path at constant speed from $x$ to $-x$ along the }%\textcolor{red}{great} semicircle determined by the tangent vector $\upsilon(x)$.
\vspace{-1.5mm}
\item }%\textcolor{red}{For $x\neq-y$,} $\lambda_2(x, y)$ }%\textcolor{red}{is} the path }%\textcolor{red}{from $x$ to $y$,} at constant speed, along the }%\textcolor{red}{shortest} geodesic arc }%\textcolor{red}{determined by $x$ and $y$.}
\end{itemize}
}\end{ejem}
}%\textcolor{red}{The generalization of Example~\ref{TCs S even} to }%\textcolor{red}{the} higher topological complexity }%\textcolor{red}{of a} subcomplex of a product of even }%\textcolor{red}{dimensional} spheres is slightly more }%\textcolor{red}{elaborate} than the corresponding generalization of Example~\ref{TCs S} in the previous section due, in part, to the additional local domain in Example~\ref{TCs S even}. So, before considering the general situation (Theorem~\ref{TC_s(X) even} below), and in order to illustrate the essential points in our construction, it will be convenient to give full details in the case }%\textcolor{red}{of} $\protect\operatorname{TC}_s(\mathbb{S}(2d))$.}
\smallskip
}%\textcolor{red}{Consider the sets}
\begin{eqnarray*}
T_0&=&\{(x_1, \ldots, x_s) \in }%\textcolor{red}{\mathbb{S}(2d)^s}\,\, | \,\, x_j\neq}%\textcolor{red}{\pm}e^{}%\textcolor{red}{0}},\,\, \text{for all}\,\, }%\textcolor{red}{j\in[s]}\},\\
T_1&=&\{(x_1, \ldots, x_s) \in }%\textcolor{red}{\mathbb{S}(2d)^s}\,\, | \,\, x_j=}%\textcolor{red}{\pm}e^{}%\textcolor{red}{0}},\,\, \text{for some}\,\, }%\textcolor{red}{j\in[s]}\}
\end{eqnarray*}
}%\textcolor{red}{and,} for each }%\textcolor{red}{partition $P$} of }%\textcolor{red}{$[s]$} and }%\textcolor{red}{each }$i \in \{0, 1\}$, $$}%\textcolor{red}{\mathbb{S}(2d)^{s}_{P, i}}=\left\{(x_1, \ldots, x_s) \in }%\textcolor{red}{\mathbb{S}(2d)^s}\,\, \Big| \!\!\!\!\begin{array}{ll} &x_{l}=\pm x_{k} \,\, \mbox{if and only if} \,\,k\mbox{ and }l \\ & \mbox{belong to the same }%\textcolor{red}{part} in} \,\, }%\textcolor{red}{P} \,\, \end{array} \right\} \cap T_i.$$ }%\textcolor{red}{The norm of the pair $(P,i)$ above is defined as} $\protect\operatorname{N}(}%\textcolor{red}{P}, i)= |}%\textcolor{red}{P}|-i$. }%\textcolor{red}{Lastly,} for $k \in }%\textcolor{red}{[s]_0}$, }%\textcolor{red}{consider the set}
\begin{equation}\label{lashaches}
H_k= \bigcup_{\protect\operatorname{N}(}%\textcolor{red}{P}, i)=k}}%\textcolor{red}{\mathbb{S}(2d)^s_{P, i}}.
\end{equation}
\begin{proposition}\label{queva}
}%\textcolor{red}{There is an optimal motion planner for $\hspace{.172mm}\mathbb{S}(2d)$ with local domains $H_k$, $k\in[s]_0$.}
\end{proposition}
\begin{proof}
}%\textcolor{red}{The optimality of such a motion planner follows by the well known fact the $s$-th topological complexity of an even sphere is $s$. On the other hand, it is obvious that $H_0,\ldots, H_s$ form a pairwise disjoint covering of $\mathbb{S}(2d)^s$. Since each $\mathbb{S}(2d)^s_{P,i}$ is clearly an ENR, it suffices to show that~(\ref{lashaches}) is a topological disjoint union (so $H_k$ is also an ENR), and that each $\mathbb{S}(2d)^s_{P,i}$ admits a local rule (all of which, therefore, }%\textcolor{red}{determine} a local rule on $H_k$).}
\medskip\noindent
}%\textcolor{red}{{\bf Topology of $H_k$}: For pairs} $(}%\textcolor{red}{P}, i)$ and $(}%\textcolor{red}{P}', i')$ }%\textcolor{red}{as above, with} $\protect\operatorname{N}(}%\textcolor{red}{P}, i)=\protect\operatorname{N}(}%\textcolor{red}{P'}, i')$ }%\textcolor{red}{and} $(}%\textcolor{red}{P}, i) \neq (}%\textcolor{red}{P'}, i')$, }%\textcolor{red}{we prove}
\begin{equation}\label{otravez}
}%\textcolor{red}{\overline{\mathbb{S}(2d)^s_{P,i}}\cap\mathbb{S}(2d)^s_{P',i'}=\emptyset=\mathbb{S}(2d)^s_{P,i}\cap\overline{\mathbb{S}(2d)^s_{P',i'}}.}
\end{equation}
If }%\textcolor{red}{$i\neq i'$, say} $i=1$ and $i'=0$, then }%\textcolor{red}{the first equality in~(\ref{otravez}) is obvious, whereas the second equality follows since} $|}%\textcolor{red}{P}| > |}%\textcolor{red}{P'}|$. }%\textcolor{red}{On the other hand, if} $i=i'$, }%\textcolor{red}{then} $|}%\textcolor{red}{P}|=|}%\textcolor{red}{P'}|$ with $}%\textcolor{red}{P}\neq}%\textcolor{red}{P'}$, and the }%\textcolor{red}{argument starting in the second paragraph of the proof of} Proposition \ref{interxp} gives~}%\textcolor{red}{(\ref{otravez})}.
\medskip\noindent }%\textcolor{red}{{\bf Local section on $\mathbb{S}(2d)^s_{P,i}$}:} }%\textcolor{red}{We assume the partition $P=\{\alpha_1, \ldots, \alpha_n\}$ is ordered in the sense indicated at the beginning of this section.} For each $}%\textcolor{red}{\beta} \subseteq }%\textcolor{red}{\alpha_1}-\{1\}$, let $$}%\textcolor{red}{\mathbb{S}(2d)^s_{P, i,\beta}}=}%\textcolor{red}{\mathbb{S}(2d)^s_{P, i}} \cap \{(x_1, \ldots, x_s) \in }%\textcolor{red}{\mathbb{S}(2d)^s}\, |\, x_1 =x_j}%\textcolor{red}{{}\Leftrightarrow{}} j \in }%\textcolor{red}{\beta},\;}%\textcolor{red}{\forall j\in[s]-1}\}. $$ }%\textcolor{red}{Since}
$$
}%\textcolor{red}{\mathbb{S}(2d)^s_{P, i}}=\bigsqcup_{}%\textcolor{red}{\beta} \subseteq }%\textcolor{red}{\alpha_1}-\{1\}}}%\textcolor{red}{\mathbb{S}(2d)^s_{P, i,\beta}}
$$
}%\textcolor{red}{is a topological disjoint union, it suffices to construct a local section on each $\mathbb{S}(2d)^s_{P, i,\beta}$.}
\medskip\noindent}%\textcolor{red}{{\bf Case $i=0$}.}
}%\textcolor{red}{As in the previous subsection, the required} local section }%\textcolor{red}{can be} defined }%\textcolor{red}{by the formula} $\sigma(x_1, \ldots, x_s)=(\sigma_1 (x_1, x_1), \ldots, \sigma_s(x_1, x_s))$ where
$$
}%\textcolor{red}{\sigma_j}=\begin{cases}
}%\textcolor{red}{\lambda_2}, & \mbox{if }j \in }%\textcolor{red}{\left([s]-\alpha_1\right) \cup \beta\cup\{1\}};\\
}%\textcolor{red}{\lambda_1}, & \mbox{}%\textcolor{red}{otherwise}.}
\end{cases}
$$
\medskip\noindent}%\textcolor{red}{{\bf Case $i=1$}.}
}%\textcolor{red}{The required local section is now defined in terms of the decomposition}
\begin{equation}\label{tdu2}
}%\textcolor{red}{\mathbb{S}(2d)^s_{P, i,\beta}}= \left( }%\textcolor{red}{\mathbb{S}(2d)^s_{P, i,\beta}} \cap }%\textcolor{red}{T}_0(}%\textcolor{red}{\alpha_1})\right)\sqcup \left(}%\textcolor{red}{\mathbb{S}(2d)^s_{P, i,\beta}} \cap }%\textcolor{red}{T}_1(}%\textcolor{red}{\alpha_1}) \right)
\end{equation}
}%\textcolor{red}{which will be shown in Lemma~\ref{auxiliando} below to be a topological disjoint union. Here} $$}%\textcolor{red}{T}_0(}%\textcolor{red}{\alpha_1})=\{(x_1, \ldots, x_s) \in }%\textcolor{red}{\mathbb{S}(2d)^s}\,\, | \,\, x_j\neq}%\textcolor{red}{\pm}e^{}%\textcolor{red}{0}}, \,\, \mbox{for all}\,\, j\in }%\textcolor{red}{\alpha_1} \}$$ and $$}%\textcolor{red}{T}_1(}%\textcolor{red}{\alpha_1})=\{(x_1, \ldots, x_s) \in }%\textcolor{red}{\mathbb{S}(2d)^s}\,\, | \,\, x_j=}%\textcolor{red}{\pm}e^{}%\textcolor{red}{0}} ,\,\, \text{for some}\,\, j\in }%\textcolor{red}{\alpha_1} \}.$$ A local section }%\textcolor{red}{on} $}%\textcolor{red}{\mathbb{S}(2d)^s_{P, i,\beta} \cap T_0(\alpha_1)}$ is defined just as in }%\textcolor{red}{the case $i=0$}, }%\textcolor{red}{whereas} a local section }%\textcolor{red}{on} $}%\textcolor{red}{\mathbb{S}(2d)^s_{P, i,\beta} \cap T_1(\alpha_1)}$ is defined }%\textcolor{red}{by the formula} $\mu(x_1, \ldots, x_s)=(\mu_1 (x_1, x_1), \ldots, \mu_s(x_1, x_s))$ where
$$
}%\textcolor{red}{\mu_j}=\begin{cases}
}%\textcolor{red}{\lambda_2}, &\mbox{if \ }}%\textcolor{red}{j \in \left([s]-\alpha_1\right)\cup\beta\cup\{1\};}\\
}%\textcolor{red}{\lambda_0}, & \mbox{}%\textcolor{red}{otherwise.}}
\end{cases}
$$
\vspace{-8mm}
\end{proof}
\begin{lema}\label{auxiliando}
}%\textcolor{red}{The decomposition~$(\ref{tdu2})$ is a topological disjoint union $($recall $i=1).$}
\end{lema}
\begin{proof}
}%\textcolor{red}{The condition ``$x_j=\pm e^0$ for some $j\in\alpha_1$'' in $T_1(\alpha_1)$ is inherited by elements in its closure, in particular
$$
\left( \mathbb{S}(2d)^s_{P, i,\beta} \cap T_0(\alpha_1)\right)\sqcup \overline{\left(\mathbb{S}(2d)^s_{P, i,\beta} \cap T_1(\alpha_1)\right)}=\emptyset.
$$
On the other hand, since $i=1$, the condition ``$x_j=\pm e^0$ for some $j\not\in\alpha_1$'' is forced on elements of $\mathbb{S}(2d)^s_{P, i,\beta} \cap T_0(\alpha_1)$ and, consequently, on elements of its closure. But the latter condition is not fulfilled by any element in $\mathbb{S}(2d)^s_{P, i,\beta} \cap T_1(\alpha_1)$.}
\end{proof}
}%\textcolor{red}{We now focus on the general situation.}
\begin{theo}{\label{TC_s(X) even}}
}%\textcolor{red}{Assume all of the $k_i$ are even. A subcomplex $X$ of the minimal CW structure on $\mathbb{S}(k_1,\ldots,k_n)$ has} $$\protect\operatorname{TC}_s(X)= }%\textcolor{red}{s(1+\dim(\mathcal{K}_X)).}$$
\end{theo}
}%\textcolor{red}{The inequality $s(1+\dim(\mathcal{K}_X))\leq\protect\operatorname{TC}_s(X)$ will be dealt with in Section~\ref{sectres} using cohomological methods; in the rest of this subsection we prove the inequality $\protect\operatorname{TC}_s(X)\leq s(1+\dim(\mathcal{K}_X))$ by constructing an explicit motion planner with $1+s(1+\dim(\mathcal{K}_X))$ local domains---given by the sets in~(\ref{tdu3}) below.}
\medskip
}%\textcolor{red}{As in previous constructions, we think of an element} $(b_1, \ldots, b_s) \in X^s$ }%\textcolor{red}{with} $b_{j}=(b_{1j}, \ldots, b_{nj})$, $j \in [s]$, }%\textcolor{red}{as an $n\times s$ matrix whose $(i,j)$ coordinate is $b_{ij}\in\mathbb{S}(k_i)$}. }%\textcolor{red}{For} $P \in \mathcal{P}$ and $k\in}%\textcolor{red}{[n]_0}$, }%\textcolor{red}{set} $\protect\operatorname{N}(P, k):= \sum_{i=1}^n |P_i|-k$, the norm of the pair $(P,k)$, and
$$
X^{s}_{P, k}:=X^s_P \cap \left\{(b_{1}, \ldots, b_s) \in }%\textcolor{red}{\mathbb{S}(k_1,\ldots,k_n)^s} \;\; \Big| \!\!\!\!\!\begin{array}{cl}&\,(b_{i1}, \ldots, b_{is}) }%\textcolor{red}{{}\in T_{1,k_i}} \mbox{ for}\\ &\text{ exactly $k$ indexes $i \in }%\textcolor{red}{[n]}$}\end{array} \right\}
$$
}%\textcolor{red}{where $T_{1,k_i}=\left\{(x_1, \ldots, x_s) \in }%\textcolor{red}{\mathbb{S}(k_i)^s}\,\, | \,\, x_j=}%\textcolor{red}{\pm}e^{}%\textcolor{red}{0}},\,\, \text{for some}\,\, }%\textcolor{red}{j\in[s]}\right\}$. The local domains we propose are given by}
\begin{equation}\label{tdu3}
W_r= \bigcup_{\protect\operatorname{N}(P, k)=r} X^{s}_{P, k}.
\end{equation}
}%\textcolor{red}{By~(\ref{Pcardinality}), the norm $\protect\operatorname{N}(P, k)$} }%\textcolor{red}{is the number of} ``row'' }%\textcolor{red}{$G$-}orbits }%\textcolor{red}{different from that of $e^0$ in any matrix $(b_1,\ldots,b_s)\in X^s_{P,k}$}. Therefore }%\textcolor{red}{the sets $W_r$ with $r\in[s(1+\dim(\mathcal{K}_X))]_0$ yield a pairwise disjoint cover of} $X^s$. }%\textcolor{red}{Our task then is to show:}
\begin{proposition}\label{tarea1.1}
}%\textcolor{red}{Each $W_r$ is an ENR admitting a local rule.}
\end{proposition}
Our proof of Proposition~\ref{tarea1.1} }%\textcolor{red}{depends on showing that~(\ref{tdu3}) is a topological disjoint union (Lemma~\ref{separadotes} below) and that each piece $X^s_{P,k}$ admits a suitably finer topological decomposition ((\ref{tdu4}),~(\ref{tdu10}), and Proposition~\ref{construfinal} below).}
\begin{lema}\label{separadotes}
}%\textcolor{red}{For} $P, P' \in \mathcal{P}$ and $k , k' \in }%\textcolor{red}{[n]_0}$ }%\textcolor{red}{with} $N(P, k)= N(P', k')$ and $(P, k) \neq (P', k')$, $$\overline{X^s_{}%\textcolor{red}{P,k}}}\cap X^s_{}%\textcolor{red}{P', k'}}= \emptyset = X^s_{}%\textcolor{red}{P, k}}\cap \overline{ X^s_{}%\textcolor{red}{P',k'}}} $$
\end{lema}
\begin{proof}
}%\textcolor{red}{Write} $P=(P_1, \ldots, P_{}%\textcolor{red}{n}})$ and $}%\textcolor{red}{P'}=(P_1', \ldots, P'_{}%\textcolor{red}{n}})$ }%\textcolor{red}{so that, by hypothesis,} $\sum_{i=1}^n |P_i| -k = \sum_{i=1}^n |P'_i| -k'.$ If $k > k'$, then $\overline{X^s_{}%\textcolor{red}{P,k}}}\cap X^s_{}%\textcolor{red}{P',k'}}= \emptyset$, }%\textcolor{red}{and since} $\sum_{i=1}^n |P_i| > \sum_{i=1}^n |P'_i|$ }%\textcolor{red}{is forced, we also get} $X^s_{}%\textcolor{red}{P, k}}\cap \overline{ X^s_{}%\textcolor{red}{P',k'}}} = \emptyset$. If $k=k'$, then $|P|= |P'|$ with $P \neq P'$ and, }%\textcolor{red}{just as for~(\ref{otravez}), the argument starting in the second paragraph of the proof of Proposition \ref{interxp} yields the conclusion.}
\end{proof}
}%\textcolor{red}{Next we work with a fixed pair} $(P, k) \in \mathcal{P} \times }%\textcolor{red}{[n]_0}$ with $P=(P_1, \ldots, P_n)$ }%\textcolor{red}{and where each} $P_i=\{\alpha_1^i, \ldots, \alpha^{i}_{n(P_i)}\}$ }%\textcolor{red}{is ordered as described at the beginning of this section.} For }%\textcolor{red}{a subset} $I\subseteq [n]$ }%\textcolor{red}{consider} the set $T_I=\{ (b_1, \ldots, b_s) \in X^s\, \, | \,\, (b_{i1}, \ldots, b_{is})}%\textcolor{red}{{}\in T_{1,k_i}} \, \,\text{if and only if}\,\, i \in I\}.$ }%\textcolor{red}{Then}~(\ref{X^s_P}) }%\textcolor{red}{yields a topological disjoint union}
\begin{equation}\label{tdu4}
X^s_{P, \, k}= \bigsqcup_{\beta, I} \left(X^s_{P, \beta} \cap T_I\right)
\end{equation}
}%\textcolor{red}{running over subsets $I\subseteq[n]$ of cardinality $k$, and $n$-}%\textcolor{blue}{tuples} $\beta=}%\textcolor{red}(\beta^1, \ldots, \beta^n}%\textcolor{red})$ of (possibly empty) subsets} $\beta^i \subseteq \alpha^i_1- \{1\}$. }%\textcolor{red}{Besides, as suggested by~(\ref{tdu2}) in the proof of Proposition~\ref{queva}, it is convenient to decompose even further each piece in~(\ref{tdu4})}. For each $i \in }%\textcolor{red}{[n]}$, let
\begin{eqnarray}\label{T_01}
}%\textcolor{red}{T}_{0}( \alpha^i_1)&=&\{(b_1, \ldots, b_s) \in }%\textcolor{red}{X^s}\, \, | \,\, b_{ij} \neq }%\textcolor{red}{\pm} e^{}%\textcolor{red}{0}} \mbox{ for all }j \in \alpha^i_1 \}, \nonumber\\
}%\textcolor{red}{T}_{1}( \alpha^i_1)&=&\{(b_1, \ldots, b_s) \in }%\textcolor{red}{X^s}\, \, | \,\, b_{ij} = }%\textcolor{red}{\pm}e^{}%\textcolor{red}{0}}\,\, \text{for some}\,\,j \in \alpha^i_1 \}
\end{eqnarray}
}%\textcolor{red}{and, for $I}%\textcolor{red}{{}= \{\ell_1, \ldots, \ell_{|I|}\}}\subseteq[n]$ and $\varepsilon =(t_1, \ldots, t_{|I|})\in\{0,1\}^{|I|}$,}
$$T_{\varepsilon}(I)= }%\textcolor{red}{T_I\cap{}}\bigcap^{}%\textcolor{red}{|I|}}_{}%\textcolor{red}{ i =1} } }%\textcolor{red}{T}_{t_i}(\alpha^{}%\textcolor{red}{\ell_i}}_1).$$ }%\textcolor{red}{In these terms there is an additional topological disjoint union decomposition}
\begin{equation}\label{tdu10}
X^s_{P, \beta} \cap T_I = \bigsqcup_{\varepsilon \in}%\textcolor{red}{\{0,1\}^{|I|}}} \left(X^s_{P, \beta} \cap T_{\varepsilon}(I)\right).
\end{equation}
}%\textcolor{red}{Proposition~\ref{tarea1.1} is now a consequence of~(\ref{tdu4}),~(\ref{tdu10}), Lemma~\ref{separadotes}, and the following result:}
\begin{proposition}\label{construfinal}
}%\textcolor{red}{For $P$, $\beta$, I, and $\varepsilon$ as above, $X^s_{P, \beta} \cap T_{\varepsilon}(I)$ is an ENR admitting a local rule.}
\end{proposition}
\begin{proof}
}%\textcolor{red}{The ENR property follow since, in fact, $X^s_{P, \beta} \cap T_{\varepsilon}(I)$ is homeomorphic to the Cartesian product of a finite discrete space and a product of punctured spheres. Indeed, the information encoded by $P$ and $\beta$ produces the discrete factor, as coordinates in a single $G$-orbit are either repeated (e.g.~in the case of $\beta$) or sign duplicated. Besides, after ignoring such superfluous information as well as all $e^0$-coordinates (determined by $I$ and $\varepsilon$), we are left with a product of punctured spheres.}
\medskip
}%\textcolor{red}{The needed local rule can be defined following the algorithm at the end of Subsection~}%\textcolor{red}{\ref{suboddmotion}.} Explicitely, let $\rho_i$ ($i=0,1,2$) denote the local rules obtained by normalizing the corresponding $\lambda_i$ (defined in Example~\ref{TCs S even}) in the same manner as the local rules $\tau_i$ were obtained right after the proof of Proposition~\ref{interxp} from the corresponding $\phi_i$. Then consider the (non-continuous) global section $\varphi\colon X^s \to \mathbb{S}}%\textcolor{red}{(k_1,\ldots, k_n)}^{J_s}$ defined through the algorithm following~(\ref{elplanmot}), except that~(\ref{lasigma}) gets replaced by $$\sigma(b_{i1}, b_{ij})=\rho_m(b_{i1}, b_{ij}), \text{ if } (b_{i1}, b_{ij}) \in }%\textcolor{red}{B_m}\text{ for }m\in\{0,1,2\}$$ where the domains $}%\textcolor{red}{B_m}$ are now those defined in Example~\ref{TCs S even}. As in the previous subsection, the point is that the restriction of $\varphi$ to $X^s_{P, \beta} \cap T_{\varepsilon}(I)$ is continuous since, in that domain, the latter equality can be written as
$$
\sigma=\begin{cases}
\rho_2, &\mbox{if \ }j \in \left([s]-\alpha_1^i\right)\cup\beta^i\cup\{1\};\\
\rho_1, &\mbox{if \ }j \in \alpha_1^i - \left(\beta^i\cup\{1\}\right)\text{ and } t_i=0;\\
\rho_0, & \mbox{if \ }j \in \alpha_1^i - \left(\beta^i\cup\{1\}\right)\text{ and } t_i=1.
\end{cases}
$$
In addition, the proof of Proposition~\ref{noleave} applies word for word to show that the image of $\varphi$ is contained in $X^{J_s}$.}
\end{proof}
\begin{remark}{\em
}%\textcolor{red}{The gap noted in Remark~\ref{intento} also holds in~\cite{CohenPruid08} when all the $k_i$ are even. The new situation is subtler in view of an additional gap (pinpointed in~\cite[Remark~2.3]{morfismos}) in the proof of~\cite[Theorem~6.3]{CohenPruid08}. Of course, the detailed constructions in this section fix the problem and generalize the result.}}
\end{remark}
\section{Zero-divisors cup-length}\label{sectres}
}%\textcolor{red}{We now show that, for a subcomplex $X$ of $\mathbb{S}(k_1,\ldots,k_n)$ where all the $k_i$ have the same parity, the cohomological lower bound for $\protect\operatorname{TC}_s(X)$ in Proposition~\ref{ulbTCn} is optimal and agrees with the upper bound coming from our explicit motion planners in the previous section. Throughout this section we use cohomology with rational coefficients, writing $H^*(X)$ as a shorthand of $H^*(X;\mathbb{Q})$.}
\medskip
}%\textcolor{red}{Recall $H^*(\mathbb{S}(k_1,\ldots,k_n))$} is an exterior algebra }%\textcolor{red}{$E(\epsilon_1,\ldots,\epsilon_n)$ where $\epsilon_i$ corresponds to the $\mathbb{S}(k_i)$ factor, so that $\protect\operatorname{deg}(\epsilon_i)=k_i$.} For $J=\{j_1, \ldots, j_k\} }%\textcolor{red}{{}\subseteq [n]}$, let $\epsilon_J= \epsilon_{j_1}\cdots\epsilon_{j_k}$. }%\textcolor{red}{The cohomology ring $H^*(X)$ is a quotient of $E(\epsilon_1,\ldots,\epsilon_n)$:}
\begin{proposition}{\label{H^*(X)}}
}%\textcolor{red}{For} a subcomplex $X$ of the }%\textcolor{red}{minimal} CW-decomposition of }%\textcolor{red}{$\hspace{.4mm}\mathbb{S}(k_1\ldots,k_n)$,} the cohomology ring $H^*(X)$ is the quotient of the exterior algebra }%\textcolor{red}{$E(\epsilon_1,\ldots,\epsilon_n)$} by the }%\textcolor{red}{monomial} ideal $I_X$ generated by those $\epsilon_J$ }%\textcolor{red}{for which} $e_J$ is not a cell of $X$.
\end{proposition}
For a proof }%\textcolor{red}{(in a more general context)} of this proposition see \cite[Theorem 2.35]{BahBende10}. }%\textcolor{red}{In particular, an additive basis for $H^*(X)$ is given by the products $\epsilon_J$ with $e_J$ a cell of $X$. We will work with the corresponding tensor power basis for $H^*(X^s)$.}
\begin{remark}\label{elhandlingparodd}}%\textcolor{red}{\em
}%\textcolor{red}{In} the next two results, the hypothesis of having a fixed parity for all the $k_i$ }%\textcolor{red}{will be} crucial }%\textcolor{red}{when handling} products of zero divisors in $H^*(X^s)$. Indeed, a typical such element has the form
$$
z=c_1\cdot\epsilon_i\otimes1\otimes\cdots\otimes1+
c_2\cdot1\otimes\epsilon_i\otimes1\otimes\cdots\otimes1+\cdots+
c_s\cdot1\otimes\cdots\otimes1\otimes\epsilon_i
$$
for $i\in[n]$ and $c_1,\ldots,c_s\in\mathbb{Q}$ with $c_1+\cdots+c_s=0$. Then, by graded commutativity, $z^2$ is forced to vanish when $k_i$ is odd. However $z^s\neq0$ if $k_i$ is even and $c_j\neq0$ for all $j\in[s]$.
}\end{remark}
}%\textcolor{red}{Proposition~\ref{ulbTCn} and the following result complete the proof of Theorem~\ref{tcsX odd}.}
\begin{proposition}\label{cotaabajoimpar}
Let $X$ }%\textcolor{red}{be as in Proposition~$\ref{H^*(X)}$. If all of the $k_i$ are odd,} then $$\protect\operatorname{N}^{}%\textcolor{red}{s}}(X)\leq\protect\operatorname{zcl}_s(H^*(X)).$$
\end{proposition}
\begin{proof}
Let $H_X=H^*(X^s)=[H^*(X)]^{\otimes s}$. For $u \in H^*(X)$ and $2 \leq }%\textcolor{red}{\ell} \leq s$, let $$u(}%\textcolor{red}{\ell})=\underbrace{u \otimes 1 \otimes \cdots \otimes 1}_{s \text{ }%\textcolor{red}{factors}}} -\underbrace{1 \otimes \cdots \otimes 1 \otimes \overset{\,}%\textcolor{red}{\ell}}{u} \otimes 1 \otimes \cdots \otimes 1}_{s \text{ }%\textcolor{red}{factors}}} \in H_{X} $$ where an $}%\textcolor{red}{\ell}$ on top of a tensor factor indicates the coordinate where the factor appears. }%\textcolor{red}{Take a cell} $e_{J_1} \times e_{}%\textcolor{red}{J}_2} \times \cdots \times e_{J_s} }%\textcolor{blue}{{}\subseteq{}} X^s $, $J_1, \ldots, J_s}%\textcolor{red}{{}\subseteq[n]}$. For $2 \leq \ell \leq s$, let
\begin{center}
$\begin{array}{ll}
\gamma( J_1, \ldots, J_{\ell}) &= \displaystyle{\prod_{j \in \Big(\bigcap_{m=1}^{\ell-1} J_m - J_{\ell} \Big) \cup J_{\ell} } \epsilon_j (\ell)}
\\
& =\displaystyle{ \sum_{\phi_{\ell} \subseteq \Big(\bigcap_{m=1}^{\ell-1} J_m - J_{\ell} \Big) \cup J_{\ell}} \pm \epsilon_{\phi^c_{\ell}} \otimes 1 \otimes \cdots \otimes 1 \otimes \overset{\ell\hspace{2mm}}{\epsilon_{\phi_{\ell}}} \otimes 1 \otimes \cdots \otimes 1}
\end{array}$
\end{center}
where $\phi_{\ell}^c$ }%\textcolor{red}{stands for} the complement of $\phi_{\ell}$ in $\Big(\bigcap_{m=1}^{\ell-1} J_m - J_{\ell} \Big) \cup J_{\ell}$. }%\textcolor{red}{It suffices to prove the non-triviality of the} product of $\protect\operatorname{N}_X(J_1, \ldots, J_s)$ zero-divisors
\begin{equation}\label{prodzcls}
\gamma (J_1, J_2) \cdots \gamma (J_1, \ldots, J_s)= \sum_{\phi_ 2, \ldots, \phi_s} \pm \epsilon_{\phi_2^c}\cdots\epsilon_{\phi^c_s} \otimes \epsilon_{\phi_2} \otimes \cdots \otimes \epsilon_{\phi_s}
\end{equation}
where the sum runs over all $\phi_\ell }%\textcolor{red}{{}\subseteq{}} \Big(\bigcap_{m=1}^{\ell-1} J_m - J_{\ell} \Big) \cup J_{\ell} $ with $2 \leq \ell \leq s$. }%\textcolor{red}{With this in mind, note that} the term
\begin{equation}\label{eltermino}
\pm \epsilon_{J_1- J_2} \cdots \epsilon_{(J_1 \cap \cdots \cap J_{\ell-1})-J_{\ell}} \cdots \epsilon_{(J_1 \cap \cdots \cap J_{s-1})-J_s} \otimes \epsilon_{J_2} \otimes \cdots }%\textcolor{red}{{}\otimes \epsilon_{J_\ell}\otimes\cdots}\otimes\epsilon_{J_s},
\end{equation}
}%\textcolor{red}{which} appears in~(\ref{prodzcls}) }%\textcolor{red}{with} $\phi_\ell=J_\ell$ for $2 \leq \ell \leq s$, is }%\textcolor{red}{a basis element} because $$\epsilon_{J_1- J_2} \cdots \epsilon_{(J_1 \cap \cdots \cap J_{\ell-1})-J_\ell} \cdots \epsilon_{(J_1 \cap \cdots \cap J_{s-1})-J_s}= \epsilon_{J_0}$$ with $J_0 }%\textcolor{blue}{{}\subseteq{}} J_1$. }%\textcolor{red}{The non-triviality of~(\ref{prodzcls}) then follows by observing that~(\ref{eltermino})} cannot arise when other summands }%\textcolor{red}{in~(\ref{prodzcls})} are expressed in terms of }%\textcolor{red}{the} basis for $H_X$. }%\textcolor{red}{In fact, each summand
\begin{equation}\label{otrossumandos}
\pm \epsilon_{\phi_2^c} \cdots \epsilon_{\phi^c_s} \otimes \epsilon_{\phi_2} \otimes \cdots \otimes \epsilon_{\phi_s}
\end{equation}
in~(\ref{prodzcls}) is either zero or a basis element and, in the latter case,~(\ref{otrossumandos}) agrees (up to sign) with~(\ref{eltermino}) only if $\phi_{\ell}=J_{\ell}$ for $\ell=2,\ldots,s$.}
\end{proof}
}%\textcolor{red}{Likewise, the proof of Theorem~\ref{TC_s(X) even} is complete by Proposition~\ref{ulbTCn} and the following result:}
\begin{proposition}\label{casiultima}
}%\textcolor{red}{Let $X$ be as in Proposition~$\ref{H^*(X)}$. If all of the $k_i$ are even, then} $$s }%\textcolor{red}{\left(1+\dim(\mathcal{K}_X)\right)} \leq \protect\operatorname{zcl}_s (}%\textcolor{red}{H^*(X)}).$$
\end{proposition}
\begin{proof}
For $u \in H^*(X)$, set $$}%\textcolor{red}{\overline{u}}= \left(\hspace{.5mm}\sum_{i=1}^{s-1} 1 \otimes \cdots \otimes 1 \otimes \overset{i}{u} \otimes 1 \otimes \cdots \otimes 1\right)- 1 \otimes \cdots \otimes 1 \otimes (s-1) u}%\textcolor{red}{{\,}\in H_X.}$$
}%\textcolor{red}{Fix} a maximal cell $e_L$ of $X$ where $L=\{\delta_1, \ldots, \delta_{\ell}\} \subseteq }%\textcolor{red}{[n]}$ }%\textcolor{red}{(so $\ell=1+\dim(\mathcal{K}_X)$)}. }%\textcolor{red}{A straightforward calculation yields,} for }%\textcolor{red}{$i\in[\ell]$,}
$$
(\overline{\epsilon_{\delta_i}})^s=(1-s)s!(\underbrace{\epsilon_{\delta_i} \otimes \cdots \otimes \epsilon_{\delta_i}}_{s \text{ }%\textcolor{red}{factors}}}),
$$
so
$$
\prod_{i=1}^{\ell} (\overline{\epsilon_{\delta_i}})^s= }%\textcolor{red}{\left((1-s)s!\right)^\ell} \underbrace{\epsilon_L \otimes \cdots \otimes \epsilon_L}_{s \text{ }%\textcolor{red}{factors}}}
$$
}%\textcolor{red}{which is a nonzero product of $s\ell$ zero-divisors in $H_X$.}
\end{proof}
\begin{remark}{\em
}%\textcolor{red}{The estimate $s(1+\dim(\mathcal{K}_{}%\textcolor{red}{X}}))\leq\protect\operatorname{TC}_s(X)$ can also be obtained by noticing that, in the notation of the proof of Proposition~\ref{casiultima}, $\mathbb{S}(k_{\delta_1},\ldots,k_{\delta_\ell})\cong \overline{e_L}$ is a retract of $X$ (c.f.~\cite[proof of Proposition~4]{FelixTanre09}).}
}\end{remark}
}%\textcolor{red}{It well known that, under suitable normality conditions, the higher topological complexity of a Cartesian product can be estimated by
\begin{equation}\label{laestimacion}
\protect\operatorname{zcl}_s(H^*(X))+\protect\operatorname{zcl}_s(H^*(Y))\leq \protect\operatorname{zcl}_s(H^*(X\times Y))\leq \protect\operatorname{TC}_s(X \times Y) \leq \protect\operatorname{TC}_s(X) + \protect\operatorname{TC}_s(Y),
\end{equation}
see~\cite[Proposition 3.11]{bgrt} and~\cite[Lemma~2.1]{CoFa11}. Of course, these inequalities are sharp provided $\protect\operatorname{TC}_s=\protect\operatorname{zcl}_s$ for both $X$ and $Y$. In particular, for the spaces dealt with in }%\textcolor{red}{Theorem~\ref{TC_s(X)},} $\protect\operatorname{TC}_s$ is additive in the sense that the higher topological complexity of a Cartesian product is the sum of the higher topological complexities of the factors. This generalizes the known $\protect\operatorname{TC}_s$-behavior of products of spheres, see~\cite[Corollary~3.12]{bgrt}. However, if Cartesian products are replaced by wedge sums, the situation becomes much subtler. To begin with, we remark that Theorem~3.6 and Remark~3.7 in~\cite{MR3267004}, together with~\cite[Theorem~19.1]{farber06}, give evidence suggesting that a reasonable wedge-substitute of~(\ref{laestimacion}) (for $s=2$) would be given by
$$
\protect\operatorname{max}\{\protect\operatorname{TC}_2(X),\protect\operatorname{TC}_2(Y),\protect\operatorname{cat}(X\times Y)\}\leq\protect\operatorname{TC}_2(X\vee Y)\leq\protect\operatorname{max}\{\protect\operatorname{TC}_2(X),\protect\operatorname{TC}_2(Y),\protect\operatorname{cat}(X)+\protect\operatorname{cat}(Y)\}.
$$
We show that both of these inequalities hold as equalities for the spaces dealt with in the previous section (c.f.~\cite[Proposition~3.10]{CohenPruid08}). More generally:}
\begin{proposition}\label{farberdranishnikov}
}%\textcolor{red}{Let $X$ and $Y$ be subcomplexes of $\hspace{.6mm}\mathbb{S}(k_1 \ldots, k_n)$ and $\hspace{.6mm}\mathbb{S}(k_{n+1}, \ldots, k_{n+m})$ respectively. If $\protect\operatorname{cat} (X) \geq \protect\operatorname{cat} (Y)$ and all the $k_i$ have the same parity, then}
$$}%\textcolor{red}{\protect\operatorname{TC}_s(X \vee Y)= \protect\operatorname{max} \{\protect\operatorname{TC}_s(X), \protect\operatorname{TC}_s(Y), \protect\operatorname{cat}(X^{s-1})+ \protect\operatorname{cat} (Y)\}.}$$
\end{proposition}
\begin{proof}
}%\textcolor{red}{If all the $k_i$ are even, the conclusion follows directly from Theorem~\ref{TC_s(X) even} and Remark~\ref{downtocat} at the end of the paper. In fact $\protect\operatorname{TC}_s(X\vee Y)=\protect\operatorname{TC}_s(X)$ under the present hypothesis.}
\smallskip
}%\textcolor{red}{Assume now that all the $k_i$ are odd, and think of $X \vee Y$ as a subcomplex of $X \times Y$ inside $\mathbb{S}(k_1, \ldots, k_n, k_{n+1}, \ldots, k_{n+m})$, so that $\mathcal{K}_{X\vee Y}$ }%\textcolor{red}{is the disjoint union of $\mathcal{K}_{X}$ and $\mathcal{K}_{Y}$.} Since $\protect\operatorname{cat} (X)=\dim(\mathcal{K}_X)+1\geq\protect\operatorname{cat} (Y)=\dim( \mathcal{K}_Y)+1$, for maximal simplexes $J_1,\ldots,J_s$ of $\mathcal{K}_{X\vee Y}$ we see
\begin{equation}\label{word4word}
}%\textcolor{red}{\protect\operatorname{N}}_{X \vee Y} (J_1, \ldots, J_s) \leq \begin{cases}
\protect\operatorname{TC}_s(X), & \text{if} \,\, J_1, \ldots, J_s \subseteq [n] ; \\
\protect\operatorname{TC}_s(Y), & \text{if} \,\, J_1, \ldots, J_s \subseteq \{n+1, \ldots, n+m\} ; \\
(s-1)\protect\operatorname{cat} (X)+ \protect\operatorname{cat} (Y), & \text{otherwise}.
\end{cases}
\end{equation}
Therefore $\protect\operatorname{TC}_s(X \vee Y) \leq \protect\operatorname{max} \{\protect\operatorname{TC}_s(X), \protect\operatorname{TC}_s(Y), \, (s-1)\protect\operatorname{cat}(X)+ \protect\operatorname{cat} (Y)\}$. The reverse inequality holds since each of $\protect\operatorname{TC}_s(X)$, $\protect\operatorname{TC}_s(Y)$, and $(s-1) \protect\operatorname{cat} (X)+ \protect\operatorname{cat} (Y)$ can be achieved as a $\protect\operatorname{N}_{X \vee Y} (J_1, \ldots, J_s)$ for a suitable combination of maximal simplexes $J_i$ of $\mathcal{K}_{X\vee Y}$.}
\end{proof}
\section{}%\textcolor{red}{The unrestricted} case}
}%\textcolor{red}{We now prove Theorem~\ref{TC_s(X)} in} the general case, }%\textcolor{red}{that is for} $X$ a subcomplex of $\hspace{.3mm}\mathbb{S}(k_1, \ldots, k_n)$ where all }%\textcolor{red}{the} $k_i$ are positive integers with no restriction on their parity. }%\textcolor{red}{As usual, we start by establishing the upper bound.}
\subsection{Motion planner}\label{stn4.1}
}%\textcolor{red}{Consider the disjoint union decomposition $[n]= J_E \sqcup J_O $ where $J_E$ is the collection of indices $i\in[n]$ for which $k_i$ is even (thus $i \in J_O$ if and only if $k_i$ is odd)}. For a subset $K \subseteq J_E$ and $P \in \mathcal{P}$, }%\textcolor{red}{let $X^s_{P, K}\subseteq X^s$ and $\protect\operatorname{N}(P, K)$, the norm of $(P, K)$, be defined by}
\begin{itemize}
\item $X^s_{P, K}= X_P^s \cap \left\{(b_1, \ldots, b_s) \in X^s\; \Big|\hspace{-3.6mm} \begin{array}{ll} & \text{for }%\textcolor{red}{each $(i,j)\in K\times[s],\;\;$}} b_{ij}\neq \pm e^0, \,\,\,\, }%\textcolor{red}{\text{while}}\\ & }%\textcolor{red}{\text{for each } i \in J_E- K \text{ there is}\,\, j\in [s] \text{ with } b_{ij}= \pm e^0}\!\!
\end{array} \right\}$
\item $\protect\operatorname{N}(P, K)= |P|+ |K|$ where $|P|$ is defined in~}%\textcolor{red}{(\ref{lanormadeP}).}
\end{itemize}
}%\textcolor{red}{This extends the definitions of $X^s_{P, k}$ and $\protect\operatorname{N}(P, k)$ done when all the $k_i$ are even.}
\smallskip As in the cases where all }%\textcolor{red}{the $k_i$} have the same parity, the }%\textcolor{red}{higher topological complexity of a} subcomplex $X$ of $\mathbb{S}(k_1, \ldots, k_n)$, }%\textcolor{red}{now} with no restrictions on the parity of the sphere }%\textcolor{red}{factors}, is encoded just by }%\textcolor{red}{the} combinatorial information }%\textcolor{red}{on the cells} of $X$. Consider
\begin{equation}\label{sjdhfuxji}
\mathcal{N}^s(X)= \protect\operatorname{max} \left\{\left.\protect\operatorname{N}_X(J_1, \ldots, J_s) + \Big|\bigcap_{i=1}^s J_i \cap J_E \Big|\;\; \right| \,\, }%\textcolor{red}{J_1,\ldots,J_s\in\mathcal{K}_X} \right\}
\end{equation}
where $\protect\operatorname{N}_X(J_1, \ldots, J_s)$ is defined in~(\ref{nx}) for $\mathcal{K}=\mathcal{K}_X$. }%\textcolor{red}{Since both $\protect\operatorname{N}_X(J_1, \ldots, J_s)$ and $|\bigcap_{i=1}^s J_i \cap J_E|$ are monotonically non-decreasing functions of the $J_i$'s, the definition of $\mathcal{N}^s(X)$ can equally be given using only maximal simplexes $J_i\in\mathcal{K}_X$.} }%\textcolor{red}{Further, by~(\ref{sy}),} $\mathcal{N}^s(X)$ can be rewritten as
\begin{equation}\label{gtc_sX}
\mathcal{N}^s(X)= \protect\operatorname{max} \left\{\displaystyle{\sum_{i=1}^s |J_i| - \Big| \bigcap_{i=1}^s J_i \cap J_O \Big|} \,\, | \,\, e_{J_i} \,\, \text{is a cell of } \,\, X, \, \text{for all}\,\, i \in [s] \right\}.
\end{equation}
\begin{theo}\label{TC_s(X), gene}
}%\textcolor{red}{For} a subcomplex $X$ of $\hspace{.4mm}\mathbb{S}(k_1, \ldots, k_n)$, $$\protect\operatorname{TC}_s(X)= \mathcal{N}^s(X).$$
\end{theo}
}%\textcolor{red}{Theorem~\ref{TC_s(X), gene} generalizes Theorems~\ref{tcsX odd} and~\ref{TC_s(X) even}. This is obvious} when all }%\textcolor{red}{the $k_i$} are odd }%\textcolor{red}{for then both $\mathcal{N}^s(X)$ and $\protect\operatorname{N}^s(X)$ agree with} $$\protect\operatorname{max} \left\{\left.\displaystyle{\sum_{i=1}^s |J_i| - \Big| \bigcap_{i=1}^s J_i \Big|} \,\, \right| \,\, }%\textcolor{red}{J_1,\ldots,J_s\in\mathcal{K}_X} \right\},$$ }%\textcolor{red}{whereas if all the $k_i$} are even, $$\mathcal{N}^s(X)= \protect\operatorname{max} \left\{\displaystyle{\sum_{i=1}^s |J_i| } \,\, \big| \,\, }%\textcolor{red}{J_1,\ldots,J_s\in\mathcal{K}_X} \right\} =s (1+ \dim \mathcal{K}_X).$$
}%\textcolor{red}{The estimate $\mathcal{N}^s(X)\leq\protect\operatorname{TC}_s(X)$ in Theorem~\ref{TC_s(X), gene} will be proved in the next subsection by extending the cohomological methods in Section~\ref{genzero}. Here we prove the estimate $\protect\operatorname{TC}_s(X)\leq\mathcal{N}^s(X)$ by constructing} an optimal motion planner }%\textcolor{red}{with $\mathcal{N}^s(X)+1$ local rules. The corresponding local domains will be obtained} by clustering subsets $X^s_{P, K}$ }%\textcolor{red}{for which the pair} $(P, K)\in \mathcal{P}\times 2^{J_E}$ }%\textcolor{red}{has a fixed} norm. }%\textcolor{red}{In detail,} for $j \in [\mathcal{N}^s(X)]_0$ }%\textcolor{red}{let}
\begin{equation}\label{lagejotatopologicamente}
G_j:= \bigcup_{\protect\operatorname{N}(P, K)=j} X^s_{P, K}.
\end{equation}
\begin{lema}\label{lemma4.2pairwisedisjointcovering}
}%\textcolor{red}{The sets $G_0,\ldots,G_{\mathcal{N}^s(X)}$ yield a pairwise disjoint covering of $X^s$.}
\end{lema}
\begin{proof}
}%\textcolor{red}{It is easy to see that $G_{j}\cap G_{j'}=\emptyset$ for $j\neq j'$.} Let $b=(b_1, \ldots, b_s) \in e_{J_1} \times \cdots \times e_{J_s} \subseteq X^s$, where $J_j \subseteq [n]$ for $j \in [s]$. As in Lemma \ref{coverodd}, we have
\begin{equation}\label{cg1nueva}
\sum_{i=1}^n |\{G\cdot b_{ij}\,\, |\,\, j\in[s]\}|-n \leq \sum_{j=1}^s \, |J_{j}| - \Big|\bigcap_{j=1}^s J_j \Big|= \protect\operatorname{N}_X(J_1, \ldots, J_s).
\end{equation}
Moreover, it is clear that
\begin{equation}\label{cg2nueva}
\Big| \left\{ }%\textcolor{red}{i \in J_E} \; | \,\, b_{ij} \neq \pm e^0, \, \, \forall j\in [s]\right\} \Big| \leq \Big| \bigcap_{i=1}^s J_i \cap J_E \Big|.
\end{equation}
Thus, if $P \in \mathcal{P}$ is the type of $b$, and }%\textcolor{red}{$K \subseteq J_E$ is determined by the condition that} $b \in X^s_{P, K}$, }%\textcolor{red}{then $N(P,K)=|P|+|K|\leq\mathcal{N}^s(X)$ in view of~(\ref{Pcardinality}), (\ref{cg1nueva}) and~(\ref{cg2nueva}).}
\end{proof}
\begin{lema}\label{emptyinterg}
}%\textcolor{red}{$(\ref{lagejotatopologicamente})$ is a topological disjoint union. Indeed,}
\begin{equation}\label{disuniong}
X^s_{P,K} \cap \overline{X^s_{P', K'}}= \emptyset=\overline{ X^s_{P,K}} \cap X^s_{P', K'}
\end{equation}
for $(P, K), (P', K') \in \mathcal{P} \times 2^{J_E}$ provided that $(P, K) \neq (P', K')$ and $\protect\operatorname{N}(P, K)=\protect\operatorname{N}(P', K')$.
\end{lema}
}%\textcolor{red}{The} following }%\textcolor{red}{observation} will be useful }%\textcolor{red}{in the proof of Lemma~\ref{emptyinterg}:}
\begin{remark}\label{noindex}{\em
Let $K, K' \subseteq 2^{J_E}$ and }%\textcolor{red}{$P, P' \in \mathcal{P}$. If} there exists an index $i \in }%\textcolor{red}{K-K',}$ }%\textcolor{red}{then}
\begin{itemize}
\item $b_{ij} \neq \pm e^0$ for all $j \in [s]$ }%\textcolor{red}{provided} $b=(b_1, \ldots, b_s)\in X^s_{P, K}$.
\item $b_{ij_0}= \pm e^0$ for some $j_0 \in [s]$ }%\textcolor{red}{provided} $b=(b_1, \ldots, b_s) \in X^s_{P', K'}$.
\end{itemize}
Therefore, $X^s_{P,K} \cap \overline{X^s_{P', K'}}= \emptyset$.
}\end{remark}
\begin{proof}[Proof of Lemma~$\ref{emptyinterg}$] }%\textcolor{red}{There are three possibilities:}
\medskip
\noindent \textbf{Case} $K=K'$. In this case, one conclude that $P \neq P'$ with $|P|=|P'|$, since $(P, K) \neq (P', K')$ and $\protect\operatorname{N}(P, K)=\protect\operatorname{N}(P', K')$. The desired equalities follow from Proposition \ref{interxp}.
\medskip
\noindent\textbf{Case} $P=P'$. In this case we have $K \neq K'$ with $|K|=|K'|$. Then, there }%\textcolor{red}{exist} indexes $i, i'}%\textcolor{red}{\in[n]}$ such that $i \in }%\textcolor{red}{K - K}'$ and $i' \in }%\textcolor{red}{K'-K.}$ Therefore, equalities~(\ref{disuniong}) follow from Remark \ref{noindex}.
\medskip
\noindent\textbf{Case} $P\neq P'$ and $K\neq K'$. }%\textcolor{red}{Without loss of generality we can assume} $|P|> |P'|$. Then there exists $i \in [n]$ such that $|P_i|> |P_i'|$, thus $X^s_{P,K} \cap \overline{X^s_{P', K'}}= \emptyset$. Moreover, }%\textcolor{red}{since} $|K| < |K'|$ }%\textcolor{red}{is forced,} there exits $i \in }%\textcolor{red}{K' - K,}$ }%\textcolor{red}{so that} $\overline{X^s_{P,K}} \cap X^s_{P', K'}= \emptyset$ by }%\textcolor{red}{Remark~\ref{noindex}.}
\end{proof}
}%\textcolor{red}{Lemmas~\ref{lemma4.2pairwisedisjointcovering} and~\ref{emptyinterg} reduce the proof of Theorem~\ref{TC_s(X), gene} to checking that each $X^s_{P,K}$ is an ENR admitting a local rule. Thus, troughout the remaining of this subsection we fix a pair $(P, K) \in \mathcal{P} \times 2^{J_E}$ with $P=(P_1, \ldots, P_n)$ and where each $P_i=\{\alpha_1^i, \ldots, \alpha^i_{n(P_i)}\}$ is assumed to be ordered as indicated at the beginning of Section~\ref{secciondosmeramera}.}
\medskip }%\textcolor{red}{Our analysis of $X^s_{P,K}$ depends on establishing a topological decomposition of $X^s_{P,K}$. To start with, note the topological} disjoint union decomposition $$X^s_{P,K}=\,\bigsqcup_{\beta} \,X^s_{P, K} \cap X^{}%\textcolor{red}{s}}_{P, \beta}$$ where the union runs over all $\beta=(\beta^1, \ldots, \beta^n)$ as in~(\ref{X^s_P}). }%\textcolor{red}{But we need a further splitting of each term $X^s_{P, K} \cap X^s_{P, \beta}$.}
\medskip Let $I=\{\ell_1, \ldots, \ell_{|I|}\}$ }%\textcolor{red}{stand for} $J_E- K$ and, for each $i \in [n]$, consider the subsets }%\textcolor{red}{$T_0(\alpha_1^i)$ and $T_1(\alpha_1^i)$ defined in~(\ref{T_01}).} }%\textcolor{red}{For each} $\epsilon=(t_1, \ldots, t_{|I|}) \in \{0, 1\}^{|I|}$ }%\textcolor{red}{define} $$}%\textcolor{red}{T_{\epsilon}}=\bigcap_{i = 1}^{|I|} T_{t_i}( \alpha_1^{\ell_i}).$$ }%\textcolor{red}{We then get a topological disjoint union decomposition} $$X^s_{P, K} \cap X^s_{P, \beta}\,= \,\bigsqcup_{}%\textcolor{red}{\epsilon \in \{0, 1\}^{|I|}}} \, X^s_{P, K} \cap X^s_{P, \beta} \cap}%\textcolor{red}{T_{\epsilon}.}$$ Therefore, }%\textcolor{red}{the updated task is the proof of:}
\begin{lema}\label{updatedtarea}
}%\textcolor{red}{Each $ X^s_{P, K, \beta, \epsilon} := X^s_{P, K} \cap X^s_{P, \beta} \cap }%\textcolor{red}{T_{\epsilon}}$ is an ENR admitting a local rule.}
\end{lema}
\begin{proof}
}%\textcolor{red}{The ENR assertion follows just as in the first paragraph of the proof of Proposition~\ref{construfinal}. The construction of the local rule is also similar to the those at the end of Subsections~\ref{suboddmotion} and~\ref{laspares}, and we provide the generalized details for completeness.}
\smallskip For $i=0,1$ and $j=0, 1, 2$, let $\tau_i$ and $\rho_j$ be the local rules, }%\textcolor{red}{with corresponding local domains $A_i$ and $B_j$,} obtained }%\textcolor{red}{in Subsections~\ref{suboddmotion} and~\ref{laspares}} by normalizing the local rules $\phi_i$ and $\lambda_j$ given in Examples \ref{TCs S} and \ref{TCs S even} }%\textcolor{red}{---see the proof of Proposition~\ref{construfinal} and the considerations following the proof of Proposition~\ref{interxp}.}
\medskip
As }%\textcolor{red}{before,} it is useful to keep in mind that elements $(}%\textcolor{red}{b}_1, \ldots, }%\textcolor{red}{b}_s) \in X^s$, with $}%\textcolor{red}{b}_j=(}%\textcolor{red}{b}_{1j}, \ldots, }%\textcolor{red}{b}_{nj})$ for $j\in }%\textcolor{red}{[s]}$, }%\textcolor{red}{can be thought of as matrices $(b_{i,j})$ whose columns represent the various stages in $X$ through which motion is to be planned (necessarily along rows). Again, we follow a pivotal strategy. In detail, in terms of the notation set at the beginning of the introduction for elements in the function space $X^{J_s}$,} consider the map
\begin{equation}\label{elplanmotbis}
}%\textcolor{red}{}%\textcolor{green}{\varphi}\colon X^s} \to \mathbb{S}}%\textcolor{red}{(k_1,\ldots, k_n)}^{J_s}
\end{equation}
}%\textcolor{red}{given by} $}%\textcolor{green}{\varphi}\left((}%\textcolor{red}{b}_1, \ldots, }%\textcolor{red}{b}_s)\right)=\left(}%\textcolor{green}{\varphi_1}(}%\textcolor{red}{b}_1, }%\textcolor{red}{b}_1), \ldots , }%\textcolor{green}{\varphi_s}(}%\textcolor{red}{b}_1, }%\textcolor{red}{b}_s)\right)$ where, }%\textcolor{red}{for $j\in[s]$,} $$}%\textcolor{green}{\varphi_j}(}%\textcolor{red}{b}_1, }%\textcolor{red}{b}_j)= \left(}%\textcolor{green}{\varphi_{1j}} (}%\textcolor{red}{b}_{11}, }%\textcolor{red}{b}_{1j}), \ldots, }%\textcolor{green}{\varphi_{nj}}(}%\textcolor{red}{b}_{n1}, }%\textcolor{red}{b}_{nj})\right)$$ }%\textcolor{red}{is the path in $\mathbb{S}(k_1,\ldots,k_n)$, from $b_1$ to $b_j$, whose $i$-th coordinate} $}%\textcolor{green}{\varphi_{ij}}(}%\textcolor{red}{b}_{i1}, }%\textcolor{red}{b}_{ij})$, }%\textcolor{red}{}%\textcolor{red}{$i\in[n]$,} is the path in $S^{k_i}$, from $b_{i1}$ to $b_{ij}$,} defined by $$}%\textcolor{green}{\varphi_{i,j}} (}%\textcolor{red}{b}_{i1},}%\textcolor{red}{b}_{ij})(t)=\begin{cases} }%\textcolor{red}{b}_{i1}, & 0 \leq t \leq t_{ }%\textcolor{red}{b}_{i1}}, \\ }%\textcolor{red}{\sigma}(}%\textcolor{red}{b}_{i1}, }%\textcolor{red}{b}_{ij})(t-t_{}%\textcolor{red}{b}_{i1}}), & t_{}%\textcolor{red}{b}_{i1}} \leq t \leq 1. \end{cases}$$ Here $t_{}%\textcolor{red}{b}_{i1}}=\frac{1}{2} -d(}%\textcolor{red}{b}_{i1}, }%\textcolor{red}{e^0})$ and
\begin{equation}\label{lasigmabis}
}%\textcolor{red}{\sigma(b_{i1}, b_{ij})}=\begin{cases}\tau_0}%\textcolor{red}{(b_{i1}, b_{ij})}, & }%\textcolor{red}{\text{if } i\in J_O \text{ and }} (}%\textcolor{red}{b}_{i1}, }%\textcolor{red}{b}_{ij}) \in A_0; \\ \tau_1}%\textcolor{red}{(b_{i1}, b_{ij})}, & }%\textcolor{red}{\text{if } i\in J_O \text{ and }} (}%\textcolor{red}{b}_{i1}, }%\textcolor{red}{b}_{ij}) \in A_1;\\
\rho_0(b_{i1}, b_{ij}), & }%\textcolor{red}{\text{if } i\in J_E \text{ and }} (b_{i1}, b_{ij}) \in B_0;\\
\rho_1(b_{i1}, b_{ij}), & }%\textcolor{red}{\text{if } i\in J_E \text{ and }}(b_{i1}, b_{ij}) \in B_1;\\
\rho_2(b_{i1}, b_{ij}), & }%\textcolor{red}{\text{if } i\in J_E \text{ and }}(b_{i1}, b_{ij}) \in B_2. \end{cases}
\end{equation}
}%\textcolor{red}{Although }%\textcolor{green}{$\varphi$} is not continuous, its restriction $}%\textcolor{green}{\varphi}_{P, K, \beta, \epsilon}$ to $X^s_{P, K, \beta, \epsilon}$ is, for then~(\ref{lasigmabis}) takes the form}
$$
\sigma=\begin{cases}\tau_1, & \mbox{$}%\textcolor{red}{i\in J_O,}\,\,\, j \notin \alpha^i_{1}$ or $j \in }%\textcolor{red}{\beta^i\cup\{1\}}$}; \\ \tau_0, & \mbox{$}%\textcolor{red}{i\in J_O,}\,\,\, j \in \alpha^i_{1}$ and $j \notin }%\textcolor{red}{\beta^i\cup\{1\}}$};\\
\rho_2, & i\in J_E,\,\,}%\textcolor{red}{j \notin \alpha^i_{1}\,\, \text{or} \,\, j \in }%\textcolor{red}{\beta^i\cup\{1\}}};\\
\rho_1, & i\in J_E,\,\, j\in \alpha_1^i - \left(\beta^i\cup\{1\}\right)\text{ and } t_i=0;\\
\rho_0, & i\in J_E,\,\, j \in \alpha_1^i - \left(\beta^i\cup\{1\}\right)\text{ and } t_i=1. \end{cases}
$$
}%\textcolor{red}{Moreover,} $}%\textcolor{green}{\varphi}_{P, K, \beta, \epsilon}$ is clearly a section for $e_s^{}%\textcolor{red}{\mathbb{S}(k_1,\ldots,k_n)}}$, }%\textcolor{red}{while the} fact that $\varphi_{P, K, \beta, \epsilon}$ actually takes values in $X^{J_s}$ is }%\textcolor{red}{verified with an argument identical to the one proving} Proposition~\ref{noleave}.
\end{proof}
\subsection{Zero-divisors cup-length}\label{genzero}
We }%\textcolor{red}{next} show that, for a subcomplex $X$ of $\mathbb{S}(k_1,\ldots,k_n)$ (with no restrictions on the parity of the $k_i$, $i \in [n]$), the cohomological lower bound for $\protect\operatorname{TC}_s(X)$ in Proposition~\ref{ulbTCn} is optimal and agrees with the upper bound coming from our explicit motion }%\textcolor{red}{planner} in the previous }%\textcolor{red}{subsection. Here} we use same considerations and notation as in Section \ref{sectres}.
\begin{proposition}\label{cotaabajoimparbis}
}%\textcolor{red}{A subcomplex $X$ of $\,\mathbb{S}(k_1,\ldots,k_n)$ has} $$\mathcal{N}^{}%\textcolor{red}{s}}(X)\leq\protect\operatorname{zcl}_s(H^*(X)).$$
\end{proposition}
\begin{proof}
}%\textcolor{red}{We use the tensor product ring $H_X$, and the elements $u(\ell)\in H_X$ for} $u \in H^*(X)$, }%\textcolor{red}{as well as the elements $\gamma(J_1,\ldots,J_\ell)\in H_X$ for $J_1,\ldots,J_\ell\in\mathcal{K}_X$ defined for $2 \leq }%\textcolor{red}{\ell} \leq s$ at the beginning of the proof of Proposition~\ref{cotaabajoimpar} (but this time we will only need the latter elements in the range $3\leq\ell\leq s$). In addition,} let $J'= \bigcap_{j=1}^s J_j \cap J_E$ and consider
\begin{eqnarray}
\bar{\epsilon}_{J'}&=& \prod_{j \in J'} (\epsilon_j \otimes 1 \otimes \cdots \otimes 1-1 \otimes \epsilon_j \otimes 1 \otimes \cdots \otimes 1)^2\label{elemento34decorrecion}\\
&=& (-2)^{|J'|} \epsilon_{J'} \otimes \epsilon_{J'} \otimes 1 \otimes \cdots \otimes 1 \nonumber
\end{eqnarray}
and
\begin{eqnarray}
\bar{\gamma}(J_1, J_2)&=& \displaystyle{\prod_{j \in (J_1-J_2) \cup (J_2-J')} \epsilon_j (2)}
\label{elemento35decorreccion} \\ \nonumber
&=& \displaystyle{ \sum_{\phi_{2} \subseteq (J_1-J_2 )\cup (J_2-J')} \pm \epsilon_{\phi^c_{2}} \otimes \epsilon_{\phi_{2}} \otimes 1 \otimes \cdots \otimes 1}
\end{eqnarray}
}%\textcolor{red}{where, as in the proof of Proposition~\ref{cotaabajoimpar}, $\phi^c_2$ stands for the complement of $\phi_2$ in $ (J_1-J_2 )\cup (J_2-J')$. Then}
\begin{equation}\label{prodzclsbis}
\bar{\epsilon}_{J'}\cdot \bar{\gamma} (J_1, J_2) \cdot\prod_{\ell=3}^{s} \gamma (J_1, \ldots, J_{\ell})= \sum_{\phi_ 2, \ldots, \phi_s} \pm 2^{|J'|} \epsilon_{J'} \epsilon_{\phi_2^c} \cdots\epsilon_{\phi^c_s} \otimes \epsilon_{J'} \epsilon_{\phi_2} \otimes }%\textcolor{red}{\epsilon_{\phi_3}}\otimes\cdots \otimes \epsilon_{\phi_s}
\end{equation}
}%\textcolor{red}{where, for $3\leq\ell\leq s$,
$$
\phi_\ell }%\textcolor{red}{{}\subseteq{}} \Big(\bigcap_{m=1}^{\ell-1} J_m - J_{\ell} \Big) \cup J_{\ell}
$$
with $\phi^c_{\ell}$ standing for the complement of $\phi_{\ell}$ in $\Big(\bigcap_{m=1}^{\ell-1} J_m - J_{\ell} \Big) \cup J_{\ell}$ ---here we are using the notation in Proposition~\ref{cotaabajoimpar}. Recalling that} $$\protect\operatorname{N}_X(J_1, \ldots, J_s)=\sum_{\ell=2}^s \left(\Big| \bigcap_{m=1}^{\ell-1} J_m- J_{\ell} \Big|+ \Big| J_{\ell} \Big| \right),$$ }%\textcolor{red}{we easily see that the left-hand side of~(\ref{prodzclsbis})} is a product of $\protect\operatorname{N}_X(J_1, \ldots, J_s)+ |\bigcap_{j=1}^sJ_j \cap J_E|$ zero-divisors. }%\textcolor{red}{Thus, by~(\ref{sjdhfuxji}),} it suffices to prove the non-triviality of the }%\textcolor{red}{right-hand side of~(\ref{prodzclsbis}).} }%\textcolor{red}{With this in mind, note that} the term
\begin{equation}\label{elterminobis}
\pm 2^{|J'|} \, \epsilon_{J'} \, \epsilon_{J_1- J_2} \, }%\textcolor{red}{\epsilon_{(J_1 \cap J_2)-J_3}} \cdots \epsilon_{(J_1 \cap \cdots \cap J_{s-1})-J_s} \otimes \epsilon_{J_2} \otimes \cdots \otimes\epsilon_{J_s},
\end{equation}
}%\textcolor{red}{which} appears in~(\ref{prodzclsbis}) }%\textcolor{red}{with} \,$\phi_\ell=J_\ell$ for $3 \leq \ell \leq s$ and $\phi_2 =J_2-J'$, is }%\textcolor{red}{a basis element} because $$ \epsilon_{J'} \cdot\epsilon_{J_1- J_2} \cdots \epsilon_{(J_1 \cap \cdots \cap J_{\ell-1})-J_\ell} \cdots \epsilon_{(J_1 \cap \cdots \cap J_{s-1})-J_s}= \epsilon_{J'}\cdot \epsilon_{(J_1- \cap_{j=1}^s J_j)}= \epsilon_{J_0}$$ with $J_0 }%\textcolor{blue}{{}\subseteq{}} J_1$. }%\textcolor{red}{The non-triviality of~(\ref{prodzclsbis}) then follows by observing that~(\ref{elterminobis})} cannot arise when other summands }%\textcolor{red}{in~(\ref{prodzclsbis})} are expressed in terms of }%\textcolor{red}{the} basis for $H_X$. }%\textcolor{red}{In fact, each summand
\begin{equation}\label{otrossumandosbis}
\pm 2^{|J'|} \epsilon_{J'} \epsilon_{\phi_2^c} \cdots \epsilon_{\phi^c_s} \otimes \epsilon_{J'} \epsilon_{\phi_2} \otimes }%\textcolor{red}{\epsilon_{\phi_3}} \otimes\cdots \otimes \epsilon_{\phi_s}
\end{equation}
in~(\ref{prodzclsbis}) is either zero or a basis element and, in the latter case,~(\ref{otrossumandosbis}) agrees (up to sign) with~(\ref{elterminobis}) only if $\phi_{\ell}=J_{\ell}$ for $\ell=3,\ldots,s$, and $\phi_2= J_2-J'$.}
\end{proof}
\begin{remark}{\em
}%\textcolor{red}{The factors~(\ref{elemento34decorrecion}) and~(\ref{elemento35decorreccion}) adjust the product~(\ref{prodzcls}) of zero divisors in the proof of Proposition~\ref{cotaabajoimpar} so to account for the differences noted in Remark~\ref{elhandlingparodd}.}
}\end{remark}
}%\textcolor{red}{We close the section by noticing that Proposition~\ref{farberdranishnikov} holds without restriction on the parity of the sphere dimensions $k_1,\ldots,k_{n+m}$. }%\textcolor{red}{That is:
\begin{proposition} Let $X$ and $Y$ be subcomplexes of $\hspace{.6mm}\mathbb{S}(k_1 \ldots, k_n)$ and $\hspace{.6mm}\mathbb{S}(k_{n+1}, \ldots, k_{n+m})$ respectively. If $\protect\operatorname{cat} (X) \geq \protect\operatorname{cat} (Y)$, then
$$}%\textcolor{red}{\protect\operatorname{TC}_s(X \vee Y)= \protect\operatorname{max} \{\protect\operatorname{TC}_s(X), \protect\operatorname{TC}_s(Y), \protect\operatorname{cat}(X^{s-1})+ \protect\operatorname{cat} (Y)\}.}$$ \end{proposition}}
The argument given in the second paragraph of the proof of Proposition~\ref{farberdranishnikov} applies word for word in the unrestricted case (replacing, of course, $\protect\operatorname{N}_{X \vee Y} (J_1, \ldots, J_s)$ by $\sum_{i=1}^s|J_i|-|\bigcap_{i=1}^sJ_i\cap J_O|$ in~(\ref{word4word}) and in the last line of that proof).}
\section{Other polyhedral product spaces}\label{othersec}
}%\textcolor{red}{Polyhedral product spaces have recently been the focus of intensive research in connection to toric topology and its applications to other fields. In this section we determine the higher topological complexity of polyhedral product spaces $Z(\{(X_i,\star)\},\mathcal{K})$ for which each factor space $X_i$ admits a {\it $\protect\operatorname{TC}_s$-efficient homotopy cell decomposition,} concept that is defined next.}
\medskip
}%\textcolor{red}{Recall that the spherical cone length of a path connected space $Y$, denoted here by $\protect\operatorname{cl}(Y)$, is the least nonnegative integer $c$ for which there is a \emph{length-$c$ homotopy cell decomposition} $(Y_0,\ldots,Y_c)$ of $Y$, that is, a nested sequence of spaces $Y_0\subseteq\cdots\subseteq Y_c$ so that $Y_0$ is a point (the base point of all the $Y_i$'s), $Y_c$ has the (based) homotopy type of $Y$ and, for $0\leq i<c$, $Y_{i+1}$ is the (reduced) cone of a (based) map $\pi_i\colon W_i\to Y_i$ whose domain $W_i$ is a finite wedge of spheres (of possibly different dimensions). In such a situation, we refer to $Y_i$, to $Y_i-Y_{i-1}$, and to $\pi_i$, respectively, as the $i$-th layer, the $i$-th stratum, and the $i$-th attaching map of the homotopy cell decomposition. If no such integer $c$ exists, we set $\protect\operatorname{cl}(Y)=\infty$. In these terms we say that $Y$ admits a $\protect\operatorname{TC}_s$-efficient homotopy cell decomposition when $\protect\operatorname{TC}_s(Y)=s\protect\operatorname{cl}(Y)$. The adjective ``$\protect\operatorname{TC}_s$-efficient'' is motivated by the following standard fact:}
\begin{lema}\label{decomposicionminimal}
}%\textcolor{red}{For a path connected space $X$, $\protect\operatorname{TC}_s(X)\leq s\protect\operatorname{cl}(X)$.}
\end{lema}
}%\textcolor{red}{The proof of Lemma~\ref{decomposicionminimal} given below makes use of products of homotopy cell decompositions, which is a standard construction in view of the finiteness condition on the number of cells in a given strata. For instance, the product of two homotopy cell decompositions $(Y_0,\ldots,Y_c)$ and $(Z_0,\ldots,Z_d)$, of $Y$ and $Z$ respectively, is the homotopy cell decomposition of $Y\times Z$ given by the sequence $(P_0,\ldots,P_{c+d})$ with $P_i=\bigcup_{j+k=i} Y_j\times Z_k$ and where we take the usual (Cartesian product) attaching maps.}
\begin{proof}[Proof of Lemma~$\ref{decomposicionminimal}$]
}%\textcolor{red}{Let $(X_0,\ldots,X_c)$ be a minimal homotopy cell decomposition of $X$. The product decomposition on $X^s$ has length $sc$, so the results follows from the fact that the sectional category of a fibration is bounded from above by the spherical cone length of its base.}
\end{proof}
}%\textcolor{red}{Known examples of spaces admitting a $\protect\operatorname{TC}_s$-efficient homotopy cell decomposition are:}
\begin{enumerate}
\item }%\textcolor{red}{Wedge sums of spheres (with the single exception of a wedge with a single summand given by an odd dimensional sphere).}
\item }%\textcolor{red}{Simply connected closed symplectic manifolds admitting a cell structure with no odd dimensional cells.}
\item }%\textcolor{red}{Configuration spaces on odd dimensional Euclidean spaces.}
\end{enumerate}
}%\textcolor{red}{All such examples satisfy, in addition, the equality $\protect\operatorname{TC}_s=\protect\operatorname{zcl}_s$, a condition that will be part of Theorem~\ref{pps} below. In particular, our result implies that the list of examples above can be extended to polyhedral product spaces constructed from the three types of spaces already listed.}
\begin{definition}
}%\textcolor{red}{For an $n$ tuple $}%\textcolor{blue}{\gamma=(c_1,\ldots,c_n)}$ of nonnegative integers, we define the }%\textcolor{blue}{\emph{$\gamma$-weighted} dimension} of an abstract simplicial complex $\mathcal{K}$ with vertices $[n]$ as}
$$
}%\textcolor{red}{}%\textcolor{blue}{\dim_{\gamma}}(\mathcal{K})=\protect\operatorname{max}\left\{ c_{i_1}+\cdots+c_{i_\ell} \,\, | \,\, 1\leq i_1<\cdots<i_\ell\leq n\mbox{ \hspace{.5mm}and \hspace{.3mm}} \{i_1,\ldots,i_\ell\}\in\mathcal{K}\right\}-1.}
$$
\end{definition}
}%\textcolor{red}{Theorem~\ref{TC_s(X) even} is generalized by:}
\begin{theo}\label{pps}
}%\textcolor{red}{Let $X=Z(\{(X_i,\star)\},\mathcal{K})\subseteq\prod_{i=1}^nX_i$ be the polyhedral product space associated to a family of pointed spaces $X_1,\ldots,X_n$, and an abstract simplicial complex $\mathcal{K}$ with vertices $[n]$. Assume that, for }%\textcolor{blue}{each} $i\in[n]$,}
\begin{itemize}
\item }%\textcolor{red}{$\protect\operatorname{TC}_s(X_i)=\protect\operatorname{zcl}(H^*(X_i;\mathbb{Q}))$, and}
\item }%\textcolor{red}{$X_i$ admits a $\protect\operatorname{TC}_s$-efficient $($and necessarily minimal, in view of Lemma~$\ref{decomposicionminimal})$ homotopy cell decomposition.}
\end{itemize}
}%\textcolor{red}{Then $X$ also satisfies the two hypothesis above and, in addition, $\,\protect\operatorname{TC}_s(X)=s(1+}%\textcolor{blue}{\dim_{\gamma}}(\mathcal{K}))\,$ where $}%\textcolor{blue}{\gamma=(\protect\operatorname{cl}(X_1),\ldots,\protect\operatorname{cl}(X_n))}$.}
\end{theo}
\begin{proof}[Proof of Theorem~$\ref{pps}$]
}%\textcolor{red}{For $i\in[n]$ let $(X_{0,i},X_{1,i},\ldots,X_{c_i,i})$ be a $\protect\operatorname{TC}_s$-efficient (and necessarily minimal, in view of Lemma~\ref{decomposicionminimal}) homotopy cell decomposition of $X_i$. By the homotopy invariance of the polyhedral product functor, we can assume that $X_i=X_{c_i,i}$ for all $i\in[n]$. Let $(P_0,\ldots,P_c)$ be the product homotopy cell decomposition on $\prod_iX_i$ where $c=\sum_ic_i$, and let $P'_i=P_i\cap X$ for $i\in[c]_0$. Note that $(P'_0,\ldots,P'_c)$ is a homotopy cell decomposition of $X$ for which $$P'_{1+}%\textcolor{blue}{\dim_{\gamma}}(\mathcal{K})}=P'_{2+}%\textcolor{blue}{\dim_{\gamma}}(\mathcal{K})}=\cdots=P'_c,$$ so $\protect\operatorname{TC}_s(X)\leq s(1+}%\textcolor{blue}{\dim_{\gamma}}(\mathcal{K}))$ in view of Lemma~\ref{decomposicionminimal}. To see that this is an equality (so that $(P'_0\ldots,P'_{1+}%\textcolor{blue}{\dim_{\gamma}}(\mathcal{K})})$ is $\protect\operatorname{TC}_s$-efficient), choose $1\leq i_1<\cdots<i_\ell\leq n$ with $\{i_1,\ldots,i_\ell\}\in\mathcal{K}$ and $c_{i_1}+\cdots+c_{i_\ell}=1+}%\textcolor{blue}{\dim_{\gamma}}(\mathcal{K})$, and note that
\begin{eqnarray*}
\protect\operatorname{TC}_s(X) & \geq &\protect\operatorname{TC}_s(X_{i_1}\times\cdots\times X_{i_\ell}) \\
&\geq& \protect\operatorname{zcl}_s(H^*(X_{i_1}\times\cdots\times X_{i_\ell};\mathbb{Q})) \\
&\geq& \sum_{j=1}^\ell\protect\operatorname{zcl}_s(H^*(X_{i_j};\mathbb{Q})) \\
&=& s\sum_{j=1}^\ell c_{i_j} \\
&=& s(1+}%\textcolor{blue}{\dim_{\gamma}}(\mathcal{K})).
\end{eqnarray*}
The second and third inequalities hold by Proposition~\ref{ulbTCn} and~\cite[Lemma~2.1]{CoFa11}, respectively, whereas the first inequality holds since, as explained in the first paragraph of the proof of Proposition~4 in~\cite{FelixTanre09}, $X_{i_1}\times\cdots\times X_{i_\ell}$ is (homeomorphic to) a retract of $X$. To complete the proof, note that, as above, $\protect\operatorname{zcl}_s(H^*(X;\mathbb{Q}))$ is bounded from above by $\protect\operatorname{TC}_s(X)$ and from below by $\protect\operatorname{zcl}_s(H^*(X_{i_1}\times\cdots\times X_{i_\ell};\mathbb{Q}))$---and that the last two numbers agree.}
\end{proof}
\begin{remark}\label{downtocat}{\em
}%\textcolor{red}{The methods of this section can be applied to describe the category of suitably efficient polyhedral products. For instance, without any restriction on the parity of the sphere dimensions $k_i$, any subcomplex $X$ of $\mathbb{S}(k_1,\ldots,k_n)$ has $\protect\operatorname{cat}(X^s)=s(1+\dim(\mathcal{K}_X))$. This is just an example of a partial (but very useful) generalization of~\cite[Proposition~4]{FelixTanre09}.}
}\end{remark}
|
\section*{References}
\providecommand{\newblock}{}
|
\section{Introduction \label{intro}}
Protein folding studies \cite{Dunbar} pose a great challenge to microfluidic mixers. Proteins are composed of chains of amino acids which take on complex three-dimensional (3D) structures to achieve a wide range of biological functions \cite{Berg,Roder}. The range of applications of protein folding in research and industry is wide and includes drug discovery, DNA sequencing, and molecular analysis or food engineering \cite{Gaudet,inf,Russell}. One of the most versatile methods of initiating the process of protein folding is using changes in chemical potential (e.g., changing the concentration of a chemical species) \cite{Park, Yamaguchi, Mansur}.
In this work, we consider a class of microfluidic mixers based on diffusion from (or to) a hydrodynamically focused stream. This type of mixer was initially proposed by Brody et al. ~\onlinecite{Brody}. A geometrical representation of such a mixer is shown in Figure~\ref{curgeoa} (here our optimized mixer of Ref. \onlinecite{POF1}). The basic features of the design are as follows: It is composed of three inlet channels and a common outlet channel, and the geometry has a symmetry with the center channel. Typically, a mixture of unfolded proteins and a chemical denaturant solution is injected through the center channel and exposed to background buffers (no denaturant) streams through the two side channels. The design goal is to rapidly decrease the denaturant concentration in order to rapidly initiate protein folding in the outlet channel \cite{Dunbar}. Since the publication of Brody et al., there have been significant advances on the design of these mixers \cite{hert,refmixer,yao07} including reduction in consumption rate of reactants, methods of detection, manufacturing and, perhaps most importantly, drastic reductions of the mixing time (i.e. the time required to reach a sufficiently low denaturant concentration).
We recently studied the optimization of the shape and flow conditions of a particular hydrodynamic focused microfluidic mixer\cite{POF1}. The objective was to improve the mixing time of the best mixer designs found in literature, which exhibited mixing times of approximately 1.0 $\mu$s. To this end, we introduced a mathematical model which computes the mixing time for a given mixer geometry and injection velocities. Then, we defined the corresponding optimization problem and solved it by considering a hybrid global optimization method \cite{ijnme,jota,jogo}. This approach was carried out and presented using both 2D and 3D models. To save on computational time, much of the optimization process was conducted using the 2D model. However, our earlier work also pointed out that certain important effects
(including the impact of upper and lower mixer walls and inertial effects on the velocity field) can be appreciated only with the 3D model. We therefore performed 3D model studies to analyze such effects. The optimized mixer generated by our approach achieved a mixing time of about 0.10 $\mu$s. The shape of this optimized microfluidic mixer, its concentration distribution and the concentration evolution of a particle in its central streamline are summarized in Figure ~\ref{res2D}. The optimized side and center channel injection velocities were $u_s=$5.2 m s$^{-1}$ and $u_c=$0.2 m s$^{-1}$, respectively.
The optimization problem studied in this previous work was identified as highly nonlinear\cite{redondo1}. Further, the process has many parameters which are difficult to know with great precision in experiments. Therefore, it is important to understand and quantify the stability of the device performance with respect to these parameters. This enables identification of key parameters and so guide experimental efforts. To this aim, we have analyzed the optimized mixer to study and quantify its robustness to parameters perturbations.
\begin{figure}
\centering
\psfig{file=sol2dconc-bw.eps,width=16cm}
\caption{Optimized mixer of Reference \onlinecite{POF1} and associated mixing performance: (a) top view representation of the half of the mixer shape (symmetry with respect to $x$=0 $\mu$m ) of the mixer with a superposed grey scale plot of the denaturant normalized concentration distribution $c$ at width $z$=0 $\mu$m and (b) the time evolution of the denaturant concentration of a particle flowing along the symmetry
streamline starting from $x$=0 $\mu$m, $y$=5 $\mu$m and $z$=0 $\mu$m. \cben{The parts of the mixer shape corresponding to the protuberances Prot$_{\rm up}$ and Prot$_{\rm lo}$, introduced in Section \ref{shape}, are highlighted in sub-figure (a)}.\label{res2D}}
\end{figure}
We previously presented a very simple sensitivity analysis~\cite{POF1}. That preliminary sensitivity study consisted of random perturbations of all the parameters by taking uniform variations within a range of \textcolor{black}{$[-\beta \%,+\beta \%]$} of their value. Results showed that the mixing time variations were of the same order as the normalized perturbations considered, suggesting the optimized solution was fairly stable. See Ivorra et al~\cite{POF1} for further information.
Here, we \cben{significantly increase }the scope of our sensitivity analysis. We quantify the impact of mixing time on the key design parameters of the mixer. The objective of our study is to provide recommendations and guidelines for the fabrication of the device introduced here. More precisely, we consider and study (i) Geometrical parameters defining the mixer shape: the angle defined by the channel intersection, the shape of the channel intersection, the width of the inlet and outlet channels, the mixer depth and possible irregularities in the symmetry of the shape;(ii) Central and side injections velocities; and (iii) Physical coefficients associated with the working fluid and the concentration thresholds of the mixing time definition.
In addition to those sensitivity analysis experiments, we also analyze the uniformity of the mixing time as a function of the inlet streamline location in the inlet channel. This mixing time uniformity analysis quantifies the robustness of the mixing time through the whole inlet flow, and helps place a statistical confidence on observed mixing times. In particular, it helps quantify the so-called wall effect (due to the no-slip condition at the mixer walls, resulting in low velocity values near the walls) on mixer performance.
The last two decades have seen a large number of microfluidic device designs and their use in a wide range of applications. Most, if not all, of these devices have performance specifications which are dependent on their geometry and flow control conditions (e.g., flow rates, pressure, inlet concentrations). Despite this, the systematic study of how performance depends on intentional or untintential design parameters is rarely if ever demonstrated. For this reason, we also offer the current work as a case study describing the significant challenge and complexity of determining design robustness for microfluidics.
This article is organized as follows: Section \ref{model} introduces the 3D model used to estimate the mixing times. Section \ref{uniform} describes the mixing time uniformity analysis and the results. Section \ref{sensa} presents the numerical experiments carried out to perform the extended sensitivity analysis and deduce major conclusions and design guidelines.
\section{Microfluidic mixer modeling \label{model}}
We consider the microfluidic mixer described in Section \ref{intro}. The geometry has two symmetry planes which we use to reduce the simulation domain to a quarter of the mixer. This reduced domain is denoted by $\Omega$, as depicted in Figure \ref{curgeoa}. The mixer shape is composed of interpolated surfaces, and the inlet velocities are described by a set of parameters denoted by $\phi$, detailed in Ref. \onlinecite{POF1}.
\begin{figure}[!ht]
\centering
\psfig{file=geom3d.eps,width=16cm}
\caption{Typical three-dimensional representation of the microfluidic mixer geometry. \cben{The mixer design hydrodynamically focuses a center inlet stream using two side inlets.} In dark gray we represent the domain $\Omega$ used for numerical simulations. The geometry's \cben{two} symmetry planes are also highlighted. \label{curgeoa}}
\end{figure}
We consider guanidine hydrochloride (GdCl) as the denaturant \cite{Dunbar,Kawahara}. We assume the mixer liquid flow is incompressible \cite{refmixer}. Thus, the flow velocity and the denaturant concentration distribution are approximated by using the steady configurations of the incompressible Navier-Stokes equations coupled with the convective diffusion equation. More precisely, we consider the following system \cite{massey,hert,refmixer}:
\begin{equation}
\left \{
\begin{array}{ll}
-\nabla \cdot ( \eta ( \nabla \mathbf{u} +( \nabla \mathbf{u})^{\top} ) -p\mathbf{I}) + \rho ( \mathbf{u}
\cdot \nabla) \mathbf{u} = 0& {\rm in} ~~\Omega,\\
\nabla \cdot \mathbf{u} =0 & {\rm in} ~~\Omega,\\
\nabla \cdot ( -D \nabla c) + \mathbf{u} \cdot \nabla c = 0& {\rm in} ~~ \Omega,
\end{array}
\right. \label{eql1}
\end{equation}
where $c$ is the denaturant normalized concentration distribution, $\mathbf{u}$ is the flow velocity vector (m s$^{-1}$), $p$ is the pressure field (Pa), $D=2\times 10^{-9}$ is the diffusion coefficient of the denaturant solution in the background buffer (m$^{2}$ s$^{-1}$), $\eta=9.8 \times 10^{-4}$ is the denaturant solution dynamic viscosity (kg m$^{-1}$ s$^{-1}$) and $\rho=1010$ is the denaturant solution density (kg m$^{-3}$ ).
System (\ref{eql1}) is completed by the following boundary conditions:
For the flow velocity $\mathbf{u}$:
\begin{equation}
\left \{
\begin{array}{ll}
\mathbf{u}=0 & \hbox{on } \Gamma_{w},\\
\mathbf{u}=- u_s {\rm para_1} \mathbf{n} & \hbox{on } \Gamma_{s},\\
\mathbf{u}=- u_c {\rm para_2} \mathbf{n} &\hbox{on } \Gamma_{c},\\
p=0 \hbox{ and } ( \eta ( \nabla \mathbf{u} +( \nabla \mathbf{u})^{\top} )) \mathbf{n}=0 &\hbox{on } \Gamma_{e},\\
\mathbf{n} \cdot \mathbf{u}=0 \hbox{ and } \mathbf{t} \cdot ( \eta ( \nabla \mathbf{u} +( \nabla \mathbf{u})^{\top} ) -p\mathbf{I}) \mathbf{n}=0 &\hbox{on } \Gamma_{a},\\
\end{array}
\right. \label{boundcu}
\end{equation}
where $\Gamma_{c}$, $\Gamma_{s}$, $\Gamma_{e}$, $\Gamma_{w}$ and $\Gamma_{a}$ denote the boundaries representing the central inlet, the side inlet, the outlet, the mixer walls and the symmetry plane, respectively; $u_s$ and $u_c$ are the maximum side and center channel injection velocities (m s$^{-1}$), respectively; ${\rm para_1}$ and ${\rm para_2}$ are the laminar flow profiles, which are equal to 0 in the inlet border and to 1 in the inlet center, of the side and central inlets, respectively \cite{massey}; and $(\mathbf{t},\mathbf{n})$ is the local orthonormal reference frame along the boundary.
For the concentration $c$:
\begin{eqnarray}
\left \{
\begin{array}{ll}
\mathbf{n} \cdot (-D \nabla c + c \mathbf{u})= - c_0 \mathbf{u} & \hbox{on } \Gamma_{c},\\
c =0 &\hbox{on } \Gamma_{s},\\
\mathbf{n} \cdot (-D \nabla c )=0 &\hbox{on } \Gamma_{e},\\
\mathbf{n} \cdot (-D \nabla c + c \mathbf{u})=0 &\hbox{on } \Gamma_{w} \cup \Gamma_{a},
\end{array}
\right. \label{boundcc}
\end{eqnarray}
where $c_0 =1$ is the initial denaturant normalized concentration in the center inlet.
In this work, the mixing time of a particular mixer $\phi$, denoted by $J(\phi)$ is defined as the time required to change the denaturant normalized concentration of a typical Lagrangian stream fluid particle situated in the symmetry streamline at depth $z=0$ $\mu$m from $\alpha \%$ to $\omega \%$. It is computed by:
\begin{equation}
J(\phi)= \int_{c^{\phi}_{\omega}}^{c^{\phi}_{\alpha}} \dfrac{{\rm d}y}{\mathbf{u}^{\phi}(y)}
\label{mtcf}
\end{equation}
where $\mathbf{u}^{\phi}$ and $c^{\phi}$ denote the solution of System (\ref{eql1})-\eqref{boundcc}, when considering the mixer defined by $\phi$; and $c^{\phi}_{\alpha}$ and $c^{\phi}_{\omega}$ denote the y-coordinate of points situated along the streamline defined by the intersection of the two symmetry planes $z=0$ $\mu$m and $x=0$ $\mu$m, i.e. the y-axis, where the denaturant normalized concentration $c^{\phi}$ is $\alpha$ and $\omega$, respectively. By default, we assume $\alpha=90$\% and $\omega=30 \%$.
The numerical model used to approximate the solutions of System (\ref{eql1})-(\ref{boundcc}) and to compute (\ref{mtcf}) was implemented by coupling Matlab scripts with COMSOL Multiphysics 3.5a models.
\section{Uniformity of the mixing time \label{uniform}}
We first analyze the non-uniformity of mixing times across the focused stream for our optimized mixer $\phi_o$. Indeed, as suggested in Refs. \onlinecite{Park2,yao07}, the mixing time can be measured not only in the symmetry streamline, situated on the (x,z)=(0,0) segment, but also in other streamlines. We are interested in the uniformity of mixing times, as protein states in these mixers are quantified \cben{experimentally} within a finite probe volume which integrates signal \cben{throughout} a volume in space within the mixing region. \cben{This measurement volume is} fed in principle by all streamlines of the center inlet channel.
We consider 100 streamlines, denoted by $(sl_{i,j})_{i,j=1}^{10}$, starting from a finite set of points, which are denoted by $\Sigma_{\Gamma_{c}}$, in $\Gamma_{c}$. Here, $\Sigma_{\Gamma_{c}}= \{ P_{(i,j)} | i=1,...,10 $ and $ j=1,...,10 \}$ where $P_{(i,j)}= (\frac{i}{10} 0.9\mu$m$, \frac{j}{8} 0.75 \mu$m$)$. In the previous definition, the maximum coordinate in the x-axis (i.e., 0.9 $\mu$m) has been selected in order to avoid particles too close to the wall $\Gamma_{w}$, and the maximum coordinate in the z-axis (i.e., 0.75 $\mu$m) has been chosen as a characteristic 1.5 $\mu$m depth of field for confocal microscope imaging (i.e., extent of the measurement volume)\cite{hert}. Those streamlines are numerically approximated by considering an explicit Euler scheme and the velocity vector $\mathbf{u}$ obtained by solving System \eqref{eql1}-\eqref{boundcc} \cite{inf}.
For each streamline $sl_{i,j}$, we compute the associated mixing times, denoted by $t_{sl_{i,j}}$, in a manner similar to Equation \eqref{mtcf}. More precisely, $t_{sl_{i,j}}$ is defined as the time required by a protein within a Lagrangian fluid particle to travel from $c^{sl_{i,j}}_{90}$ to $c^{sl_{i,j}}_{30}$, where $c^{sl_{i,j}}_{90}$ and $c^{sl_{i,j}}_{30}$ denote the points within $sl_{i,j}$ with a concentration of $90$\% and $30$\%, respectively. Next, we study the spatial distribution according to the streamline starting point in $\Sigma_{\Gamma_{c}}$, the maximum value, the mean value and the standard deviation of $(t_{sl_{i,j}})_{i,j=1}^{10}$. Furthermore, we also compute the weighted mixing time value of $sl_{i,j}$, denoted by $\overline{t_{sl_{i,j}}}$ and defined as
\begin{equation} \label{mmt}
\overline{t_{sl_{i,j}}}=\frac{\omega_{i,j} t_{sl_{i,j}}}{\sum_{i,j=1}^{10} \omega_{i,j}},
\end{equation}
where $\omega_{i,j}$ denotes the velocity of a particle in the streamline $sl_{i,j}$ at its initial position $x^{\rm init}_{(i,j)}$. This choice of weight coefficients reflects the fact that the probe volume used to measure experimentally the mixing time receive particles more frequently from streamlines with the highest velocities. The maximum and standard deviation values of those weighted mixing times $(\overline{t_{sl_{i,j}}})_{i,j=1}^{10}$ are also studied.
Furthermore, due to the fact that the depth of the mixer is 10 times larger than the minimum width of the center channel, the mixing time variations in the z-axis direction are negligible in comparison to the variations in the x-axis \cite{hert}. Thus, we perform a more extensive uniformity analysis along the x-axis, by considering 100 streamlines, denoted by $(sl_{i,z=0})_{i=1}^{100}$, in the plane $z=0$ starting from the set of points $P_{(i,j)}= (\frac{i}{100} 0.9\mu$m$, 0 \mu$m$)$ in $\Gamma_{c}$. The methodology is the same as that introduced previously. In this case, we also compute the evolution of both the mean value and standard deviation of $(t_{sl_{i,z=0}})_{i=1}^{k}$ and $(\overline{t_{sl_{i,z=0}}})_{i=1}^{k}$, with $k=1,...,100$. These results will be compared with the ones presented in Ref. \onlinecite{hert}.
Our study of mixing time uniformity yielded that the mean mixing time value obtained by considering $(t_{sl_{i,j}})_{i,j=1} ^{10}$ was 0.34 $\mu$s with a standard deviation of 0.17 $\mu$s. As expected, the maximum mixing time value was reached at the streamline $sl_{10,10}$ with a value of 1.43 $\mu$s.
The mixing times $t_{sl_{i,j}}$ and weighted mixing times $\overline{t_{sl_{i,j}}}$ of the considered streamlines $(sl_{i,j})_{i,j=1}^{10}$ are presented in Figure \ref{uan}-(a). As shown,
within 0.4 $\mu$m of the centerline, the mixing times vary between 0.1 $\mu$s and 0.5 $\mu$s, and this region accounts for 60\% of the detection events (i.e., considering the sum of the weight coefficients $(\sum_{i=1}^{4}\sum_{j=1}^{10} \omega_{i,j})/(\sum_{i,j=1}^{10} \omega_{i,j})$). In contrast, the near-wall region of [0.7, 0.9] $\mu$m of the centerline have mixing times between 1 and 1.43 $\mu$s, but these streamlines contribute to only 10\% of detection events (i.e., considering $(\sum_{i=7}^{10}\sum_{j=1}^{10} \omega_{i,j})/(\sum_{i,j=1}^{10} \omega_{i,j})$).
\begin{figure}
\begin{center}
\includegraphics [width=11cm]{3dslstat.eps}\\
\includegraphics [width=11cm]{2dslstat.eps}
\end{center}
\caption{Results obtained during the analysis of the mixing time non-uniformity for the optimized mixer: \textbf{a)} Mixing and weighted mixing times obtained according to the position $(x,z)$ in $\Gamma_c$ of the initial particle for the considered streamlines, \textbf{b)} Mixing and weighted mixing times obtained as a function of the position $x$ in $\Gamma_c$ of the initial particle and for the streamlines considered in the plane $z=0$. \cben{The weighted mixing time reflects the frequency of events (measurements of proteins along said streamline) as determined by the stream-line averaged velocity. The lower velocities near the wall yield longer mixing times but are less frequent.}}\label{uan}
\end{figure}
Next, the mixing times $t_{sl_{i,z=0}}$ and weighted mixing times $\overline{t_{sl_{i,z=0}}}$ across the streamlines $(sl_{i,z=0})_{i=1}^{100}$ are plotted versus spanwise streamline position in Figure \ref{uan}-(b). For these 100 streamlines, the mean mixing time computed by considering $(t_{sl_{i,z=0}})_{i=1}^{100}$ was 0.32 $\mu$s with a standard deviation of 0.16 $\mu$s. Again, we can observe that particles near the walls exhibit higher mixing times ($>$1 $\mu$s). However, these near-wall-slow-moving particles contribute only infrequently to probe volume detection events.
We note similar phenomena were reported in Ref. \onlinecite{hert}. However, the mixer presented in that work exhibited a mean mixing time, considering streamlines in the plane z=0, of 3.1 $\mu$s with a standard deviation of 1.5 $\mu$s. The maximum mixing time value was 10 $\mu$s, obtained for the streamline closer to the wall $\Gamma_{w}$. The optimized mixer design presented here therefore offers better mixing time uniformity leading to more consistent measurements and less scatter in measurement ensembles.
\section{Sensitivity analysis of the model parameters \label{sensa}}
We here present a study of the influence of key parameters of the model described in Section \ref{model} on mixer mixing time. We vary parameters individually, fixing the values of others to the corresponding value of the optimized mixer $\phi_o$. We note that, in our previous work, we explored the impact of simultaneous perturbations on the whole set of parameters on mixer performance\cite{POF1}. We here perform the more complete influence of individual perturbations on the mixing time. We believe such individual parameter perturbation analyses are also more useful to designers in identifying key parameters and methods for fabrication. We consider the following percent variation function:
\begin{equation}
E(\phi_{p})=100\dfrac{| J(\phi_o)-J(\phi_{p})|}{J(\phi_o)}.\label{err}
\end{equation}
where $\phi_{p}$ represents the perturbed mixer.
The parameters analyzed can be classified in three categories: (i) geometrical parameters defining the mixer shape; (ii) central and side injections velocities; and (iii) physical coefficients associated with the denaturant solution and the concentration threshold in the mixing time definition.
\subsection{Geometrical parameters}
In the following computational experiments, we analyze the variation on the mixing time due to changes in: (i) the angle defined at the channel intersection;
(ii) the shape of the channel intersection; (iii) the width of the inlet and outlet channels; (iv) the mixer depth; and (v) perturbation in the symmetry of the mixer shape.
\subsubsection{Inlet intersection angle}
First, we study the angle between the x-axis and the mixer side channel, denoted by $\theta$. The optimized value $\theta=\pi/5$ is varied from 0 up to $2\pi/5$ by considering 50 equally spaced intermediate values (i.e, we perform 50 evaluations of our model). A geometrical representation of those variations is showed in Figure~\ref{angf}.
\begin{figure}[!ht]
\centering
\psfig{file=angf.eps,width=16cm}
\caption{\cben{Three mixer} shapes \cben{for inlets} intersection angles of $\theta=0$ (light grey), $\theta=\pi/5$ (white) and $\theta=2\pi/5$ (dark grey). Only the area where the shape changes is \cben{shown}. \label{angf}}
\end{figure}
Perturbations on $\theta$ have generated a mean variation in the mixing time of 3\%. Figure~\ref{angr} gives a graphical representation of the obtained results. As we can observe on this plot, the maximum variation was around 15\% and was obtained for $\theta=2\pi/5$. Furthermore, the variation was less than 4\% for angles lower than $\pi/3$, and grew up exponentially after that value. This suggests that the angle is not a sensible parameter for the mixer performance.
\begin{figure}[!ht]
\centering
\psfig{file=angr.eps,width=16cm}
\caption{\cben{Percent variation of mixing time as a function of deviation from the optimal angle $\theta$ value (denoted by X=$\pi$/5, c.f. Figure \ref{angf}). Mixing time is relatively insensitive to small errors in angle of the side channel. The distribution shows the strongly non-linear dependence of mixing time on geometry.} \label{angr}}
\end{figure}
\subsubsection{Shape of the channel intersection \label{shape}}
We now study the impact of the shape of the area where the three inlets and the outlet intersect. The shapes allowed by our model are built by considering Beziers curves and describe a 'bubble' (also called protuberance) invading the central and side inlets from the upper corner (according to y-axis) and a protuberance invading the outlet and side inlets from the lower corner. These protuberances are defined according to a restriction (due to a convenient lithographic and plasma etching limitation) of a minimum channel width of 1$\mu$m. For the sake of simplicity, those bubbles are only described by two scalar numbers Prot$_{\rm up}$ and Prot$_{\rm lo}$ in $[0,1]$, where 0 corresponds to the minimum bubble shape and unity is the maximum bubble shape of the upper and lower corner, respectively, as allowed by the model. The optimal shape corresponds to Prot$_{\rm up}$=0.8 and Prot$_{\rm lo}$=0.7. \cben{The parts of the mixer shape corresponding to Prot$_{\rm up}$ and Prot$_{\rm lo}$ are presented in Figure \ref{res2D}.} A geometrical representation of the minimum, maximum and optimal shapes of the protuberances is given in Figure \ref{cornf}.
\begin{figure}[!ht]
\centering
\psfig{file=cornerf.eps,width=16cm}
\caption{\cben{Mixer shapes highlighting the range of mixer shapes we explored. In both cases, the optimal shape is shown as a dark line. Shown are detailed views of the shape of the channels intersection near the \textbf{a)} upper (denoted by Prot$_{\rm up}$) and \textbf{b) lower (denoted by Prot$_{\rm lo}$)} corners of the intersection region.} The dark gray zones correspond to the domain, between the shape of the maximum and the minimum protuberance allowed by the model considered here (according to a minimum channel width of 1$\mu$m).
\label{cornf}}
\end{figure}
This experiment consisted of computing the mixing time of the mixer generated by considering all the possible combination of values of Prot$_{\rm up}$ and Prot$_{\rm lo}$ in $[0,1]$ with a grid step size of 0.1. This required 121 evaluations of our model. The variation of the mixing time according to analyzed values of Prot$_{\rm up}$ and Prot$_{\rm lo}$ is presented in Figure \ref{cornr} and values are reported in Table \ref{tcornr}. As shown by both the figure and the table, for values of Prot$_{\rm up}$ and Prot$_{\rm lo}$ lower than 0.5, the mixing time dramatically increased from 50\% up to 250\%. This indicates that a minimum protuberance in both upper and lower corners should be considered in order to obtain an efficient mixing time. Furthermore, when Prot$_{\rm up}$ and Prot$_{\rm lo}$ were greater than 0.5, the variation in mixing time was moderated and was lowered by 22\%, which can be considered as a reasonable value. In addition to those first results, we see that the impact on the mixing of Prot$_{\rm up}$ was greater than Prot$_{\rm lo}$. For instance, by decreasing the parameter Prot$_{\rm lo}$ from 1 to 0 and fixing the value of Prot$_{\rm up}$=1, we have generated mixing time variations up to 50\%, whereas by decreasing the parameter Prot$_{\rm up}$ and fixing Prot$_{\rm lo}$=1 we have obtained a maximum 20\% variation of mixing time. This result is consistent with the fact that the length of the lower corner is much larger than that of the upper (see Figure \ref{cornf}), thus, its influence on the mixing time is expected to be greater.
\begin{figure}[!ht]
\centering
\psfig{file=corner.eps,width=16cm}
\caption{\cben{Percent variation of} mixing time for the optimal shape (represented by X) for the protuberance magnitudes considered and described in Section \ref{shape}. Protuberance parameter values Prot$_{\rm up}$ and Prot$_{\rm lo}$ each vary from 0 to unity with a grid step size of 0.1. \cben{The details of the protuberance shape of the side channels is an important feature.} \label{cornr}}
\end{figure}
\begin{table}
\begin{center}
\caption{Percentage variation in mixing time of the optimal design value and considering the protuberances described in Section \ref{shape}. Here protuberances parameters Prot$_{\rm up}$ and Prot$_{\rm lo}$, each vary from 0 to unity with a grid step size of 0.1. The optimal shape (-) is obtained with
Prot$_{\rm up}$=0.8 and Prot$_{\rm lo}$=0.7. \label{tcornr}}
\vspace{0.5cm}
\begin{tabular}{|c||rrrrrrrrrrr|}
\hline
\backslashbox{\textbf{Prot$_{\rm lo}$}}{\textbf{Prot$_{\rm up}$}} & 0.0 & 0.1 & 0.2 & 0.3 & 0.4 & 0.5 & 0.6 & 0.7 & 0.8 & 0.9 & 1.0\\
\hline
\hline
0.0 & 251 & 220 & 192 & 167 & 144 & 126 & 106 & 91 & 79 & 71 & 89\\
0.1 & 218 & 192 & 167 & 145 & 125 & 107 & 90 & 76 & 65 & 56 & 75\\
0.2 & 186 & 163 & 142 & 123 & 105 & 89 & 75 & 62 & 51 & 42 & 61\\
0.3 & 155 & 136 & 118 & 102 & 86 & 73 & 60 & 48 & 39 & 30 & 48\\
0.4 & 125 & 110 & 95 & 82 & 69 & 57 & 47 & 37 & 28 & 20 & 36\\
0.5 & 98 & 87 & 74 & 64 & 52 & 43 & 34 & 26 & 18 & 11 & 20\\
0.6 & 74 & 65 & 54 & 45 & 37 & 29 & 22 & 15 & 9 & 3 & 15\\
0.7 & 51 & 43 & 37 & 29 & 23 & 17 & 11 & 5 & - & 4 & 8\\
0.8 & 30 & 25 & 19 & 14 & 9 & 5 & 1 & 3 & 7 & 11 & 9\\
0.9 & 19 & 7 & 3 & 2 & 3 & 7 & 9 & 12 & 14 & 17 & 9\\
1.0 & 8 & 7 & 6 & 10 & 13 & 14 & 16 & 19 & 20 & 21 & 14\\
\hline
\end{tabular}
\end{center}
\end{table}
\cben{From the previous results, we conclude that the mixing time is sensitive to the shape of these protuberances.}
\subsubsection{Channel width \label{widthe}}
We are here interested in estimating the impact of the inlets and outlet widths on the mixing time (i.e., the minimum width of these channels where the flow they carry first interact with the neighbouring streams). This study is interesting as the mixer design and general shape can be scaled geometrically and inserted into different devices. We note that the channel widths were fixed during the optimization process in Ref. \onlinecite{POF1} and were set to values suited for the mixer implementation and validation studies, as the one carried out in Ref. \onlinecite{hert}.
We considered a width denoted by $w_{c}$ $\in$ [1$\mu$m,4$\mu$m] for the central inlet, a width denoted by $w_{s}$ $\in$ [1$\mu$m,4$\mu$m] for the side inlets and a width denoted by $w_{o}$ $\in$ [2$\mu$m,18$\mu$m] for the outlet. The original optimized shape exhibited $w_{c}=2\mu$m, $w_{s}=3\mu$m and $w_{o}=10\mu$m. All possible configurations of channel widths were tested by considering a mesh of step size of 1 $\mu$m for each width, which represents a total of 272 evaluations of our model. Representations of the mixer shape with all channel widths set to their maximum or minimum values, are depicted by Figure \ref{widf}.
\begin{figure}[!ht]
\centering
\psfig{file=widf.eps,width=16cm}
\caption{Shape of the mixer for lengths of the channels set to their maximum (continuous line for $x>$0 m) values and minimum values (continuous line for $x<$0 m) . The dash-dot line corresponds to the axis $x=0$ \cben{(and the y-z symmetry plane)}. The optimal mixer shape is represented by a dashed line. We note \cben{the choice of parameterization of the mixer shape (including side inlet channel width) determines the position of the region corresponding to the channel intersection. Our geometry variations therefore considered a wide range of shapes and relative channel lengths.}
\label{widf}}
\end{figure}
In order to check the importance of each channel width on the mixing time regarding all possible configurations of other width, we considered percent variations denoted by $WE$ and the mean evolution of the mixing time $MET$ according to each width. Both processes are explained below. We illustrate the process of computing $WE$ and $MET$ in the case of $w_{c}$. This approach can be extended to $w_{s}$ and $w_{o}$.
The value $WE_{w_{c}}(j,k)$ represents a measure of the variation of the mixer mixing time according to changes in $w_{c}$ when other widths are fixed to $w_{s}=j$ and $w_{o}=k$, and is given by
\begin{equation}
WE_{w_{c}}(j,k)=100 \dfrac{1}{4} \sum_{i=1}^{4} \dfrac{| J(\phi_{i,j,k})- \hbox{mean}_{i} J(\phi_{i,j,k})|}{\hbox{mean}_{i} J(\phi_{i,j,k})}, \label{werr1}
\end{equation}
where $\phi_{i,j,k}$ denotes the mixer obtained by considering $w_{c}=i,w_{s}=j$, $w_{o}=k$ and the other parameters set to the optimal values and $\hbox{mean}_{i} J(\phi_{i,j,k})$ denotes the mean value of the mixing time obtained by varying only $i$. We compute $WE_{w_{c}}(j,k)$ for $j=1,..,4$, $k=2,...,18$ and report its mean, minimum and maximum values according to $j$ and $k$. Those results are reported in Table \ref{widt}.
The value $MTE_{w_{c}}(i)$ corresponds to the mean values of the mixing times $J(\phi_{i,j,k})$ obtained when considering $j=1,..,4$ and $k=2,..,18$. The evolution of $MTE_{w_{c}}$, $MTE_{w_{c}}$ and $MTE_{w_{c}}$ are depicted in Figure \ref{widr}.
\begin{figure}[!ht]
\centering
\psfig{file=widr.eps,width=16cm}
\caption{Dependence of mean values of mixing times as defined in Section \ref{widthe} as a function of inlet \cben{and outlet} channel widths. Shown are plots \textbf{a)} $MTE_{w_{c}}$, \textbf{b)} $MTE_{w_{s}}$ and \textbf{c)} $MTE_{w_{o}}$
\cben{which correspond to the mean mixing times obtained when fixing the value of $w_{c}$, $w_{s}$ and $w_{o}$, respectively, and let the other widths vary. Width variations have a moderate effect on mixing times.}
\label{widr}}
\end{figure}
As we can observe in Table \ref{widt}, the most sensitive widths are the side inlets and central inlet with a mean mixing time variation of about 65\%. This result is expected, since those inlets carry the denaturant solution and buffer flows, and thus affect the amount of injected products. Significant changes to these inlet geometries should be accompanied by changes in inlet velocities and performing a new optimization process as in Ref. \onlinecite{POF1}. For example, we hypothesize that variations which aim to preserve flow rate ratios should be explored first. On the other hand, outlet widths in the interval [2,13] $\mu$m (from the optimal value of 6 $\mu$m) will affect mixing time variation by only 7\%. Hence, we conclude that such errors on width have only a slight to moderate effect on mixing times. Furthermore, regarding Figure \ref{widr}, we see that the mean mixing time is lower when considering values of $w_{c}$ and $w_{s}$ in the interval [2$\mu$m,4$\mu$m] and $w_{o}$ $\in$ [1$\mu$m,12$\mu$m]. Moreover, we remark that configurations with smaller inlets and bigger outlet are the worst from an efficiency point of view.
\begin{table}
\begin{center}
\caption{Maximum, minimum and mean values of the mixing time percent variation named $WE$, defined in Section \ref{widthe}, for the widths $w_c$, $w_s$ and $w_o$.}
\label{widt}
\vspace{0.5cm}
\begin{tabular}{|c|ccc|}
\hline
\backslashbox{\textbf{Width}}{$\mathbf{WE}$} &Mean &Min &Max\\
\hline
\textbf{$\mathbf{w_c}$} &60 &27 &103 \\
\textbf{$\mathbf{w_s}$} &68 & 96 &118 \\
\textbf{$\mathbf{w_o}$} &7 &1 &57\\
\hline
\end{tabular}
\end{center}
\end{table}
\subsubsection{Mixer depth}
Next, we analyzed the effects of the mixer depth (in Z-direction). Imperfections in micro-fabrication of these mixers can result in depth variations of approximately $\pm 1\mu$m \cite{refmixer}. We thus computed the mixing time for mixers generated by considering the set of parameter $\phi_o$ and depths of $8,9,11$ and $12$ $\mu$m. The resulting mixing times (and their associated percent variation regarding the mixing time of the original mixer with a depth of 10 $\mu$m) were 0.14 $\mu$s (34\%), 0.12 $\mu$s (13\%), 0.10 $\mu$s (6\%), and 0.09 $\mu$s (13\%), respectively.
As shown by these results, perturbations of $\pm 1$ $\mu$m generate reasonable percent variations in the weighted mean mixing time between 6\% and 13\%. As described previously, this indicates that errors in the mixer depth due to manufacturing processes do not strongly affect mixing performance for these relatively deep ($\sim$10 $\mu$m) mixers. Note that the highest channel depth yields the lowest mixing time (0.09 $\mu$s for a depth equal to 12 $\mu$m versus 0.14 $\mu$s for the 8 $\mu$m depth). This result is expected, as the so-called wall effect (i.e., where the no-slip condition at the top wall results in low velocity values near the wall and near the corner where the X-Y plane meets the Y-Z plane) reduces the mixing performance near the mixer walls. Again, we see that the optimal mixer design (minimum mixing time) is influenced strongly by changes in manufacturing process (namely in achieving high aspect ratio features with deep reactive ion etching). Mixer designs with relatively high channel-depth-to-feature width ratios yield optimal results. In our study, the minimum channel width (near $y$ = -1.5 $\mu$m) was 1.1 $\mu$m.
\subsubsection{Shape Symmetry \label{ssym}}
The last geometrical aspect analyzed during this work is the impact of perturbations in the symmetry of the mixer according to the plane $x=0$ (including nonsymmetric injections velocities) on the mixer characteristics.
To this end, we considered the right half (versus quarter) of the geometry. We then randomly generated 100 nonsymmetric mixers by considering perturbations of the parameters from 0.5\% up to 50\% of the left side (respecting to $x=0$) of the mixer shape and by keeping the right side of the mixer shape to its optimal value. These mixers were then classified according to the deviation observed between the streamlines starting from $(x=0,y=0,z=0)\mu$m of the symmetric and nonsymmetric mixers at the time when the non perturbed symmetric streamline reach $\Gamma_e$. According to this classification, we then computed the mean mixing time for each category and compared it to the optimized mixer mixing time by considering the percent variation formula (\ref{err}). Deviations in the intervals [0,0.3] $\mu$m, [0.3,0.6] $\mu$m, [0.6,0.9] $\mu$m, [0.9,1.2] $\mu$m, [1.2,1.5] $\mu$m and greater than 1.5 $\mu$m generated mean mixing time percent variation of 14\%, 64\%, 114\%, 237\%, 328\% and 542\%, respectively.
As we can observe from those data, for deviations below 0.3 $\mu$m, which correspond to parameter perturbations lower than 10\% in the symmetry of shape and injection velocities, the order of the mixing time was conserved with a mixing time variation of 14\%. For greater deviations, the mixing time was dramatically increased from 64\% up to 500\%. Thus, we recommend normalized symmetry errors of less than 10\% be achieved to ensure a mixing time close to the optimal value.
\subsection{Flow injection velocities}
We studied the influence of injection velocities on mixing time. The optimized injection velocities obtained in Ref \onlinecite{POF1} were $u_s=$5.2 m s$^{-1}$ and $u_c=$0.2 m s$^{-1}$ (equivalent to a ratio $u_c/u_s$=0.0389). For this, we considered the optimized mixer and varied its side injection velocity $u_s=$ from 0.5 m/s to 9.5 m/s, with a step size of 0.5 m/s. Then, we chose $u_s$ in order to achieve $u_c/u_s$ ratios in the set $\{25,50,75,100,250\}$\% (considered as typical values) of the optimal ratio. This part of our study required a total of 95 evaluations of our model.
The results are summarized in Table \ref{velt} and Figure \ref{velr}. We can see that for $u_s$ $\in$ [4,9] m/s and $u_c$ $\in$ [0.12,0.88] m/s (i.e., a ratio $u_c/u_s$ of [0.0292,0.0973]), the mixing time has exhibited variations lower than 10\%. This suggests our mixer should be robust to small perturbations in the injection velocities. In particular, the velocity of the central inlet flow should be in the interval [0.1,0.8]m/s to obtain a reasonable mixing time. Moreover, from those results we can deduce that if the ratio and/or $u_s$ are too small, the mixing time is drastically increased (more than 1000\%). In fact, the mixer performance becomes similar to the one achieved in a previous study (see Ref. \onlinecite{ijnme}).
\begin{table}
\begin{center}
\caption{Variation in percent of the mixing time value (relative to the value of the optimal mixer indicated by -) as a function of values of the side injection velocity $u_s$ (m/s) and the injection ratio $u_c/u_s$.\label{velt}}
\vspace{0.5cm}
\begin{tabular}{|c||rrrrr|}
\hline
\backslashbox{\textbf{$\mathbf{u_c}$ (m/s)}}{\textbf{ratio}} & 0.0097 & 0.0195 & 0.0292 & 0.0389 & 0.0973\\
\hline
\hline
0.5&26317&5444.8&1542.7&668.7&126.6\\
1&6136.1&1518.6&278.4&134.2&40.5\\
1.5&2776.7&667.4&109&56.8&20.8\\
2&1631.1&388.7&57.4&30.5&12.4\\
2.5&1083.7&229.5&34.6&18&7.8\\
3&775.2&142.6&22.2&11.1&4.9\\
3.5&583.3&91.1&14.8&6.6&3\\
4&456.8&59.7&9.8&3.6&1.5\\
4.5&368.9&41.6&6.2&1.8&0.4\\
5&304.6&30&3.6&-&0.5\\
5.5&256.6&22&1.6&1.4&1.3\\
6&219.5&16.4&1&2.4&1.9\\
6.5&190.4&12.2&1.1&3.2&2.4\\
7&166.8&9&2.1&3.9&2.9\\
7.5&147.5&6.4&2.9&4.5&3.3\\
8&131.6&4.4&3.6&5&3.6\\
8.5&118.1&2.8&4.2&5.4&3.9\\
9&106.7&5&4.7&5.7&4.2\\
9.5&97&9.2&9.4&10&42\\
\hline
\end{tabular}
\end{center}
\end{table}
\begin{figure}[!ht]
\centering
\psfig{file=velr.eps,width=16cm}
\caption{Percent variation of the mixing time relative to the optimal mixer (represented by X) as a function of variations of the side injection velocity $u_s$ (m/s) and the injection ratio $u_c/u_s$. Precise control of flow rates are crucial in mixing experiments. \label{velr}}
\end{figure}
We conclude that accurate control of flow rates is crucial to achieving fast mixing. We recommend that flow rates be analyzed by experimental quantitation of inlet velocities using, for example, micron-resolution particle image velocimetry (as performed by Hertzog et al. in Ref. \onlinecite{refmixer}).
\subsection{Thermophysical Parameters}
We next studied the stability of the mixing time of the optimized mixer to changes in the thermophysical coefficients of the denaturant solution or in the concentration values needed to control the folding process. In physical experiments, these changes may result from uncertainties in conditions or solution properties (temperature, pressure, dilution, etc.)\cite{coef1} the following sections summarize.
\subsubsection{Denaturant Solution Parameters \label{denat}}
We chose for our work guanidine hydrochloride (GdCl) as a typical denaturant \cite{Dunbar,Kawahara} described by the following parameters: diffusivity in background buffer of $D=2\times 10^{-9}$ m$^{2}$ s$^{-1}$, denaturant solution dynamic viscosity of $\eta=9.8 \times 10^{-4}$ kg m$^{-1}$ s$^{-1}$ and mass density of $\rho=1010$ kg m$^{-3}$.
The thermophysical properties of GdCl solutions vary with concentration and ambient temperature: consistent with the experimental work of Refs. \onlinecite{coef2}, \onlinecite{coef3} \cben{and \onlinecite{coef4}}, (i) the density of the GdCl \cben{solutions} can vary within [1000,1700] kg m$^{-3}$; (ii) its viscosity can vary in [4,11]$\times 10^{-4}$ kg m$^{-1}$ s$^{-1}$; and (iii) its \cben{diffusivity} can vary in [1.9,13]$\times 10^{-9}$ m$^{2}$ s$^{-1}$.
We considered the impact of these parameter variations on mixing time. We varied each parameter within the aforementioned intervals using seven equispaced values. All possible configurations of parameters values were studied, which represents a total of 343 evaluations of the model. Then, similar to the work presented in Section \ref{widthe}, we computed the mean evolution of mixing time for each parameter value format both its lower and upper bound, and while varying the remaining coefficients to all their possible values.
The variations of the mean mixing time of the diffusion, density and viscosity are presented in Figure \ref{coefr}. We see that the diffusion was the most sensitive parameter, and can increase mixing time by up to 0.3$\mu$s. The other two coefficients maintained the mean mixing time close to 0.1$\mu$s. We note all of these values reasonable for the design as the order of mixing time is preserved. We further note increasing viscosity and decreasing diffusivity and density result in lower mixing time. The effect of decreasing diffusivity may at first seem counterintuitive, but mixing time is the result of a geometry- and flow-rate-dependent convective diffusion process. For example, high diffusivity can result in significant decreases of denaturant concentration within the early-focusing region of the center jet, where fluid velocities are still too low to stretch material interfaces and decrease diffusion lengths of the center jet. The latter effect is discussed by Hertzog et al. (2004) (e.g., see Figure 2 of that reference).
\begin{figure}[htb]
\centering
\psfig{file=coefr.eps,width=16cm}
\caption{Mean mixing time (in s) \cben{obtained as a function of denaturant solution} \textbf{a)} diffusivity, \textbf{b)} mass density and \textbf{c)} dynamic viscosity, and fixing the other two parameters. \label{coefr} \label{cefrr}}
\end{figure}
\subsubsection{Concentration threshold \label{econc}}
Finally, we characterize the sensitivity of the mixer to the maximum and minimum denaturant concentration values of our mixing time (see (4)). The original mixer was designed to trigger unfolding for a concentration reduction of 60\%. We here consider mixing times for denaturant concentration reductions ranging from 10\% and 92\%. To this end, for a particular threshold value denoted by $\gamma$, we identify $\alpha_\gamma$ and $\omega_\gamma$ such that $\alpha_\gamma-\omega_\gamma=\gamma$ and they produce the minimum mixing time value
\begin{equation}
J_{\gamma}(\phi)= \int_{c^{\phi}_{\omega_\gamma}}^{c^{\phi}_{\alpha_\gamma}} \dfrac{{\rm d}y}{\mathbf{u}^{\phi}(y)},
\label{mtcf2}
\end{equation}
where $c^{\phi}_{\alpha_\gamma}$ and $c^{\phi}_{\omega_\gamma}$ denote the Y-coordinates of the points situated along the streamline defined by the intersection of the two symmetry planes $z=0$ $\mu$m and $x=0$ $\mu$m, i.e. the y-axis, where the denaturant normalized concentration is $\alpha_\gamma$ and $\omega_\gamma$, respectively.
Results are presented in Figure \ref{concr}. The mixer exhibited mixing times lower than 0.4$\mu$s for up to a reduction of 90\%. We conclude that it is a robust design as the maximum reduction allowed by the flow rate ratios in this mixer was 92\%. For a 70\% denaturant concentration reduction, we observed a 0.1$\mu$s mixing time. The latter can be compared to the mixer of Ref. \onlinecite{POF1} which showed mixing times of 1$\mu$s for the same denaturant concentration reduction\cite{yao07}.
\begin{figure}[htb]
\centering
\psfig{file=concr.eps,width=16cm}
\caption{Mixing time (in s) as a function of the percentage reduction in concentration as defined in Section \ref{econc}. \cben{The concentration values inherent to the definition of mixing have significant effect on mixing time. Note the} maximum concentration reduction allowed by complete mixing far downstream is $92\%$. \label{concr}}
\end{figure}
\section{Conclusions}
We presented a detailed study of the robustness and performance of a microfluidic mixer design first presented in Ref. \onlinecite{POF1}. The mixer is for protein folding dynamics studies and can be used to initiate the folding process of a protein by diluting a local denaturant concentration in a short time interval. In Ref. \onlinecite{POF1}, we used a 3D numerical model and showed the ideal mixer shape, flow control parameters, and expected thermophysical parameters, which resulted in a mixing time of 0.10 $\mu$s. We here studied the robustness of this mixer relative to expected variations of these major design features. In particular, we studied (i) the uniformity of the mixing time through the center inlet flow and (ii) the sensitivity of the mixing time with respect to key mixer parameters. The uniformity study showed that mixing time is quite stable throughout the majority of the inlet stream (up to a distance from the walls of 0.4 $\mu$m). With respect to design robustness, we found that the details of the mixer design in the region near the channel intersections are essential to the performance, i.e., the shape the minimum channel widths near this inlet, the inlet flow velocity ratio, and possible (unwanted) asymmetries in the fabrication. Other factors such as inlet channel angles, mixer depth (above a certain minimum), fluid properties, and denaturant concentration thresholds for protein folding have significantly weaker effect on mixing time.
Our analyses \cben{may} provide a guide to designers and fabricators of protein folding mixer devices, and can be used to evaluate trade-offs between manufacturing quality, precision of flow control, and expected performance. Our work also serves as a case study associated with the general design and performance prediction of microfluidic devices, and may serve as a guide to designing complex and optimal fluidic systems. In the least, the work highlights the complexity and importance of predicting and managing uncertainty in the performance of microfluidic systems.
In Table \ref{tabres}, for each parameter, we provide the mean, maximum and standard deviation values of the mixing time percentage variation regarding the optimal mixer obtained with this sensitivity analysis.
\begin{table}
\begin{center}
\caption{Summary of major findings of our sensitivity analysis: Mean (\textbf{Mean}), Standard Deviation (\textbf{Dev}) and worst-case Maximum (\textbf{Max}) values of the mixing time in percentage of base design. For the sake of completeness, we also report the optimal value of each parameter (\textbf{Opt}) as well as the range of the considered values (\textbf{Range}). \label{tabres}}
\vspace{0.5cm}
\begin{tabular}{|c|c|c|c|c|c|}
\hline
\textbf{Parameter}& \textbf{Opt} & \textbf{Range} & \textbf{Mean} & \textbf{Dev} & \textbf{Max}\\
\hline
\multicolumn{6}{|c|}{\textbf{Intersection Angle}}\\
\hline
$\theta$ & $\pi$/5 & [0,2$\pi$/5] &3 & 3 &16 \\
\hline
\multicolumn{6}{|c|}{\textbf{Channel Intersection}}\\
\hline
Prot$_{up}$ &0.8 & [0,1] &21& 17& 51 \\
Prot$_{low}$&0.7 &[0,1]& 30& 26 & 79\\
\hline
\multicolumn{6}{|c|}{\textbf{Channel Width}}\\
\hline
$\omega_c$ ($\mu$m) &2 & [1,4] & 50 & 67 & 150 \\
$\omega_s$ ($\mu$m) &3 & [1,4] & 53 & 57 & 181 \\
$\omega_o$ ($\mu$m) &10 & [2,18] & 28 & 41 & 101 \\
\hline
\multicolumn{6}{|c|}{\textbf{Mixer Depth}}\\
\hline
Depth ($\mu$m) & 10 & [8,12] &13 & 13&34 \\
\hline
\multicolumn{6}{|c|}{\textbf{Symmetry}}\\
\hline
Symmetry (\%) & 0 & [0.5,50] &216 & 196&542 \\
\hline
\multicolumn{6}{|c|}{\textbf{Injection velocities}}\\
\hline
$u_c$ (m s$^{-1}$) &0.2 & [0.005,0.92] & 68 & 133 & 304 \\
$u_s$ (m s$^{-1}$) &5.2 & [0.5,9.5] & 51 & 153 & 669 \\
\hline
\multicolumn{6}{|c|}{\textbf{Physical coefficients}}\\
\hline
D (m$^2$ s$^{-1}$) &2 $\times 10^{-9}$ & [1.9,13]$\times 10^{-9}$ & 155 & 198 & 307 \\
$\nu$ (kg m$^{1}$ s$^{-1}$) &9.8 $\times 10^{-4}$ & [4,11]$\times 10^{-4}$ & 4 & 4 & 11 \\
$\rho$ (kg m$^{-3}$) &1010 & [1000,1700] & 2 &2 & 6 \\
$\gamma$ (\%) &60 & [10,92] & 129 & 271 & 1282 \\
\hline
\end{tabular}
\end{center}
\end{table}
\begin{acknowledgments}
This work was carried out thanks to the financial support of the Spanish “ Ministry of Economy and Competitiveness” under projects MTM2011-22658 and TIN2012-37483; the “Junta de Andaluc\'{\i}a” and the ”European Regional Development Fund (ERDF)” through projects P10-TIC-6002, P11-TIC-7176 and P12-TIC301; and the research group MOMAT (Ref. 910480) supported by "Banco Santander" and "Universidad Complutense
de Madrid". Juana López Redondo is a fellow of the Spanish ”Ram\'on y Cajal” contract program, co-financed by the European Social Fund.
\end{acknowledgments}
|
\section{Introduction}
Spintronics, which concerns the effects on transport due to the coupled spin and charge degrees of freedom of the electron, has raised intense interest due to its broad industrial applications and theoretical challenges.\cite{Wolf_2001, Zutic_2004, Awschalom_2007, Chappert_2007, Bader_2010} These magnetic transport properties underlie giant, tunneling, and colossal magnetoresistance (GMR, TMR, and CMR),\cite{Baibich_1988,Binasch_1989,Miyazaki_1995,Moodera_1995, Jin_1994} tunneling anisotropic magnetoresistance (TAMR),\cite{Gould_2004, Saito_2005} and the anomalous Hall effect (AHE).\cite{Toyosaki_2004} Both GMR and TMR are widely observed in thin films\cite{Baibich_1988,Binasch_1989,Miyazaki_1995,Moodera_1995} where the magnetic coupling between layers can be artificially tuned. Observation in bulk materials\cite{Kusters_1989, Tokura_1999} revealed that CMR can be a bulk material property. Many mechanisms were suggested for the large magnetoresistance (MR) observed in bulk materials: nanoscale phase separation of the metallic ferromagnetic and insulating antiferromagnetic clusters in manganites;\cite{Uehara_1999, Dagotto_2001} metamagnetic transitions in rare earth intermetallics;\cite{Sechovsky_1994, Janssen_1997} and metal-insulator transitions and double exchange interactions for transition metal oxides.\cite{Hwang_1995, Millis_1996, Hammer_2004, Nakamura_2009}
While structures that exhibit GMR and TMR are already widely used in electronic devices, there remains strong technological and fundamental interest in homogeneous materials that exhibit large magnetoresistive effects. Moreover, since ordinary MR effects in bulk metals are typically only a few percent, understanding any occurrences of enhanced MR effects in bulk is of fundamental interest. In the ongoing search for new magnetic materials, transition metal dichalcogenides (TMDs) may be ideal candidates, due to their layered crystal structure and ease of intercalation with magnetic elements.\cite{Wilson1969, Wilson1975, Parkin_1980_1, Parkin_1980_2} For nearly forty years, the family of layered compounds Fe$_x$TaS$_2$ has been the subject of sustained inquiry focused on a surprising variety of anisotropic ferromagnetic properties.\cite{Eibschutz_1975, Morosan_2007} Prior studies have demonstrated that tuning the Fe concentration allows control of these magnetic properties, and measurements of magnetization, MR, and the anomalous Hall effect have been effective probes of the resulting modifications in behavior.\cite{Eibschutz_1975, Eibschutz_1981, Dijkstra_1989, Narita_1994, Morosan_2007, Checkelsky_2008} Here, we report experimental characterization of such a compound, with $x~\approx$ 0.28, which exhibits MR in the ordered state exceeding 60\%, nearly two orders of magnitude larger than was previously measured. By comparing our complementary results from bulk and thin exfoliated samples, we conclude that the large observed change in resistance is intrinsic and does not result from size-dependent phenomena, such as domain wall scattering. We argue that spin disorder scattering in the presence of strong spin-orbit coupling is the mechanism behind this MR, and that this is a potential paradigm for creating homogeneous materials with large MR. These observations suggest that the TMDs are rich targets for further theoretical study and potential industrial applications.\cite{Xu_2014}
\section{Methods}
Single crystals of Fe$_{0.28}$TaS$_2$ were prepared using iodine vapor transport in a sealed quartz tube, as described elsewhere.\cite{Morosan_2007} The typical size of the resulting bulk Fe$_{0.28}$TaS$_2$ single crystals was 2$\times$2$\times$0.1 mm$^3$. Powder x-ray diffraction revealed the expected Fe$_{0.28}$TaS$_2$ phase, with the lattice parameters consistent with a composition $x$ between 0.20 and 0.34.\cite{Eibschutz_1981} Energy-dispersive spectroscopy (EDS) and inductively coupled plasma (ICP) on bulk samples as well were used to more precisely determine the Fe concentration to be $x~=$ 0.28$\pm 3 \%$. EDS data were collected using a scanning electron microscope (SEM) equipped with an energy-dispersive spectroscopy (EDS) detector. ICP data were collected using a Perkin Elmer Optima 8300 ICP-OES system. The iron concentration of the sample was derived by comparison with commercial iron pure single-element standards (Perkin Elmer). Selected area electron diffraction (SAED) was also performed at room temperature on a bulk single crystal, ground in ethanol and placed on a holey carbon TEM grid.
The exfoliated samples were prepared using the tape exfoliation method.\cite{Novoselov_2004} Bulk Fe$_{0.28}$TaS$_2$ single crystals were mechanically cleaved using blue Nitto SPV 224 tape, and the resulting exfoliated crystals were deposited onto an oxidized silicon wafer (300 nm or 2 $\mu$m oxide thickness). Metallic contacts were defined using standard electron beam lithography and development. Contact metals were then deposited by electron beam evaporation of a Ti, Cr, or Fe adhesion layer ($\sim$ 3 nm) and Au (50 nm); an extra 20 nm of Au was added by sputtering.
For the exfoliated samples, the thickness was determined using atomic force microscopy (AFM). The measured thickness, with average values between 80 and 180 nm, varied by up to 21$\%$ within each sample. Scanning electron microscopy (SEM) images showed that the exfoliated flakes had lateral dimensions on the order of 10 $\mu$m, with variation from sample to sample. Thinner samples could only be produced with lateral dimensions much smaller than 10 $\mu$m due to relatively strong bonding between the layers compared to, \textit{e.g.}, graphite. Two exfoliated samples were prepared with electrodes configured to enable Hall measurements as well as conventional MR, while a third exfoliated sample was prepared for MR alone. Voltage probes were separated by less than 5~$\mu$m in these devices.
\begin{figure}[b!]
\includegraphics[width=1.0\columnwidth,clip]{arxivfig1.eps}
\caption{\label{magnetization_and_resistivity} (a) ZFC (solid symbols) and FC (open symbols) temperature-dependent magnetic susceptibility of a bulk sample measure in an applied field H = 0.1 T, H $\parallel$ c. Inset: The Curie temperature $T_C$ is determined from the minimum in $dM/dT$ (solid symbols) and an inflexion point in $d\rho/dT$ (line).and. (b) Temperature-dependent resistivity of both bulk (open symbols) and exfoliated (solid line) samples.}
\end{figure}
Temperature- and field-dependent magnetization data for bulk Fe$_{x}$TaS$_2$ were collected in a Quantum Design (QD) Magnetic Property Measurement System (MPMS). Temperature- and magnetic field-dependent AC resistivity measurements for both bulk and exfoliated Fe$_{0.28}$TaS$_2$ were performed in a QD Physical Property Measurement System (PPMS) using standard four-probe methods. Additional Hall resistivity data were collected using a five probe configuration for both the bulk and the exfoliated samples. Angle-dependent transport measurements were performed on an exfoliated sample mounted on a QD horizontal rotator insert, which allowed the sample to be rotated relative to the magnetic field direction.
\section{Results and discussion}
\begin{figure}[b!]
\includegraphics[width=1.0\columnwidth,clip]{arxivfig2.eps}
\caption{\label{magnetization_at_temperatures} H $\parallel~c$ (full symbols) field-dependent magnetization $M(H)$ data at various temperatures, together with the $T$ = 1.8~K, $H \parallel ab$ (open symbols) isotherm. For clarity, the two close isotherms ($H \parallel c$ for $T$ = 200~K and $H \parallel ab$ for $T$ = 1.8~K) are only shown for $H <$ 0 and $H >$ 0, respectively. }
\end{figure}
\begin{figure*}
\includegraphics[width=1.0\columnwidth,clip]{arxivfig3.eps}
\caption{\label{GMR} MR of (a) bulk and (b) exfoliated samples at selected temperatures for $H \parallel c$, and the current $i$ $\parallel$ $ab$. }
\end{figure*}
Fe$_x$TaS$_2$ is a unique intercalated transition metal dichalcogenide (TMD), with its strong and non-monotonic dependence of the magnetic properties (the ground state - ferromagnetic or antiferromagnetic, the ordering temperature) on the Fe concentration $x$.\cite{Eibschutz_1981, Dijkstra_1989,Narita_1994} It has been shown that a 3$\%$ difference in the Fe concentration (from $0.25$ to $0.28$) causes a modification of $T_C$ as large as 90 K (from 160 K to 70 K),\cite{Morosan_2007, Dijkstra_1989} {while increasing $x$ from $x~<~0.40$ to $x~\geq~0.40$\cite{Dijkstra_1989,Narita_1994} results in a change of the magnetic interactions from ferro- (FM) to antiferromagnetic (AFM). In the current Fe$_{0.28}$TaS$_2$ single crystals, the $H \parallel c$ temperature-dependent magnetic susceptibility measurements (Fig.~\ref{magnetization_and_resistivity}a) are consistent with the onset of FM order below $\sim$ 70 K upon cooling. The H = 0 temperature dependent resitivity data $\rho(T)$ on bulk (open symbols) and exfoliated (solid line) samples are virtually identical, as can be seen in Fig.~\ref{magnetization_and_resistivity}b. The weakly linear decrease in $\rho(T)$ at high T is indicative of the poor metal behavior in both bulk and exfoliated samples, while a drop below 70 K is consistent with loss of spin disorder scattering in the FM state. The derivatives of the ZFC magnetization data $dM/dT$ (symbols, inset) and the bulk resistivity data $d\rho/dT$ (line, inset) suggest that the Curie temperature T$_C$ is close to 68.8 K, if $T_C$ is determined from the minimum in $dM/dT$ and the inflection point in $d\rho/dT$ (vertical dashed line). The $T_C$ value is consistent with the reported $T_C$ for Fe$_{0.28}$TaS$_2$.\cite{Dijkstra_1989} We do find the onset of irreversibility in the zero-field-cooled (ZFC, solid symbols) and field-cooled (FC, open symbols) $M(T)$ data occurs around 150 K, well above T$_C$ for $x$ = 0.28 and very close to that for $x$ = 0.25.\cite{Morosan_2007} {This may be due to a small amount of Fe ions forming a commensurate superstructure as in Fe$_{0.25}$TaS$_2$, which, however, has very little effect on the transport properties where the transition is not even visible.}}
Remarkable behavior is observed in field-dependent magnetization and resistivity measurements with the magnetic field H along the reported easy axis H$\parallel$ $c$.\cite{Morosan_2007} The magnetization isotherms $M(H)$ of the bulk single crystals (Fig.~\ref{magnetization_at_temperatures}) reveal a sharp switching, similar to that for both Fe$_{0.28}$TaS$_{2}$\cite{Eibschutz_1981} and Fe$_{0.25}$TaS$_{2}$ compounds.\cite{Morosan_2007} The switching field $H_S$ is defined as the field where the magnetization crosses zero and where, as will be shown, the MR $\Delta \rho/ \rho_0$ and Hall resistivivity $\rho_{xy}$ display rapid changes as a function of H $\parallel$ c. In this study, both $H_S$ and the sharpness of the transition decrease with increasing temperature. $H_S$ at 1.8 K has the highest value of 6.23 T, while at T = 4 K, H$_S$ = 5.5 T, very close to value reported for Fe$_{0.28}$TaS$_{2}$.\cite{Eibschutz_1975} A second step-like feature in $M(H)$ appears for 7.5 $\leq~T~\leq$ 80 K and disappears when $T~>$ 200 K. While this could simply be attributed to the small amount of Fe$_x$TaS$_2$ phase with 0.25 $\leq$ x $\leq$ 0.28, this scenario is inconsistent with the absence of the additional M(H) step at the lowest temperatures. Another possible explanation for the second step-like feature could be heat release during the dynamic switching process in the bulk crystals, which could alter the shape of M(H). We do note that the magnetic and transport measurements are reproducible after the samples remain at low temperatures for long periods of time, and after performing multiple field sweeps at different sweep rates. Moreover, the $H \parallel c$ resistivity data $\rho$(H) in Fig.\ref{GMR} and anomalous Hall resistivity in Fig.\ref{AHE} feature a sharp jump at $H_S$.
\begin{figure*}
\includegraphics[width=1.0\columnwidth,clip]{arxivfig4.eps}
\caption{\label{AHE} Anomalous Hall resistivity for (a) bulk and (b) exfoliated samples at selected temperatures for $H$ $\parallel$ $c$, and the current $i$ $\parallel$ $ab$.}
\end{figure*}
MR is a crucial measurement for inferring information about the interactions between itinerant charge carriers and the magnetic degrees of freedom in a variety of magnetic materials.\cite{McGuire_1975, PhysRevB.57.R2037} The MR is defined as:
\[
\frac{\Delta \rho}{\rho_0} = \frac{\rho_{xx}(H)-\rho_{xx}(0)}{\rho_{xx}(0)}
\]
where $\rho_{xx}(H)$ is the value of the resistivity in a magnetic field $H$. The $\Delta \rho/ \rho_0$ measurements, with magnetic field H applied along the $c$ axis, were performed at selected temperatures for both bulk and exfoliated Fe$_{0.28}$TaS$_2$ single crystals (Fig.~\ref{GMR}a and b respectively). Below $T_C~\approx~68.8$ K, as the magnetic field H increases from 0 to 9 T, $\Delta \rho/ \rho_0$ smoothly increases to its maximum value at $H_S$ and sharply drops in a very narrow $H$ interval $\Delta H$, followed by a nearly linear decrease up to the maximum measured field $H$ = 9 T. When the magnetic field direction was reversed, the same change in $\Delta \rho/ \rho_0$ was observed, resulting in a bow-tie shape of $\Delta \rho/ \rho_0$ after one full cycle of field sweeping.
Qualitatively, this MR field-dependence resembles that for Fe$_{0.25}$TaS$_2$.\cite{Morosan_2007} However, the absolute $\Delta \rho/ \rho_0$ values are remarkably high in Fe$_{0.28}$TaS$_2$ (full symbols, Fig. \ref{GMR}a), nearly two orders of magnitude larger than that observed for x = 0.25 (open symbols, Fig. \ref{GMR}a). In both bulk and exfoliated Fe$_{0.28}$TaS$_2$ crystals, the largest $\Delta \rho/ \rho_0$ close to 60$\%$ was observed at T = 4 K (blue, Fig. \ref{GMR}). Furthermore, both $\Delta \rho/ \rho_0$ and $H_S$ decreased with increasing temperature, and the bow-tie shape of the $\Delta \rho/ \rho_0$ curves disappears above $T_C$ when $\Delta \rho /\rho_{0}$ becomes nearly linear for the whole measured field range. It should be noted that $\Delta H$ is much smaller in bulk ($\sim$ 0.04 T) than in the exfoliated sample ($\sim$ 0.8 T) at lower temperatures, and becomes comparable ($\sim$ 0.3 T) in both as the temperature exceeds 10 K. The broadening of the transition with increasing T in the bulk seems natural, while the opposite effect (sharpening) in the exfoliated sample emphasizes the role of the long range interplanar coupling in Fe$_{0.28}$TaS$_2$. This may imply that long range coupling exists between the Fe ions in different layers, which is weakened in the exfoliated sample, even when 100 nm thick.
The observed magnitude of the MR in Fe$_{0.28}$TaS$_{2}$, comparable to that seen in GMR and TMR systems, is remarkably large for a homogeneous bulk material not going through a phase transition (as in CMR systems). A useful point of comparison is (Ga,Mn)As, which has a similar $\rho$ vs $T$ response.\cite{PhysRevB.57.R2037} This latter material exhibits ordinary AMR, a spin-orbit coupling effect,\cite{McGuire_1975} which is typically at most a few percent in bulk materials based on 3$d$ transition metals. To gain insight into the very large MR in Fe$_{0.28}$TaS$_2$ it is necessary to correlate with other field-dependent measurements, like anomalous Hall effect (AHE). As previously observed in Fe$_{0.25}$TaS$_2$ and Fe$_{0.28}$TaS$_2$,\cite{Checkelsky_2008, Dijkstra_1989} the Hall resistance $\rho_{xy}$ for both bulk and exfoliated samples displays hysteresis below $T_{C}$, with jumps at $\pm~H_{S}$ (Fig.~\ref{AHE}). As was the case for $\Delta \rho /\rho_{0}$ (Fig. ~\ref{GMR}), $\rho_{xy}$ has a sharper jump at $\pm~H_{S}$ in the bulk sample than in the exfoliated one below 4 K, but then became comparable at higher temperatures. When $H$ exceeds $\pm~H_{S}$, the Hall resistivity $\rho_{xy}$ becomes almost linear in field, a result of the ordinary Hall effect contribution. For temperatures above $T_{C}$, only the ordinary Hall effect is observed, as $\rho_{xy}(H)$ is again nearly linear in $H$. Note that the Hall coefficient $R_H$ in Fe$_{0.28}$TaS$_2$ does not change sign throughout the ordered state, in contrast to the situation in Fe$_{0.25}$TaS$_2$.\cite{Checkelsky_2008} Converting into the Hall conductivity, the change in $\sigma_{xy}$ when passing through $H_{S}$ at 4~K is close to 200~S/cm, essentially the same as that seen in the $x=0.25$ compound,\cite{Checkelsky_2008} and exceeding the values typically seen in (Ga,Mn)As by a factor of five\cite{Nagaosa_2010}. These results imply that the spin-orbit coupling is very strong in this material and is very similar in the $x=0.28$ and $x=0.25$ compositions.
We must consider candidate mechanisms to explain the magnetotransport properties of the Fe$_{0.28}$TaS$_2$ single crystals, in particular the remarkably large H $\parallel$ $c$ MR. One natural possibility is AMR,\cite{McGuire_1975} parametrized in terms of the resistivities measured with the current density $\mathbf{J}$ parallel or perpendicular to the magnetization $\mathbf{M}$, $\rho_{||}$ and $\rho_{\perp}$, respectively. Generally the difference between the two $\rho_{\Delta} \equiv \rho_{||}-\rho_{\perp}$ is positive. The prior work\cite{Checkelsky_2008} on the $x=0.25$ compound ascribed the small (a maximum $\Delta \rho/\rho_{0} \approx 1.5\%$ at 1.5~K) MR for H $\parallel$ c to a $\rho_{\Delta}$ of +260$~\mu\Omega$-cm and a splaying of the spins as $H \rightarrow H_{S}$ by about 0.1$^{\circ}$. The large value of $\rho_{\Delta}$ is consistent in that case with in-plane MR measurements out to very high fields, showing $\Delta \rho/\rho_{0} \approx 40\%$ for H $\perp$ c and $H = 31$~T, corresponding to a tilting of $M$ away from the $c$ axis by around 15$^{\circ}$.\cite{Checkelsky_2008} Note that in these $x=0.25$ in-plane measurements at 10~K, an in-plane field of several Tesla is able to cant $M$ suffficiently to produce a measured AMR of several percent.
\begin{figure*}
\includegraphics[width=1.0\columnwidth,clip]{arxivfig7.eps}
\caption{\label{AngleDep} Angle-dependent measurements on an exofoliated sample of the longitudinal MR (left) and Hall resistivity (right) as a function of magnetic field H, for H $\parallel$ $c$, and the current $i$ $\parallel$ $ab$. (a) Data at T = 30 K for various field orientations relative to the $c$ axis. (b) Comparison of $H\parallel$ $c$ and $H\parallel$ $ab$ data for T = 10 and 30 K.}
\end{figure*}
In our $x~=~0.28$ compound, it is not unreasonable to assume a similar magnitude of $\rho_{\Delta}$, given the similarity of the spin-orbit coupling (inferred from the anomalous Hall conductivities) and the switching fields. Our observed magnitude of $\Delta \rho/\rho_{0}$ for $H\parallel~c$ would then imply a canting or splaying of the spins by tens of degrees immediately prior to magnetization reversal ($|H|~\lesssim~|H_S|$). Indeed, a significant rounding of $M(H)$ (Fig. \ref{magnetization_at_temperatures}) and $\sigma_{xy}(H)$ (Fig. \ref{GMR}) near $H_S$ for $H \parallel~c$ below, \textit{e.g.}, 10 K would at first glance seem to be compatible with this idea. However, angular dependent MR measurements on Fe$_{0.28}$TaS$_2$ strongly disfavor this possibility. Fig. \ref{AngleDep} displays MR $\Delta \rho/\rho_{0}$ (left) and $\rho_{xy}(H)$ (right) data for (a) different field orientations relative to the $c$-axis and constant temperature $T$ = 30 K, and (b) two extreme field orientations: $H \parallel c$ and $H \parallel ab$ for $T$ = 10~K (navy) and 30~K (orange). Within the AMR scenario of canting or splaying of the spins, one would expect significant canting of the magnetization when $H \parallel ab$ if such reorientation of $\mathbf{M}$ could happen with $H$ antialigned to $\mathbf{M}$ along $c$. Instead, there is almost no detectable magnetoresistive or anomalous Hall response for $H \parallel$ $ab$, and the magnetization response along that field direction (open symbols, Fig. \ref{magnetization_at_temperatures}) is correspondingly weak. This is in contrast to the $x=0.25$ case described above. These observations suggest that the easy axis of magnetization is strongly aligned with the $c$-axis, given that an in-plane field of 8 T is insufficient to produce any detectable MR or Hall signal. Thus ordinary AMR seems incompatible with the full ensemble of data, and AMR in the $x=0.28$ case appears to be quite different than at $x=0.25$.
\begin{figure}
\includegraphics[width=1.0\columnwidth,clip]{arxivfig8.eps}
\caption{\label{SAED}
SAED pattern of Fe$_{0.28}$TaS$_2$ crystal showing two concentric hexagonal sets of spots: the main structure (bright, large circles) and superlattice reflections (faint, small circles). The superstructure unit cell (small hexagonal cell) appears rotated by 90$\degree$ from the main structure unit cell (large hexagonal cell).}
\end{figure}
Giant magnetoresistance (GMR)\cite{Baibich_1988,Binasch_1989} is another mechanism capable of producing magnetoresistive effects of tens of percent. GMR results from the interplay between spin-split band structure and the density of states available for each spin species for scattering at the Fermi level. A magnetically inhomogeneous material can exhibit GMR due to current flow between differently aligned magnetic domains.\cite{Xiao_1992} To be a plausible explanation of our data would require that the magnetic domain structure of the material evolve as $H \rightarrow H_{S}$ so that charge transport is forced to take place across an increasingly large number of boundaries between antialigned domains. This can be tested through magneto-optic studies of the domain structure (beyond the scope of the present work). However, the micro-scale exfoliated samples have transport properties that look very similar to those of the bulk crystals, including a lack of any step-like features in the MR or anomalous Hall data as a function of field. This suggests that the flipping of discrete domains near $H_{S}$ and resultant GMR are unlikely to be responsible for the observed large MR. Note further that the domains observed via magneto-optic methods in the $x=0.25$ composition\cite{Vannette_2009} are typically tens of $\mu$m in extent. In the current x = 0.28 exfoliated samples the few-$\mu$m spacing of the voltage probes combined with the lack of any discrete magnetoresistive or AHE signatures in these devices implies that any domains would have to be much smaller than the $\mu$m scale - very different than the $x=0.25$ case, and difficult to image. Conversely, the similarity in the $M(H)$ data between this study and previous measurements on Fe$_{0.25}$TaS$_2$ suggests that the domain structures are likely very similar. Therefore, domain wall scattering is unlikely the cause of the large observed MR.
We suggest that the mechanism for the extremely large $H \parallel c$ MR and the near-absence of MR when $H \parallel ab$ is spin disorder scattering.\cite{Haas_1968,Otto_1989} The prominant drop in $\rho(T)$ when $T$ falls below $T_{C}$ is readily apparent in Fig.~\ref{magnetization_and_resistivity}, showing that spin disorder scattering accounts for approximately 50\% of the total scattering relevant to the resistivity above $T_{C}$. In the case of large spin-orbit coupling (SOC, as indicated by the size of the anomalous Hall conductivity in this material), it is not surprising that spin disorder can be so important. Rather than carrier-magnon scattering or Kondo physics, with the strong anisotropy and SOC the proposed mechanism for the large MR in the current x = 0.28 system is scattering from a (quasistatic) disordered arrangement of antialigned moments. In the presence of strong SOC, such spin disorder can be very effective at scattering carriers relative to ordinary potential disorder, since it mixes spin channels and therefore permits greater phase space for scattering.
When electron diffraction measurements are performed on the Fe$_{0.28}$TaS$_2$ single crystals (Fig.~\ref{SAED}), two concentric sets of spots are observed in the $ab$ plane, each with sixfold symmetry. The bright spots (large circles) are due to the main TaS$_2$ phase, while the faint spots (small circles) are assumed to result from an ordered Fe superstructure. When compared to the diffraction patterns presented in a recent study by Horibe $et~al.$,\cite{Horibe2014} the present SAED pattern appears more similar to that of Fe$_{1/3}$TaS$_2$ than that of Fe$_{1/4}$TaS$_2$, with the interior hexagon rotated by 90\degree in relation to the outer one and the resulting superstructure close to $\sqrt{3}\times\sqrt{3}$. The appearance of the superstructure spots in the electron diffraction (Fig.~\ref{SAED}) indicates that it may be useful to think about the $x~=~0.28$ system as a compound with a commensurate $x~=~0.25$ Fe structure with additional Fe local moments ($x~=~0.25~+~\delta$), or $x~=~0.33$ Fe structure with missing Fe local moments ($x~=~0.33~-~\delta$) with very small $\delta$ ($\delta~\leq~0.05$). In either case, the moments in a disordered environment, while coupled ferromagnetically to the bulk, would be expected to have weaker exchange interactions\cite{Ko_2011} than those on the superstructure sites, and hence easier to antialign with the field as $(H\parallel c) \rightarrow H_{S}$. The maximum MR for this field orientation is seen at $H_{S}$ as the spins reverse their orientation, leading to an increase in scattering comparable to the spin-disorder contribution to $\rho$. In other words, during the MR measurement, the antialignment of a significant fraction of the local moments as the field strength is increased (antiparallel to the bulk magnetization) results in enhanced scattering and increased resistance. Once the remaining spins flip to become aligned with the external field, spin disorder scattering is greatly reduced, causing a sharp drop in resistance. Canting of the moments is disfavored by the large magnetic anisotropy, while enhanced scattering (relative to potential scattering) is favored due to strong SOC and channel mixing.
Additional experiments can be used to test this hypothesis. This explanation assumes a population of weakly-coupled, easier-to-reorient spins due to deviations from the $x~=~0.25$ stoichiometry. One would therefore expect a monotonic increase in the the $H\parallel$ $c$ MR as $x$ is increased from $x~=~0.25$ to $x~=~0.28$. The dynamics of the spin reorientation should also be manifested in the MR response in this case, though no field sweep rate dependence has been observed so far. Optical perturbation of the local moment orientation would also be expected to lead to large resistive effects.
\begin{figure}[h]
\includegraphics[width=1.0\columnwidth,clip]{arxivfig5.eps}
\caption{\label{determination_Hs} Determination of the switching field $H_{S}$ for (a) bulk and (b) exfoliated samples from $M(H)$ (blue), MR (black) and anomalous Hall resistivity (red). The vertical dashed line marks the switching field $H_S$, as determined from the field values where M(H) and $\rho_{xy}$ cross H = 0, and where the fastest drop in $\Delta\rho/\rho_0$ occured.}
\end{figure}
In conclusion, we show that Fe$_{0.28}$TaS$_2$ single crystals display remarkably large MR, up to 60$\%$, when the applied magnetic field $H\parallel c$. Both the magnetization and transport properties appear nearly insensitive to sample thickness down to $\sim$ 100 nm, as measurements on bulk and exfoliated single crystals are nearly indistinguishable. As is illustrated in Fig. \ref{determination_Hs} for $T = 10$~K, the switching field $H_S$ values observed from magnetization and magneto-transport measurements on both bulk and exfoliated samples are very close at all temperatures up to $T_{C}$. The resulting temperature dependence of $H_S$ (squares) and $\Delta \rho/\rho_0$ at $H_S$ (circles) shown in Fig. \ref{Comparison} is indeed identical for both the bulk (full symbols) and exfoliated (open symbols) samples. Moreover, the non-monotonic change with x of the ordering temperature $T_C$ and switching field values $H_S$ between the Fe$_{0.28}$TaS$_2$ system and the previously reported Fe$_{0.25}$TaS$_2$ superstructure\cite{Morosan_2007}, and, more significantly, the nearly two order of magnitude enhancement of MR in the former compound, appear to be consistent with a scenario of disordered Fe moments mixed with a Fe superstructure. This scenario is even more plausible, given the experimental evidence we present to rule out other likely possibilities, such as AMR or an analog of GMR due to domain structure. The spin disorder scattering scenario reveals a design principle for intrinsically magnetoresistive materials without the need for multilayers or metal-insulator transitions coupled to magnetism. Conditions favoring maximal MR would include: single crystal materials, so that grain boundary, potential disorder, and surface scattering do not limit the mean free path; ferromagnetism with very strong uniaxial anisotropy, to favor moment flipping rather than canting as H is increased; and very strong spin-orbit coupling, magnifying the scattering cross-section of ``misaligned" spins. Transition metal dichalcogenides intercalated with various amounts of magnetic metals are promising materials where these optimal intercalation conditions may be achieved to maximize the observed MR, and such studies are currently underway.
\begin{figure}[h]
\includegraphics[width=1.0\columnwidth,clip]{arxivfig6.eps}
\caption{\label{Comparison} Comparison of $H_S$ and magnetoresistivity peak height values as a function of temperature for bulk (solid symbols) and exfoliated (open symbols) samples. $H_S$ increased monotonically with decreasing temperature, while the magnetoresistivity peak height increased with decreasing temperature until 4 K, and then decreased at lower temperatures. Inset: Image of a typical bulk sample (left), and false-color SEM image of a typical exfoliated sample with metal contacts (right).}
\end{figure}
\FloatBarrier
\section{Acknowledgements}
W.J.H. H. J., and D.N. acknowledge support from DOE BES award DE-FG02-06ER46337. E.M., C.W.C. and A.M. acknowledge support by DOD PECASE. J.S. acknowledges support by the Alexander von Humboldt Foundation. We thank Dr. Wenhua Guo for performing electron diffraction.
|
\section{Introduction}\label{sect:intro}
Teichm\"uller harmonic map flow, introduced in the joint work \cite{RT} with Peter Topping for closed surfaces, is
a geometric flow that is designed to change parametrised surfaces into critical points of the area.
Indeed, for closed surfaces in a non-positively curved target manifold,
the flow \textit{always} succeeds in changing, or more generally decomposing, the initial surface into (a union of) branched minimal
immersions through globally defined smooth solutions \cite{RT2}.
Here we take a first step to generalize this approach to the problem of flowing surfaces with boundaries to a solution of the Douglas-Plateau problem of finding a minimal surface spanning given boundary curves.
Namely, given two disjoint, closed $C^3$ Jordan curves $\Gamma_\pm$ in Euclidean space
we investigate how to flow a surface of cylindrical type in order to find a minimal surface spanning the two boundary curves.
As in \cite{RT} this flow will be constructed as a gradient flow of the Dirichlet energy
$$E(u,g)=\frac12\int_{C_0}\abs{du}_g^2 dv_g$$
considered as a function of two variables: a map $u:C_0\to \ensuremath{{\mathbb R}}^n$ parametrising the evolving
surface over a fixed domain, here the cylinder $C_0=[-1,1]\times S^1$, and a Riemannian metric $g$ on the domain.
We remark that if a pair $(u,g)$ is a critical point of $E$ then it is also a critical point of the area, to be more precise either a constant map or a (possibly branched) minimal immersion
\cite{GOR}.
A key idea of Teichm\"uller harmonic map flow is to consider $E$ on the set of equivalence classes of maps and metrics modulo the symmetries of $E$, compare \cite{RT}. Thus we identify
\begin{itemize}
\item $(u,g)\sim(u,\lambda\cdot g)$ for all functions $\lambda:C_0\to\ensuremath{{\mathbb R}}^+$ due to the conformal invariance of $E$, and
\item $(u,g)\sim (u\circ f,f^* g)$ for diffeomorphisms $f:C_0\to C_0$ that are homotopic to the identity.
\end{itemize}
As in \cite{RT} we shall then define Teichm\"uller harmonic map flow from cylinders as an $L^2$ gradient flow on the resulting set of equivalence classes
$$\ensuremath{{\mathcal A}}:=\{(u,g): u:C_0\to \ensuremath{{\mathbb R}}^n \text{ so that }
u\vert_{\{\pm1\}\times S^1} \text{ parametrises } \Gamma_\pm, \,g \text{ a metric on } C_0\}/\sim.$$
One important point to be understood in order to truly define such a flow
is how to define an $L^2$-metric on $\ensuremath{{\mathcal A}}$ and the closely related question of how to best represent a curve in $\ensuremath{{\mathcal A}}$ through pairs of maps and metrics.
In the definition of Teichm\"uller harmonic map flow on closed surfaces in \cite{RT}
we chose a canonical representative by asking that $g$ has constant curvature $K_g\equiv 1,0,-1$ (depending on the genus) and that $\norm{\partial_t g}_{L^2}$ is minimal.
For the resulting $L^2$-gradient flow this means that the symmetries are used to maximally simplify the evolution equation for the metric component in the sense that $g$ only moves by the part of the gradient of $E$ that is
orthogonal to the action of the symmetries. At the same time, the map component evolves with the full gradient, i.e. the tension.
As a result, cf. \cite{Rexistence},
for closed surfaces the evolution of the metric turns out to be well controlled as long as $\inj(M,g(t))\nrightarrow 0$,
while the map component shows a similar behaviour as the solutions of
the corresponding flow for fixed metrics,
i.e. the harmonic map heat flow of Eells-Sampson, which is well understood for closed domain surfaces and maps into general target manifolds.
In the present setting of flowing surfaces with boundary the situation is somewhat different, mainly because our boundary condition is not of Dirichlet-type,
but only of Plateau-type, i.e. prescribing the boundary values only
up to reparametrisation. As such, even for a fixed metric $g_0$, one cannot expect strong regularity results for the gradient flow of $u\mapsto E(g_0,u)$ unless one imposes a three-point-condition for $u\vert_{{\partial C_0}}$.
For maps $u$ parametrised over the disc $(D_1(0),g_{eucl})$ such a gradient flow of maps was introduced and studied by
Chang and Liu in \cite{CL1,CL2,CL3} who considered both maps into Euclidean space and into Riemannian manifolds.
The more general case of flowing to discs of prescribed mean curvature (and prescribed Plateau-boundary condition)
has been considered more recently by Duzaar and Scheven \cite{D-S}.
They show that an isoperimetric condition on the prescribed mean curvature ensures the existence of global weak solutions
and that these solutions subconverge to a
disc with the prescribed mean curvature.
In both cases, the flows are given as equations for only a map
component $u$. This is consistent with our approach as the special structure of the disc makes it unnecessary to also evolve a metric on the domain; namely,
the moduli space of the disc consists of only one point and one can furthermore pull-back any map by a suitable M\"obius transform to obtain a map that obeys a three-point-condition. Since M\"obiustransforms do not change the conformal structure this means that one can replace $(u,g_{eucl})$ by a representative of the same point of $\ensuremath{{\mathcal A}}$
whose map component satisfies a three-point-condition without having to adapt the metric component at all.
These special features of maps and metrics on the disc
are not present for any other surface with boundary, though in case of the cylinder the moduli space has a very simple structure as it is one dimensional. But even in this case, the group of conformal diffeomorphisms from $(C_0,g)$ to itself is not sufficiently large to
impose a three-point-condition for the map component without having to adjust the metric suitably. As
some kind of restriction on how $u\vert_{{\partial C_0}}$ parametrises the boundary curves $\Gamma_{\pm}$ is needed to obtain a flow that admits global solutions, we shall thus use the symmetries in a slighly different way than in the case of closed surfaces.
Namely, we use only most, but not all, symmetries to ensure that the evolution of the metric is regular, while also setting aside
a number of degrees of freedom ($3$ per boundary curve) to prevent a formation of singularities of the map
at the boundary by imposing a three-point-condition.
We remark that
the Douglas-Plateau problem has been considered by many authors and we refer to the books \cite{book-minimal}, \cite{Colding-Minicozzi}, \cite{Jost}, \cite{Struwe-book}
and the references therein for an overview of existing results. What we would like to point out is the well known fact
that while one can in general not prescribe the topological type of a minimal surface, one always obtains a minimal surface that is parametrised either over the original
domain, for us the cylinder, or over a surface of a simpler topological type, in the present situation two discs.
The paper is organised as follows.
To begin with, we discuss how to best represent curves in the set of equivalence classes
$\ensuremath{{\mathcal A}}$ and consequently give the precise definition of the flow. We then state our main results which
guarantee the existence of global weak solutions for arbitrary initial data,
see Theorem \ref{thm1}, as well subconvergence to either a minimal cylinder or to two minimal discs spanning the given
boundary curves, at least for solutions for which the three-point-condition does not
degenerate, see Theorem \ref{thm2}.
The rest of the paper is then dedicated to the proof of these results. In section \ref{sect:short-time}
we prove short-time existence based on a time-discretisation
scheme and derive a priori estimates on the map and metric component which are crucial for both the proof of existence and
of asymptotic convergence. This asymptotic analysis is carried out in section \ref{sect:asympt} but before that, in section
\ref{sect:long-time}, we establish that solutions exist for all times.
\section{Definition of the flow}
\subsection{Representing a curve in $\ensuremath{{\mathcal A}}$: Admissible variations}\label{sect:admissible}
As preparation for the definition of Teichm\"uller harmonic map flow on cylinders we discuss ways of representing curves in the set of equivalence classes $\ensuremath{{\mathcal A}}$ through suitably chosen pairs of
maps and metrics. We do not claim that our choice is canonical but rather that it is designed for the purpose of obtaining a gradient flow of energy that admits global regular solution.
To begin with, we need to identify a suitable representative of a conformal class $\texttt{c}$ of (smooth) metrics on $C_0$.
While one can always consider constant curvature, here flat, metrics with geodesic boundary curves, it turns out that
this particular representative is in general not the natural one to flow surfaces with boundary towards minimal surfaces. In particular, one would like to avoid the possibility
that a boundary curve of the domain (on which we after all impose our boundary condition) can shrink to a point and thus be lost.
For the cylinder we shall thus consider smooth metrics compatible with $\texttt{c}$ which have constant curvature $-1$ and for which
the boundary curves have both the same constant geodesic
curvature.
We first recall the following standard fact of complex analysis
\begin{lemma}\label{lemma:conf}
To any smooth conformal structure $\texttt{c}$ on $C_0$ there exists a unique number $Y>0$ such that $(C_0,\texttt{c})$ is conformally equivalent to
($[-Y,Y]\times S^1,ds^2+d\theta^2)$.
\end{lemma}
On such a cylinder $([-Y,Y]\times S^1,ds^2+d\theta^2)$ we can then use the following hyperbolic metrics whose structure is well known from the Collar lemma \cite{randol}
\begin{lemma}
\label{lemma:metric-cyl-1}
On $([-Y,Y]\times S^1,ds^2+d\theta^2)$ there is a one parameter family of \textit{collar metrics}
$$g_{\ell}=\rho_\ell(s)^2(ds^2+d\theta^2)$$
where
$$\rho_\ell(s)=\frac{\ell}{2\pi\cos(\frac{\ell s}{2\pi})}, \quad \ell\in(0,L_0(Y)), \quad L_0:=\frac{\pi^2}{Y},$$ which are all hyperbolic and whose boundary curves have the same constant geodesic curvature $\kappa\equiv \kappa_{\ell,Y}$.
\end{lemma}
As admissible metrics for our flow we shall thus consider
\begin{equation}\begin{aligned}
\ensuremath{{\mathcal M}}_{-1}:=\{f^* g: \, &f:C_0\to [-Y,Y]\times S^1 \text{ smooth diffeomorphism}, \\
&g=g_\ell \text{ a collar metric as in Lemma \ref{lemma:metric-cyl-1}}, \,\ell,Y\in (0,\infty)\}.\end{aligned}\end{equation}
While Lemma \ref{lemma:metric-cyl-1} does not yet give a canonical representative of a conformal class,
such a choice can be made so that the following splitting of the tangent space is respected
\begin{lemma}\label{lemma:splitting}
For any $g\in\ensuremath{{\mathcal M}}_{-1}$ we have
$$T_g\ensuremath{{\mathcal M}}_{-1}=\{L_X g: X\in \Gamma(TC_0)\}\oplus \text{Re}({\cal H}} %{\Upsilon(g))\oplus span\{\psi_g^2\cdot g\}.$$
where $\text{Re}({\cal H}} %{\Upsilon(g)\}$ is $L^2(C_0,g)$-orthogonal to $\{L_X g\}\oplus span\{\psi_g^2\cdot g\}$.
Here ${\cal H}} %{\Upsilon(g)$ is the real vector space of quadratic differentials that are holomorphic in the interior of
$C_0$, continuous upto the boundary and whose traces on ${\partial C_0}$ are real.
Furthermore $\Gamma(TC_0)$ stands for the space of smooth vectorfields on $C_0$ which are tangential
to ${\partial C_0}$ on ${\partial C_0}$ and $\psi_g:C_0\to \ensuremath{{\mathbb R}}$ is characterised by
$$\psi_g^2\cdot g=f^*\big(\tfrac{d}{d\ell}(\rho_\ell^2)(ds^2+d\theta^2)\big) \vert_{\ell=\ell_0}$$
for $g=f^*g_{\ell_0}$, $g_\ell$ the collar metrics of Lemma \ref{lemma:metric-cyl-1}.
\end{lemma}
We recall that
for the cylinder the space
${\cal H}} %{\Upsilon$
is simply made up by elements of the form $cdz^2$, $c\in\ensuremath{{\mathbb R}}$, $z=s+i\theta$, for collar coordinates $(s,\theta)\in [-Y,Y]\times S^1$ as in Lemma \ref{lemma:metric-cyl-1}.
We also remark that the orthogonality relation claimed in the lemma is a simple consequence of the fact that the real part of a holomorphic quadratic differential is trace and divergence free.
This lemma implies that the most \textit{efficient} way (i.e. with least $L^2$ velocity) to lift a curve $[g(\cdot)]$ from
Teichm\"uller space $\ensuremath{{\mathcal M}}_{-1}/\ensuremath{{\mathcal D}}_0$
to $\ensuremath{{\mathcal M}}_{-1}$ is as a \textit{horizontal} curve, moving only in the direction of $Re({\cal H}} %{\Upsilon(g))$. Here $\ensuremath{{\mathcal D}}_0$ denotes
the space of smooth diffeomorphisms from $C_0$ to itself that are homotopic to the identity.
For cylinders we can describe such horizontal curves of
metrics explicitly by the following lemma which is proved in the appendix
\begin{lemma}\label{lemma:horizontal-family}
Let $\eta>0$ be any fixed number. We define $Y=Y_\eta:(0,\infty)\to (0,\infty)$ by
$$Y(\ell)=\tfrac{2\pi}{\ell}\big(\tfrac\pi2-\text{atan}(\eta\cdot \ell) \big)$$
and $f_\ell=f_\ell^\eta:C_0\to [-Y(\ell),Y(\ell)]\times S^1$ by
$$f_\ell(x,\phi)=(s_\ell^\eta(x),\phi)=\big(
\tfrac{2\pi}{\ell} \text{atan}(\tfrac{\ell_0}\ell\cdot \tan(\tfrac{\ell_0}{2\pi}x))\, ,\,\phi\big)$$
where $\ell_0=\ell_0^\eta$ is determined through the condition
$Y(\ell_0)=1$.
Then the family $G_\ell:=(f_\ell)^*\big(\rho_\ell^2(s)(ds^2+d\theta^2)\big)$
is horizontal, i.e. for every $\ell$
$$\frac{d}{d\ell}G_\ell\in Re({\cal H}} %{\Upsilon(C_0,G_\ell)).$$
\end{lemma}
Since $Y(\cdot)$ is a bijection, we can combine the above result with Lemmas
\ref{lemma:metric-cyl-1} and \ref{lemma:conf} to conclude that any horizontal curve of metrics in $\ensuremath{{\mathcal M}}_{-1}$ must be of the form
$f^*(G_{\ell(t)}^\eta)$ for some fixed $\eta>0$ and a fixed diffeomorphism $f:C_0\to C_0$.
As such, we shall from now on consider $\eta>0$ to be fixed and will in particular allow all constants to depend on this number as well as on the boundary curves $\Gamma_\pm$
(and their parametrisations $\alpha_\pm$) without further mentioning this.
To describe the space of admissible maps, we first recall that
the prescribed boundary curves $\Gamma_{\pm}$ are assumed to be disjoint, closed $C^3$ Jordan curves of which we shall fix
proper $C^3$ parametrisations
$$\alpha_\pm:S^1\to \Gamma_\pm.$$
We then consider maps in the space
$$H_\Gamma^1(C_0,g):=\{u\in H^1((C_0,g),\ensuremath{{\mathbb R}}^n) \text{ such that } u:{\partial C_\pm} \to\Gamma_\pm \text{ is weakly monotone }\}$$
i.e. $H^1$ maps so that the traces $u\vert_{\partial C_\pm}$ can be written in the form
$$u\vert_{\partial C_\pm}=\alpha_\pm\circ \varphi_\pm$$
for some weakly monotone functions $\varphi_\pm:S^1\to S^1$.
Here and in the following we identify ${\partial C_\pm}:=\{\pm 1\}\times S^1$ with $S^1$ when convenient.
It is well known that the space $H_\Gamma^1$ is not closed under weak $H^1$ convergence
as one can find sequences of maps with bounded energy for which the boundary
curves $\Gamma_\pm$ are parametrised
over smaller and smaller arcs of ${\partial C_\pm}$ thus resulting in a weak limit that no longer spans $\Gamma_\pm$.
The standard way to deal with this loss of completeness is to impose a three-point-condition. So we shall restrict the set of admissible maps for our flow to
\begin{equation}\begin{aligned}
H_{\Gamma,*}^1(C_0)&:=\{u\in H_\Gamma^1(C_0): u\vert_{{\partial C_\pm}}=\alpha_\pm\circ\varphi_\pm \text{ for } \varphi_\pm
\text{ satisfying }\\
&\qquad \qquad \varphi_\pm(\theta_k)=\theta_k \text{ for } \theta_k=\frac{2\pi}{3} k, \,k=0,1,2\}.
\end{aligned}\end{equation}
To compensate for the (in our case $6$) lost degrees of freedom we need to allow the metric to move not only in
horizontal direction but also through the pull-back by select diffeomorphisms.
For this purpose we will define (and discuss) a suitable family of diffeomorphism $h_{b,\phi}$,
$\phi=(\phi^+,\phi^-)\in\ensuremath{{\mathbb R}}^2$, $b=(b^+,b^-)\in \ensuremath{{\mathbb C}}^2$, with $\abs{b^\pm}<1$
later on in section \ref{sect:diffeos}. We remark that by using the one Killing field that is available for the cylinder we could reduce the number of degrees
of freedom to $5$ instead of $6$
(e.g. by asking that $\phi^++\phi^-=0$) though this would not lead to a significant simplification.
All in all we then say that a curve $(u,g)(t)$ is an admissible representative of a curve of equivalence classes $[(u,g)(t)]\in\ensuremath{{\mathcal A}}$ if
\begin{itemize}
\item $g(t)=h_{\phi(t), b(t)}^*\tilde g(t) $ for a horizontal curve of metrics $\tilde g(\cdot)\in\ensuremath{{\mathcal M}}_{-1}$
and continuous families of parameters $(b,\phi)(\cdot)\in \Omega_h$ where
$\Omega_h:= (D_1(0))^2\times \ensuremath{{\mathbb R}}^2\subset \ensuremath{{\mathbb C}}^2\times \ensuremath{{\mathbb R}}^2$ is the domain of parameters for the diffeomorphisms $h_{b,\phi}$ that will be defined in
section \ref{sect:diffeos}.
\item $u(t)\in H_{\Gamma,*}^1(C_0,\ensuremath{{\mathbb R}}^n)$ for every $t$.
\end{itemize}
We remark that we can and will assume without loss of generality that the initial metric $\tilde g(0)$ of the horizontal curve $(\tilde g)$ is given by one of the
metrics $G_{\ell}$ described in Lemma \ref{lemma:horizontal-family}
simply by pulling-back the whole problem (including the parametrisations $\alpha_\pm$ used in the three-point-condition) by a fixed diffeomorphism. As such we shall from now on consider metric in the set
$$\widetilde{\mathcal{M}}:=\{ h_{b,\phi}^*G_\ell:\, (b,\phi)\in \Omega_h, \,\ell\in (0,\infty)\}.$$
\subsection{Definition of the flow}\label{sect:def}
As we consider a problem with a Plateau boundary condition, the space of admissible variations does not form a vectorspace.
As such the flow that we shall define will not be governed by a system of PDEs with prescribed boundary values but rather, for the map component, by a partial differential inequality.
To motivate the following definition we first make some general formal computations, which we of course do not claim to
be new in any way, but which are rather included for the convenience of the reader. We remark in particular that the differential inequalities we derive correspond to the ones obtained in \cite{CL1} and \cite{D-S} in case of the domain being a disc.
Given a functional $\mathcal{F}$ defined on some (Hilbert)manifold $B$
we may want to define a gradient flow under the restriction that the velocity $\partial_t w$ at each time is constrained to some closed
convex cone $X(w(t))\subset T_{w(t)}B$, e.g. because we want to constrain the flow to some convex set $A$ and thus the velocity to the corresponding solid tangent cone.
One can formally define such a gradient flow by asking that
\begin{equation}\label{eq:formal-grad-1}
\partial_t w=P^{X(w)}\bigg(-\nabla\mathcal{F} (w) \bigg)
\end{equation}
where $P^{X(p)}:T_pB\to X(p)$ is the nearest point projection.
We then observe that a variational formulation can be given by asking that at each time the velocity
$\partial_t w$ is given by a variation $\tfrac{d}{d\eps}\vert_{\eps=0} w_\varepsilon$ of $w_0=w(t)$ which is
admissible in that $\tfrac{d}{d\eps}\vert_{\eps=0} w_\varepsilon\in X(w(t))$
and which, among all such variations, minimises the functional
\begin{equation} \label{eq:formal-grad-2
\tfrac{d}{d\eps}\vert_{\eps=0}\mathcal{F}(w_\varepsilon)+\tfrac12\norm{\tfrac{d}{d\eps}\vert_{\eps=0} w_\varepsilon}^2 .\end{equation}
In practice, such a formulation often asks for more regularity, in particular of $\partial_t w$, than what we can a priori expect of a weak solution. So consider instead 2-parameter families $w_{\varepsilon,\delta}$
with $w_{\varepsilon,0}=w_\varepsilon$ such that each of the families $ w_{\cdot,\delta}$ gives again an admissible variation of $w(t)$. Then if
$w_{\varepsilon,0}=w_\varepsilon$ minimises \eqref{eq:formal-grad-2}
we must have that
\begin{equation} \label{eq:formal-grad-3}
\tfrac{d}{d\delta}\vert_{\delta=0}\tfrac{d}{d\eps}\vert_{\eps=0}\mathcal{F}(w_{\varepsilon,\delta})
+\langle\tfrac{d}{d\delta}\vert_{\delta=0}\tfrac{d}{d\eps}\vert_{\eps=0} w_{\varepsilon,\delta},
\partial_t w \rangle \geq 0
\end{equation}
which gives not only a weaker condition than \eqref{eq:formal-grad-2} but often requires less regularity of $\partial_t w$ than \eqref{eq:formal-grad-2}.
Going back to our problem of defining a gradient flow of the Dirichlet energy on the set $\ensuremath{{\mathcal A}}$ we recall that the (negative) $L^2$-gradient of
the energy with respect to the metric variable can be written in the form
$\frac14Re(\Phi(u,g))$, where $\Phi(u,g)$ is the Hopf-differential which is given in isothermal coordinates $(s,\theta)=(s_g,\theta_g)$ of $(C_0,g)$ by
$\Phi(u,g)=(\abs{u_s}^2-\abs{u_\theta}^2-2i\langle u_s,u_\theta\rangle)\cdot dz^2$ , $z=z_g=(s+i\theta)$.
The weak formulation \eqref{eq:formal-grad-3} thus translates to the condition that
with $v:=\tfrac{d^2}{d \varepsilon d\delta}\vert_{\delta=\varepsilon=0} u_{\varepsilon,\delta}$ and
$h=\tfrac{d^2}{d \varepsilon d\delta}\vert_{\delta=\varepsilon=0} g_{\varepsilon,\delta}$
\begin{equation}\begin{aligned}
\left[\int\langle dv,du\rangle_g +v \cdot \partial_tu \, dv_g\right]+\int \langle -\frac14\text{Re}(\Phi(u,g))+
\partial_t g, h\rangle \, dv_{g}\geq 0
\end{aligned}\end{equation}
for all variations $(u_{\varepsilon,\delta},g_{\varepsilon.\delta})$ of map and metric that are admissible in the sense described above.
On the one hand, the resulting differential inequality for $g$
can be simply recast as a differential equation
\begin{equation} \label{eq:evol-g}
\partial_t g=\frac14P^\mathcal{V}_g(Re(\Phi(u,g))\end{equation}
to be solved on $\widetilde{\mathcal{M}}$. Here $P^\mathcal{V}_g$ is the $L^2$-orthogonal projection
onto the tangent space $\mathcal{V}(g)$ of $\widetilde{\mathcal{M}}=\{ h_{b,\phi}^*G_\ell:\, (b,\phi)\in \Omega_h, \,\ell\in (0,\infty)\}$
which is given by
\begin{equation} \label{eq:VS-variations} \mathcal{V}(g):=Re({\cal H}} %{\Upsilon(g))\oplus \{L_{h_{b,\phi}^*X}g: X\in \mathcal{X}(b,\phi)\},\end{equation}
for $g=h_{b,\phi}^*G_\ell$
where
$\mathcal{X}(b,\phi)$ is the $6$ dimensional space of vectorfields generating the diffeomorphisms $h_{b,\phi}$, compare section \ref{sect:diffeos}.
On the other hand, admissible variations of the map can be described as follows.
Given $u\in H_{\Gamma,*}^1(C_0)$ we let $\varphi_\pm$ be such that $u\vert_{\partial C_\pm}=\alpha_\pm\circ \varphi_\pm$.
Then functions of the form $\alpha_\pm\circ (\varphi_\pm+\varepsilon\cdot \beta_\pm+O(\varepsilon^2))$ are again monotone parametrisations of $\Gamma_{\pm}$ at least for $\varepsilon$ in a small onesided interval
$[0,\varepsilon_0)$ if $\beta_\pm$ can be written in the form $\beta_\pm=\lambda_\pm\cdot (\psi_\pm-\varphi_\pm)$ for some numbers $\lambda_\pm>0$ and weakly monotone functions $\psi_\pm$.
As variations $\tfrac{d}{d\eps}\vert_{\eps=0} u$ of the map component we thus consider elements of
\begin{equation}\begin{aligned}
T_u^+H^1_{\Gamma,*}(C_0):=\{v\in H^1(C_0,\ensuremath{{\mathbb R}}^n): &\, v\vert_{\partial C_\pm}=\lambda_\pm \alpha'_\pm(\varphi)\cdot (\psi_\pm-\varphi_\pm) \text{ for } \lambda_\pm>0 \text{ and }\\ &
\psi_\pm\in C^0(S^1,S^1) \text{ weakly monotone with } \psi_\pm(\theta_k)=\theta_k\}.\end{aligned}\end{equation}
We remark in particular that to any $v\in T_u^+H^1_{\Gamma,*}(C_0)$ there is a onesided variation $(u_\varepsilon)\subset H_{\Gamma,*}^1(C_0)$, $\varepsilon\in [0,\varepsilon_0)$ with $\tfrac{d}{d\eps}\vert_{\eps=0} u_\varepsilon=v$, see \cite[Lemma 2.1]{D-S}.
As
$T_u^+ H^1_{\Gamma,*}$ is in general not a vectorspace, but only a convex cone, we cannot reduce the resulting partial differential inequality
\begin{equation} \label{eq:evol-u}
\langle du, dw\rangle_{L^2(C_0,g)}+\int w\cdot \partial_t u\, dv_g\geq 0 \text{ for all } w\in T^+_u H^1_{\Gamma,*}(C_0)
\end{equation}
to a PDE with a standard boundary condition though one immediately obtains that
$u$ satisfies the heat equation
$\partial_t u=\Delta_g u$
in the interior.
Furthermore, as pointed out in \cite{D-S}, the additional
condition that
$$\int\langle du, dw\rangle+\Delta_g u \cdot w \,dv_g\geq 0 \text{ for all } w\in T^+_u H^1_{\Gamma,*}, $$
can be seen as a weak Neumann-type boundary condition.
Given that \eqref{eq:evol-g} and \eqref{eq:evol-u} were motivated by the idea that $\partial_t(u,g)$ should minimise the functional
\eqref{eq:formal-grad-2}, the so called stationarity condition, asking that
\begin{equation}\label{eq:stationar}
\frac14\int Re( \Phi(u,g))L_X g dv_g+\int Du(X)\cdot \Delta_g u dv_g =0 \text{ for every } X\in \Gamma(TC_0)_*
\end{equation}
where
$$\Gamma(TC_0)_*:=\{Y\in \Gamma(TC_0): Y(\pm1, \theta_k)=0\},$$ results if one considers
variations of the form $(u(t+\varepsilon)\circ f_\varepsilon, g)$.
Similarly one expects the energy to be non-increasing along the flow, compare \eqref{eq:energy-cond} below.
All in all, we define
\begin{defn}
A weak solution of Teichm\"uller harmonic map flow on the cylinder $C_0$ is represented by a curve of maps
$$u\in L^\infty([0,T),H^1_{\Gamma,*}(C_0,\ensuremath{{\mathbb R}}^n))\cap H^1([0,T)\times C_0)$$
and a curve of metrics
$g\in C^{0,1}([0,T),\widetilde{\mathcal{M}})$
which satisfy
\begin{equation}
\label{eq:fluss-map}
\int_{[0,T]\times C_0}\langle du, dw\rangle_{g}+ \partial_t u\cdot w\, dv_{g(t)} \, dt \geq 0
\text{ for all }w\in L^2([0,T],T^+_u H^1_{\Gamma,*}(C_0))
\end{equation}
and
\begin{equation} \label{eq:fluss-metric}
\partial_t g=\tfrac14P^\mathcal{V}_g(Re(\Phi(u,g)) \qquad \text{for a.e. } t.
\end{equation}
Such a weak solution is called stationary if it satisfies \eqref{eq:stationar} for almost every $t$, and we say $(u,g)$ satisfies the energy inequality if
for almost every $t_1<t_2$
\begin{equation} \label{eq:energy-cond} E(u,g)(t_1)-E(u,g)(t_2)\geq {\tfrac12} \int_{t_1}^{t_2}\int_{C_0}\abs{\partial_t u}^2 dv_g dt+ \int_{t_1}^{t_2}\norm{\partial_t g}_{L^2(C_0,g)}^2 dt.\end{equation}
\end{defn}
\subsection{Main results}
For the flow we just defined we will prove the following two main results
\begin{thm}[Existence of global solutions]\label{thm1}
Let $\Gamma_{\pm}$ be two disjoint closed $C^3$ Jordan curves.
Then to any initial data $(u_0,g_0)\in H_{\Gamma,*}^1(C_0)\times \widetilde{\mathcal{M}}$
there exists a stationary weak solution $(u,g)$ of Teichm\"uller harmonic map flow which is
defined for all times, smooth in the interior of $C_0$ and satisfies the energy inequality.
\end{thm}
The above solution flows to a minimal surface in the sense that
\begin{thm}[Asymptotics]\label{thm2}
Let $(u,g)(t),\, t\in [0,\infty)$, be a stationary weak solution of Teichm\"uller harmonic map flow that satisfies the energy inequality
and for which the three-point-condition does not degenerate in the sense that
$\limsup_{t\to \infty} 1-\abs{b^\pm(t)}>0$.
Then there is a sequence of times $t_i\to\infty$ such that the equivalence classes $[(u,g)(t_i)]$ converge to a critical point of the area in one of the following ways:
\begin{itemize}
\item[I](Non-degenerate case)
If $inj(C_0,g(t_i))\nrightarrow 0$ for $i\to \infty$ then
$f_i^*(u(t_i),g(t_i))$ converges to a limit $(u_\infty,g_\infty)$ where $g_\infty\in \widetilde{\mathcal{M}}$ and
where $u_\infty\in H^1_{\Gamma,*} (C_0, \ensuremath{{\mathbb R}}^n)\cap C^0(C_0)$
is a (possibly branched) minimal immersion.
Here
\begin{equation} \label{eq:diffeos-asympt}
f_i:=h_{0,2\pi n_i}, \quad n_i^\pm=\lfloor\tfrac{\phi^\pm(t_i)}{2\pi}\rfloor
\end{equation}
and the convergence for the metric component is smooth convergence on all of $C_0$ while the maps converge
uniformly on the whole cylinder $C_0$ as well as strongly in $H^1(C_0)$ and weakly in $H^2_{loc}(C_0\setminus \bigcup_{j,\pm}P_j^\pm)$ away from the points
$P_j^\pm=(\pm1,\theta_j)$, $j=0,1,2$ at which the three-point-condition is imposed.
\item[II](Degenerate case)
If $\inj(C_0,g(t_i))\to 0$
then
$f_i^*(u(t_i),g(t_i))$ converges locally on $C_0\setminus \big(\{0\}\times S^1\big)$ to a limit
$ (u_\infty, g_\infty)$
which is such that
\begin{itemize}
\item[\textbullet]
each of the cylinders $(C_\pm,g_\infty)$, $C_{\pm}:=\{0<\pm s\leq 1\}\times S^1$
is isometric to the hyperbolic cusp
$$([0,\infty)\times S^1, \rho_0(s)^2(ds^2+d\theta^2)),\quad \rho_0(s)=\frac1{2\pi \eta+s}$$
\item[\textbullet]
The two maps $u_\infty\vert_{C_\pm}$ can be extended across the punctures to give
two (possibly branched) minimal immersion
$\bar u_\infty^\pm\in H^1_{\Gamma_\pm, *}(\overline{D})\cap C^0(\overline{D})$ parametrised over closed disc in $\ensuremath{{\mathbb R}}^2$ each of which spans the corresponding boundary curve $\Gamma^\pm$.
\end{itemize}
Here the convergence is smooth local convergence for the metrics and weak $H^2_{loc}$ convergence on $(C_-\cup C_+)\setminus \bigcup_{j,\pm} P_j^\pm$
as well as locally uniform and
strong $H^1_{loc}$ convergence on $C_-\cup C_+$ for the maps
and the diffeomorphisms are again given
by \eqref{eq:diffeos-asympt}.
\end{itemize}
\end{thm}
We remark that while we only obtain convergence in $H^1\cap C^0$ respectively in $H^2$ away from $P_j^\pm$, the limit $u_\infty$ is indeed far more regular than that.
Namely, classical regularity theory for solutions of the Plateau-Problem, see e.g. \cite{Struwe-book} or \cite{book-minimal},
yields that $u_\infty$ is of class $C^{2,\alpha}$, $\alpha<1$, upto the boundary.
\section{Short-time existence of solutions}\label{sect:short-time}
We shall prove short-time existence of solutions to arbitrary initial data
based on a time discretisation scheme. We remark that this method has been carried out successfully to obtain solutions of several other geometric flows, e.g. by Haga et. al. \cite{Haga} for harmonic map flow and by Moser \cite{Moser} for biharmonic map flow,
and that also the solutions for the evolution to minimal discs by Chang-Liu \cite{CL1} respectively to discs of prescribed mean curvature of Duzaar-Scheven \cite{D-S} were obtained this way.
A key part of this section consists in proving suitable a priori estimates for the
approximate solutions resulting from such a time discretisation.
For some of these estimates we will be able to appeal to work of Duzaar and Scheven \cite{D-S} whose delicate estimates allowed them
to prove
$H^2$ bounds upto the boundary but away from the points $P_j^\pm$ despite their equation being non-linear.
In the present paper the challenges are somewhat different as we do not have to deal with a non-linear equation for the map component
but instead have to understand the interplay of the map and the metric component of the flow. What makes this particular aspect of the flow quite delicate,
is that this relation involves a non-local projection operator.
This forces us to prove estimates that are valid not just near most boundary points but rather in neighbourhood of every boundary point,
including the points $P_j^\pm$ at which we impose the three-point-condition.
\subsection{The time discretisation scheme}\label{sect:time-dicr}
To begin with we outline the time-discretisation scheme and show that it is well defined.
Given an initial pair $(u_0,g_0)\in H_{\Gamma,*}^1(C_0)\times \widetilde{\mathcal{M}}$
and a (small) number $h>0$ we construct an approximate solution of Teichm\"uller harmonic map flow
using the following time-discretisation:
For $j=0,1,2,..$ we let $t_j=t_j^h=j\cdot h$, set $u^h(t)=u_0$ for $t\in [t_0,t_1]$ and then construct iteratively the approximate solution
$(u^h,g^h)$ on the interval $[ t_j^h, t_{j+1}^h]$ as follows:
First determine
$g^{h}(\cdot)$ on $( t_j^h, t_{j+1}^h]$ as the solution of
\begin{equation}
\label{eq:gh}
\partial_t g(t):=\frac14P_{g(t)}^\mathcal{V}( \Phi(u^h(t_j),g(t)) \text{ with } g(t_j)=g^h(t_j).
\end{equation}
Then select $u^h(t_{j+1})$ as a minimiser of the functional $\mathcal{F}_{g^h(t_{j+1}), u^h(t_{j})}^h$ where
\begin{equation} \label{eq:minim-fhj}
\mathcal {F}_{g,v}^h(w)=E(w,g)+\frac1{2h}\norm{w-v}_{L^2(C_0,g)}^2.\end{equation}
The existence of a minimiser of this functional is assured by the direct method of calculus of variation
thanks to the $H^1$-weak-lower semicontinuity of $u\mapsto E(u,g)$ as well as the Courant Lebesgue Lemma and the resulting equicontinuity of the traces $u\vert_{\partial C_0}$, c.f. appendix \ref{sect:Courant}.
To be more precise, we have
\begin{lemma}\label{lemma:ex-minimiser}
For any $g\in \ensuremath{{\mathcal M}}_{-1}$, any map $\bar u\in H^{1}_{\Gamma,*}(C_0)$ and any $h>0$ there exists a minimiser $w\inH_{\Gamma,*}^1(C_0)$ of
$$\mathcal{F}(w)=E(w,g)+\tfrac1{2h}\norm{w-\bar u}_{L^2(C_0,g)}^2$$
and $w$ satisfies
\begin{equation}\label{eq:E-L-eq-minimiser}
\int_{C_0} \langle dw,dv\rangle_g+\tfrac 1h\cdot(w-\bar u)\cdot v \,dv_g\geq 0 \text{ for all } v\in T^+_wH_{\Gamma,*}^1(C_0)
\end{equation}
in particular $\Delta_g w=\tfrac 1h\cdot(w-\bar u)$ in the interior of $C_0$.
Furthermore $w$ satisfies the stationarity equation
\begin{equation} \label{eq:stat.eq-min} \frac14\int Re(\Phi(w,g)) \cdot L_Xg dv_g+\int dw(X)\cdot \Delta_g w \,dv_g=0 \text{ for all } X\in \Gamma(TC_0)_*\end{equation}
and the energy inequality
\begin{equation} \label{est:energy-ineq-EL}
E(w,g)+\tfrac1{2h}\cdot \norm{w-\bar u}_{L^2(C_0,g)}^2\leq E(\bar u,g).\end{equation}
\end{lemma}
We remark furthermore that the minimiser $w$ is bounded by
\begin{equation} \label{est:Linfty}
\norm{w}_{L^\infty(C_0)}\leq \norm{\bar u}_{L^\infty(C_0)} \end{equation}
so that the $L^\infty$ norm of the map component of the flow is non-increasing in time.
Indeed, if the above estimate would not be satisfied, we could compose $w$ with the nearest point projection
to the ball with radius $\norm{u}_{L^\infty}$ to obtain a function with smaller energy $\mathcal{F}$.
In the present setting of metrics on a cylinder,
short-time existence of a solution to the differential equation \eqref{eq:gh} on $\widetilde{\mathcal{M}}$ is a simple
consequence of the fact that ${\cal H}} %{\Upsilon$ is one-dimensional since this means that
the evolution of
the metric could be expressed as a system of (in total $7$) ordinary differential equations.
As such it is easy to check that the projection satisfies the following Lipschitz- estimates
\begin{lemma}\label{lemma:Lip-Projection}
Let
$K$ be a compact subset of the set of admissible metrics
$\widetilde{\mathcal{M}}=\{g=h^*_{b,\phi}G_\ell, \ell\in (0,\infty), (b,\phi)\in \Omega_h\}$ and let $P^{\mathcal{V}}_g$ be the $L^2$-orthogonal projection onto $\mathcal{V}(g):=T_g\widetilde{\mathcal{M}}$.
Then
$$\norm{P^\mathcal{V}_{g_1}(Re(\Psi_1))-P^\mathcal{V}_{g_2}(Re(\Psi_2))}_{C^k}\leq C\norm{g_1-g_2}_{C^k}\cdot \norm{\Psi_1}_{L^1(C_0)}
+C\norm{Re(\Psi_1)-Re(\Psi_2)}_{L^1(C_0)}$$
for all $g_{i}\in K$, quadratic differentials $\Psi_i$ on $(C_0,g_i)$, $i=1,2$ and $k\in \ensuremath{{\mathbb N}}$, where $C$ depends only on $k$ and $K$.
Here and in the following
the $C^k$ norms are computed with respect to a fixed coordinate chart.
\end{lemma}
We remark that the real part of the Hopf-differential can be written equivalently as
\begin{equation} \label{eq:Hopf-coord}
Re(\Phi(u,g))=2u^*g_{\ensuremath{{\mathbb R}}^n}-\abs{du}_g^2\cdot g
\end{equation}
so that we can bound the differences of Hopf-differentials by
\begin{equation}
\norm{Re(\Phi(u,g))-Re(\Phi(\tilde u,\tilde g))}_{L^1}\leq C\norm{g-\tilde g}_{C^0}+C\norm{u-\tilde u}_{H^1}\end{equation}
for a constant $C$ that depends only on a bound for the energies $E(u,g)$, $E(\tilde u,\tilde g)$.
Combined, Lemmas \ref{lemma:ex-minimiser} and \ref{lemma:Lip-Projection} thus
imply that solutions of the time discretisation scheme exist for as long as
the injectivity radius $inj(C_0,g)=2\ell$ of the domain $(C_0,g)$ is bounded away from zero and infinity and the parameters $(b,\phi)$
remain in a compact set of $\Omega_h$.
We furthermore remark that the energy of such an approximate solution is non-increasing, namely on the open interval $(t_k,t_{k+1})$ it
decreases by
\begin{equation*}\begin{aligned}
E((u^h,g^h)(t_k))-E((u^h(t_k),g^h(t_{k+1}))&=\int_{[t_k,t_{k+1}]}
\norm{\partial_t g^h}_{L^2(C_0,g^h)}^2 dt\\
&=\frac1{16}\int_{[t_k,t_{k+1}]} \norm{P^\mathcal{V}(Re(\Phi(g^h(t), u^h(t_k))}_{L^2(C_0,g^h)}^2 dt
\end{aligned}\end{equation*}
while at $t_{k+1}$ there is a further loss of energy of no less than
$$E(u^h(t_k),g^h(t_{k+1}))-E((u^h,g^h)(t_{k+1}))\geq \frac1{2h}
\norm{u^h(t_k)-u^h(t_{k+1})}^2_{L^2(C_0,g^h(t_{k+1}^{h}))},$$
compare \eqref{est:energy-ineq-EL}.
All in all we can thus estimate
\begin{equation}\label{est:energy-ineq-approx-scheme}
E((u^h,g^h)(t_k))\leq
E((u^h,g^h)(t_{\tilde k}))- \int_{t_{\tilde k}}^{t_k}
{\tfrac12}\norm{D_t^{h} u^h}_{L^2(C_0,\tilde g^h)}^2
+ \norm{\partial_t g^h}_{L^2(C_0,g^h)}^2 dt\end{equation}
for $\tilde k<k$ where compute the norm of the difference quotient
$$D_t^{h} u^h(x,t)=\tfrac1{h} \cdot(u^h(x,t+h)-u^h(x,t))$$
with respect to the piecewise constant curve of metrics
$\tilde g^h\vert_{[t_k,t_{k+1})}=g^h(t_{k+1})$.
We obtain in particular that the length of the curve of metrics is bounded uniformly by
$$L_{L^2}(g^h\vert_{ [0,t]})\leq E_0^{1/2} t^{1/2} \text{ for every }h>0,$$
$E_0$ an upper bound on the initial energy $E(u(0),g(0))$.
Consequently, to any given $(u_0,g_0)$ there exist numbers $\varepsilon>0$, $T>0$ and $\bar C<\infty$ such that for every parameter $h>0$ the
solution of the time-discretisation scheme exists at least on the interval $[0,T]$
and so that the metric component $g^h=h_{b,\phi}^*G_{\ell}$ satisfies estimates of the form
\begin{equation} \label{ass:inj-apriori}
\inj(C_0,g^h)=2 \ell\geq \varepsilon \text{ and } \abs{b_j^\pm}\leq 1-\varepsilon,\quad \abs{\phi^\pm}\leq \bar C \end{equation}
on this interval. \cmt{add reference}
Similarly we could obtain an upper bound for $\ell$ on $[0,T]$, but we remark
that $\ell$
is indeed bounded from above uniformly in time
in terms of only the initial energy.
Namely, let $\delta_\Gamma:=\text{dist}(\Gamma_+,\Gamma_-)>0$ be the distance between our prescribed disjoint boundary curves. Then the
energy of any map $w\in H^1_\Gamma(C_0)$ with respect to a metric $g$ of the form $g=h_{b,\phi}^*G_\ell$
is bounded from below by
\begin{equation*}\begin{aligned}
E(w,h_{b,\phi}^*G_\ell)
& =E(\tilde w,G_\ell)=\frac12\int_0^{2\pi}\int_{-Y(\ell)}^{Y(\ell)}\abs{\tilde w_s}^2+\abs{\tilde w_\theta}^2 ds d\theta\\
&\geq \frac{ c}{Y(\ell)}\big(\int_0^{2\pi}\int_{-Y(\ell)}^{Y(\ell)}\abs{\tilde w_s} ds d\theta\big)^2
\geq \frac{\tilde c \delta_\Gamma^2}{Y(\ell)},
\end{aligned}\end{equation*}
for some $\tilde c>0$ and $\tilde w=w\circ h_{b,\phi}^{-1}$. The resulting lower bound on $Y(\ell)$ results in an upper bound for $\ell$ and thus the injectivity radius
of the form
\begin{equation}\label{est:apriori-upper-ell}
\ell\leq \bar L(E_0),\end{equation}
where $\bar L(E_0)$ depends only on an upper bound $E_0$ on the initial energy $E(u_0,g_0)$ (and as usual
the geometric setting and the fixed number $\eta$).
This uniform upper bound on $\ell$ and the resulting
control on the metrics $G_\ell$ near the boundary $\partial C_0$ will allow us to
derive a priori bounds for the map component near
$\partial C_0$ that are independent of $\inj(C_0,g)$. This will be crucial for the asymptotic analysis (in the degenerate case) carried out later on.
Conversely, all estimates near the boundary will depend on $b$ as the case $\abs{b_\pm}\to 1$ corresponds to the degeneration
(in collar coordinates) of the three-point-condition.
To prove that the above time-discretisation scheme converges to a solution of the flow, we need to derive a priori estimates for the map component where we will distinguish between
\begin{itemize}
\item the interior of the cylinder where standard estimates for the heat equation apply
\item the boundary region away from the points $P_k^\pm=(\pm1,\theta_k)$, on which we shall be able to appeal to results of Duzaar-Scheven \cite{D-S}
\item the region near the points $P_k^\pm$ at which the three-point-condition is imposed.
\end{itemize}
\subsection{A priori estimates in the interior and near general boundary points}
Let $\bar u$ be any fixed map, $h>0$, let $g=h_{b,\phi}^*G_\ell$ be a metric as in Lemma \ref{lemma:horizontal-family} and
let $u$ be a minimiser of the functional
$\mathcal{F}_{g,\bar u}^h=E(u,g)+\tfrac1{2h}\norm{u-\bar u}_{L^2(C_0,g)}^2.$
Then, setting $f=\frac1h( u-\bar u)$ we know
that
\begin{equation}
\label{eq:EL-minimiser-f}
\int_{C_0}\langle du,dv\rangle dv_g+\int_{C_0}v \cdot f dv_g\geq 0
\text{ for } v\in T^+_uH_{\Gamma,*}^1(C_0).\end{equation}
In particular
$\Delta_g u=f$ in the interior, so standard elliptic estimates combined with the upper bound \eqref{est:apriori-upper-ell}
on $\ell$ yield
\begin{lemma}
\label{lemma:interior-minimiser}
Given any $\ell_0>0$, $E_0<\infty$ and any $\delta>0$ there exists a constant $C<\infty$
such that the following holds true. Let
$u\in H_{\Gamma,*}^1(C_0)$ be any map of energy $E(u,g)\leq E_0$
which satisfies \eqref{eq:EL-minimiser-f} for some $f\in L^2(C_0)$
and $g\in \widetilde{\mathcal{M}}$ with $2\inj(C_0,g)>\ell_0$. Then
\begin{equation} \label{est:H2-int}
\int_{[-1+\delta,1-\delta]\times S^1}\abs{\nabla_g^2 u}^2 +\abs{d u}_g^4 dv_g\leq C\cdot \big(E(u,g)+\norm{f}_{L^2(C_0,g)}^2).\end{equation}
\end{lemma}
We remark that
the region in which $G_\ell$ degenerates as $\ell\to 0$ is contained in what corresponds to
arbitrarily small cylinders
$[-\delta,\delta] \times S^1$ with respect to the fixed coordinates
$(x,\theta)\in C_0=[-1,1]\times S^1$ since we use hyperbolic rather than flat metrics to represent a conformal class. Therefore
\begin{rmk}
The analogue of \eqref{est:H2-int} is valid with a constant independent of $\ell_0$
on every compact region of $(-1,1)\times S^1\setminus (\{0\}\times S^1)$
and for every metric $g\in \widetilde{\mathcal{M}}$ as well as for metrics $h_{b,\phi} G_{\ell=0}$ described in \eqref{eq:G0}.
\end{rmk}
Near the boundary but away from the points $P_j^{\pm}$ we can use the results of Duzaar-Scheven \cite{D-S}
which apply to more general (in particular non-linear) equations.
Namely, as explained in appendix \ref{sect:Courant}, we can derive the following a priori estimates from Theorem 8.3 of \cite{D-S}
\begin{prop}\label{prop:Duzaar-Scheven}
For any $b_0<1$, $\ell_0>0$ and $E_0<\infty$ there exist constants $\varepsilon_1,r_0>0$ and $C<\infty$ such that the following holds true.
Let $g=h_{b,\phi}^*G_\ell$ for some $\ell\geq \ell_0$ and $\abs{b^\pm}\leq b_0.$
Suppose furthermore that
$f\in L^2(C_0,g)$ and that $u\in H_{\Gamma,*}^1(C_0)$ has energy $E(u,g)\leq E_0$.
Then, if $u$ is so that \eqref{eq:EL-minimiser-f} is satisfied for all variations
$v\in T^+H_{\Gamma,*}^1(C_0)$ with support in a ball $B_r^g(p)$, where
$p\in C_0$ and $r\in(0,r_0)$ are such that
$$B_r^g(p)\cap \{P_j^\pm\}=\emptyset$$
and if the energy on this ball is small in the sense that
$$E(u,B_r^g(p)):={\tfrac12}\int_{B_r^g(p)}\abs{du}^2 dv_g\leq \varepsilon_1$$
then
$u\in H^2(B_{r/2}^g(p),g)$ with
\begin{equation} \label{est:D-S} \int_{B_{r/2}^g(p)}\abs{\nabla^2_g u}^2 +\abs{\nabla_g u}^4\,dv_g\leq \frac{C}{r^2}E(u,B_r^g(p))+C\int_{B_r^g(p)}\abs{f}^2 dv_g.\end{equation}
\end{prop}
Here and in the following we denote geodesic balls in $(C_0,g)$ by $B_r^g(p):=\{\tilde p\in C_0: d_g(\tilde p,p)<r\}$ and
compute the energy on a geodesic ball with respect to the corresponding metric unless indicated otherwise.
As we shall use this and the subsequent lemmas to control the map near the boundary, it is important to remark
\begin{rmk}\label{rem:no-lower-e-bound}
The above result is valid also without imposing a lower bound on $\ell$, and in particular also for the metric $G_0$ defined in \eqref{eq:G0}, as long as one considers only points contained in a compact subset $K\subset C_0\setminus\{0\}\times S^1$ and allows the constants to depend also on this set $K$.
Similarly, on compact sets $K\subset\subset C_0\setminus(\{0\}\times S^1)$ the a priori estimates derived in the subsequent Lemma \ref{lemma:H^2-general-boundary} and in Corollary \ref{cor:H^1-strong-estimates} are all valid for metrics $h_{b,\phi}^*G_\ell$ with $\ell\geq 0$
and $\abs{b_\pm}\leq b_0<1$, again with a constant that also depends on $K$.
\end{rmk}
As our target is Euclidean space which 'supports no bubbles', i.e. for which there are no non-trivial harmonic maps from $S^2$,
we can furthermore
exclude a concentration of the energy near general points of the boundary
\begin{lemma}\label{lemma:H^2-general-boundary}
To any numbers $\Lambda,E_0<\infty$, $b_0<1$, $d, \ell_0>0$ and any $\varepsilon>0$ there exists a radius $r>0$ such that the following holds true.
Let $u\in H^1_{\Gamma,*}(C_0)$ be a map of energy $E(u,g)\leq E_0$ which satisfies \eqref{eq:EL-minimiser-f} for a function
$f\in L^2(C_0,g)$ with $\norm{f}_{L^2}\leq \Lambda$,
a metric $g=h_{b,\phi}^*G_\ell$, with $\ell\geq \ell_0$ and $\abs{b^\pm}\leq b_0$ and variations $v\in T_u^+H_{\Gamma,*}^1(C_0)$ with $\supp(v)\subset C^*:=C_0\setminus \bigcup_{j,\pm}P_j^\pm$.
Then the energy is small
\begin{equation} \label{est:smallness-energy}
E(u,B_r^g(p))\leq \varepsilon\end{equation}
on balls around arbitrary points
$$p\in C^*(d)=C_{g}^*(d):=C_0\setminus \bigcup_{j,\pm}B_d^{g}(P_j^\pm).$$
In particular, the estimate
$$\norm{u}_{H^2(C^*(d))}\leq C\cdot E(u,g)+C\norm{f}_{L^2}^2$$
holds true with a constant $C$ that depends only on $\Lambda,E_0, \ell_0$, $b_0$ and $d$.
\end{lemma}
\begin{proof}
In order to prove the first part of the lemma we argue by contradiction.
So assume that for some numbers
$\varepsilon,d>0$, $b_0<1$ and $E_0, \Lambda<\infty $
there is a
sequence of $(u_i,g_i,f_i)$ as in the lemma
and a sequence of radii $\tilde r_i\to 0$ for which
$\sup_{x\in C_{g_i}^*(d)}E(u,B_{r_i}^{g_i}(x))>\varepsilon.$
Here we can of course assume that $\varepsilon\leq \varepsilon_1$, the number of Proposition \ref{prop:Duzaar-Scheven}.
We first prove
\textit{Claim:}
There exist radii $r_i\to 0$, points $p_i\in C_{g_i}^*( d/2)$
and numbers $\lambda_i\to\infty$ so that
$$E(u,B_{r_i}^{g_i}(p_i))=\varepsilon=\max_{p\in B_{\lambda_ir_i}^{g_i}(p_i)}E(u,B_{r_i}^{g_i}(p)).$$
To prove this claim let us first choose points $y_i$ and radii $r_i\to 0$ so that
$E(u,B_{r_i}^{g_i}(y_i))=\varepsilon=\max_{p\in C^*_{g_i}(d)}E(u,B_{r_i}^{g_i}(p)).$
Then the claim is trivially true for $p_i=y_i$ unless the points $y_i$ converge to the boundary (relative to $C_0$)
of the set $C_{g_i}^*(d)$ defined in the lemma.
So assume that, after passing to a subsequence,
$\text{dist}_{g_i}(y_i,P_j^{\pm})\to d $ for one of the point $P_j^\pm$, say for $P_0^+$.
We then consider
concentric annuli
$$A_i^k:=B_{d-2kr_i}^{g_i}\setminus B_{d-2(k+1)r_i}^{g_i}(P_0^+ )$$
constructed so that two balls of radius $r_i$ one having its centre in $A_k$ the other in $A_{k+2}$ are always disjoint.
Thus the number of such annuli that contain a point $p$ for which $E(u,B_{r_i}^{g_i}(p)) >\varepsilon$ can be no more than
$K_\varepsilon=2\lfloor\frac{E_0}{\varepsilon}\rfloor$.
We now choose $N_i\to \infty$ so that for $i$ large $ K_\varepsilon \cdot (N_i+1)r_i\leq d/2$ and observe that
$B_{d}^{g_i}\setminus B^{g_i}_{d/2}(x_0)$ must contain an annulus $\bigcup_{k=k_i}^{k_i+N_i-1} A_i^k$ of thickness
$N_i r_i$
which does not contain any point $p$ with $E(u,B_{r_i}^{g_i}(p))>\varepsilon$. Possibly reducing $r_i$ so that the maximum of
$p\mapsto E(u,B_{r_i}^{g_i}(p))$ on $B_{2d}^{g_i}\setminus B_{d-k_ir_i}^{g_i}(x_0)$ is equal to $\varepsilon$ and selecting $p_i$
to be a point at which this maximum is
achieved then implies the claim.
Based on this claim we now derive a contradiction using a standard blow-up argument where we distinguish between\\
\textit{Case 1:} \quad $\text{dist}_{g_i}(p_i,{\partial C_0})r_i^{-1}\to \infty$ for some subsequence
\\\\
\textit{Case 2:}\quad$\text{dist}(p_i,{\partial C_0})r_i^{-1}\leq C$ for some constant $C<\infty$
In the first case, we rescale the maps to maps $v_i(x)=u(\exp_{p_i}^{g_i}(r_ix))$ that are defined on larger and larger balls in $\ensuremath{{\mathbb R}}^2$.
We also observe that as always in such a bubbling argument the resulting metrics can be written as
$(\exp_{p_i}^{g_i}(r_i\cdot ))^*g_i=r_i^2 \tilde g_i$ for metrics $\tilde g_i$ that converge locally to the euclidean metric $g_{eucl}$.
In particular, the $H^2$ bounds on subsets of $(C_0,g_i)$ obtained in the previous lemmas
imply that the $H^2$-norm of the maps $v_i$
on compact subsets of $(\ensuremath{{\mathbb R}}^2,g_{eucl})$ are bounded uniformly and that
$\Delta v_i\to 0$ locally in $L^2$ since we have assumed $\norm{\Delta_{g_i}u_i}_{L^2}$ to be bounded.
Thus, after passing to a subsequence, we conclude that $v_i$ converges strongly in $H^1_{loc}$ and weakly in $H^2_{loc}$
to a limit $v_\infty:\ensuremath{{\mathbb R}}^2\to \ensuremath{{\mathbb R}}^n$ which is harmonic and has bounded energy and is thus constant. At the same time
$E(v_\infty,D_1(0))=\lim_{i\to\infty} E(v_i,D_1(0),\tilde g_i)=\lim_{i\to\infty} E(u_i, B_{r_i}^{g_i}(p_i))>\varepsilon$
resulting in the desired contradiction.
In the second case we rescale not around the points $p_j$ themselves,
but rather around their nearest point projection $\tilde p_j$ to the boundary of $C_0$. The resulting maps,
defined on larger and larger subsets of the halfplane $\mathbb{H}=\{y\in\ensuremath{{\mathbb R}}^2, y_2\geq 0\}$,
satisfy uniform $H^2$ bounds on compact sets of $\mathbb{H}$ and have energy at least $\varepsilon$ in the ball $B_{C+1}(0)\cap \mathbb{H}$.
We again obtain a harmonic limit with bounded energy, now defined only on the halfplane, but
furthermore constant on the axis $\partial \mathbb{H}$, since the maps $u_i\vert_{{\partial C_0}}$ are
equicontinous, compare Corollary \ref{lemma:courant-lebesgue-consequence} in the appendix.
Thus also this limit must be constant leading again to a contradiction.
This completes the proof of the first part of the lemma.
The second part now immediately follows from
Proposition \ref{prop:Duzaar-Scheven} and the first part if we choose $\varepsilon=\varepsilon_1$ to be the constant of that proposition.
\end{proof}
A further consequence of Proposition \ref{prop:Duzaar-Scheven} is
\begin{cor}\label{lemma:cont}
Let $u\in H^1_{\Gamma,*}(C_0)$ be a function for which
\eqref{eq:EL-minimiser-f} is satisfied for some
$f\in L^2(C_0)$ and $g=h_{b,\phi}^*G_\ell$, $\ell>0$, and for all
variations $v\in T_u^+H_{\Gamma,*}^1(C_0)$ with $\supp(v)\subset C^*:=C_0\setminus \bigcup_{j,\pm}P_j^\pm$. Then $u$
is continuous on all of $C_0$, in particular in the points $P_j^\pm$ where the three-point-condition is imposed.
Furthermore, let $u_i\in H^1_{\Gamma,*}(C_0)$ be any sequence of maps with uniformly bounded energy for which
\eqref{eq:EL-minimiser-f} is satisfied for functions
$f_i$ with $\sup_i \norm{f_i}_{L^2(M,g_i)}<\infty$
and metrics
$g_i=h_{b_i,\phi_i}^*G_{\ell_i}$ with
$\sup_{i} \abs{b_i^\pm}<1 \text{ and } \inf_i\ell_i>0$ and again for variations $v\in T_{u_i}^+H_{\Gamma,*}^1(C_0)$ with $\supp(v)\subset C^*$.
If no energy concentrates at the points $P_j^\pm$ in the sense that
\begin{equation} \label{ass:no-con}\lim_{r\to 0} \sup_{i} E(u_i, B^{g_i}_{r}(P_j^\pm))=0\quad j=0,1,2\end{equation}
then the maps $u_i$ are equicontinuous on all of $C_0$.
\end{cor}
\begin{proof} Let $u_i$ be a sequence of maps as described in the lemma.
As Lemma \ref{lemma:H^2-general-boundary} yields uniform $H^2$-bounds and thus equicontinuity for the $u_i$'s on
any compact subset of
$C^*$ it is enough to prove that to any number $\varepsilon>0$ there exists a radius $r_0$ so that
$$\mathop{{\mathrm{osc}}}\limits_{B_{r_0}^{g_i}(P_j^\pm)} u_i<\varepsilon,\quad j=0,1,2.
$$
To begin with, we recall from Corollary \ref{lemma:courant-lebesgue-consequence} that
the traces $u_i\vert_{\partial C_0}$ are equicontinuous so that
for $r_0$ sufficiently small
\begin{equation} \label{est:osc-u-infty-1}
\mathop{{\mathrm{osc}}}\limits_{B_{r_0}^{g_i}} u_i\leq \sup_{r\in (0,r_0]} \mathop{{\mathrm{osc}}}\limits_{\partial B_{r}^{g_i}}
u_i+\mathop{{\mathrm{osc}}}\limits_{B_{r_0}^{g_i}\cap {\partial C_0}} u_i\leq
\sup_{r\in (0,r_0]} \mathop{{\mathrm{osc}}}\limits_{\partial B_{r}^{g_i}}
u_i+
{\tfrac12} \varepsilon\end{equation}
where $\partial B_{r}\subset \mathring C_0$ is the boundary relative to $C_0$ and where we consider balls with centre $P_j^\pm$ unless indicated otherwise.
We bound the oscillation over $\partial B_r$
by deriving suitable
$H^2$-estimates on annuli $A_r^i:=B_{\frac54 r}^{g_i}\setminus B_{\frac34 r}^{g_i}$.
First of all, by \eqref{ass:no-con} we can assume that $r_0$ is small enough so that $E(u_i, B_{2r_0}^{g_i})<\varepsilon_1$, the number of
Proposition \ref{prop:Duzaar-Scheven}.
Then, given any $r\in (0,r_0]$ we cover the above annulus $A_r^i$ by balls $B_{r/4}^{g_i}(x_k)$ so that the corresponding balls with double the radius are contained in
$B_{2r}^{g_i}\setminus \{P_j^\pm\}$ and so that
no point is contained in more than $K$ of these larger balls $B_{r/2}^{g_i}(x_k)$, $K$ independent of $r$ and $i$.
Since \eqref{eq:E-L-eq-minimiser} is satisfied for $(u_i,g_i,f_i)$, at least for variations supported on $C^*$, we can apply
Proposition \ref{prop:Duzaar-Scheven} to bound
\begin{equation} \label{est: osc-est-annulus} \int_{B_{r/4}^{g_i}(x_k)}\abs{\nabla_{g_i}^2 u_i}^2 \,dv_{g_i}
\leq C r^{-2} E(u_i,B_{r/2}^{g_i}(x_k) )+C \norm{f_i}_{L^2(B_{r/2}^{g_i}(x_k))}^2\end{equation}
and thus also
\begin{equation} \label{est:H2-uinfty} \int_{A_r^i}\abs{\nabla_{g_i}^2 u_i}^2\, dv_{g_i}
\leq C r^{-2} E(u_i,B_{2r}^{g_i})+C\norm{f_i}_{L^2(B_{2r}^{g_i})}^2
\end{equation}
for a constant $C$ that is independent of $i$.
Observe that while the oscillation is invariant under the rescaling $\tilde u_i(x)=u_i(exp_{P_j^\pm}(rx))$, the left-hand-side of the above estimate transforms as
\begin{equation}\begin{aligned} \int_{A_1(0)}\abs{\nabla^2_{g_{eucl}} \tilde u_i }^2 dx
&\leq Cr^2 \int_{A_r}\abs{\nabla_{g_i}^2 u_i}^2\,dv_{g_i}
+C E(u_i,B_{2r}^{g_i})\\
&\leq C E(u_i,B_{2r_0}^{g_i})+Cr^2\norm{f_i}_{L^2(B_{2r_0}^{g_i})}^2.
\end{aligned}\end{equation}
Applying the Sobolev embedding theorem on a suitable subset of the fixed half-annulus $A_1(0)\subset\ensuremath{{\mathbb R}}^2$ thus allows us to conclude that
$$\big(\mathop{{\mathrm{osc}}}\limits_{\partial B_r^{g_i}} u_i\big)^2\leq C\cdot \int_{A_1(0)}\abs{\nabla_{g_{eucl}}^2 \tilde u_i}^2 +\abs{d \tilde u_i}^2 dx
\leq
C E(u_i,B_{2r_0}^{g_i})+Cr^2\norm{f_i}_{L^2(B_{2r_0}^{g_i})}^2$$
for every $r\in (0,r_0]$
again with constants that are independent of $i$.
Since the $L^2$-norms of $f_i$ are uniformly bounded and since we assumed that there is no concentration of energy at the points $P_j^\pm$ we can thus choose $r_0$ small enough so that the above expression is less than $\varepsilon/2$ as desired.
\end{proof}
\begin{rmk}\label{rem:no-lower-e-bound-2}
We observe that the claim of the above corollary remains true on arbitrary compact regions
of $C_0\setminus (\{0\}\times S^1)$ also without the assumption of a lower bound on $\ell$. Similarly, as all arguments are carried out locally, knowing that $u$ satisfies \eqref{eq:E-L-eq-minimiser} for variations supported in a set
$U\setminus \bigcup_{j,\pm}P_j^\pm$ is sufficient to conclude that $u$ is continous on every compact subset $K$ of $U$
with modulus of continuity depending only on $K$ and the bounds on $\abs{b^\pm}$, $E(u,g)$ as well as the local $L^2$-norm of $f$.
\end{rmk}
\subsection{No concentration of energy at points $P_j^{\pm}$.}
As the flow of metrics is determined in terms of the (non-local) projection of the Hopf-differential onto $\mathcal{V}$,
we need to exclude the possibility that a non-trivial amount of energy (and thus possibly of $L^1$ norm of $\Phi$) is concentrating
near one of the points $P_j^\pm$. Such a concentration of energy would be lost in a limiting process meaning that we could not expect that
the evolution of the limiting metric would be described by the projection of the limiting Hopf-differential.
To this end we prove the following key lemma
\begin{lemma}\label{lemma:key-est}
To any given numbers $\Lambda, M,E_0<\infty$, $b_0<1$ and $\varepsilon, d_0>0$ there exists a radius $r>0$ such that
the following holds true. Let
$u\in H_{\Gamma,*}^1(C_0)$ be a map with energy $E(u,g)\leq E_0$ that is bounded by $\norm{u}_{L^\infty} \leq M$ and that
weakly solves the differential inequality \eqref{eq:EL-minimiser-f} for a metric
$g=h_{b,\phi}^*G_\ell$ for which $\abs{b^\pm}\leq b_0$ and for a function $f$ with $\norm{f}_{L^2(C_0,g)}\leq \Lambda$.
Then the estimate
$$E(u,B_r^g(x_0))<\varepsilon$$
holds true for\textit{ every } point $x_0=(x,\theta)\in C_0$ with $\abs{x}\geq d_0$, in particular for $x_0=P_j^\pm$.
\end{lemma}
\begin{proof}
Thanks to Lemma \ref{lemma:H^2-general-boundary}
it is sufficient to establish the claim for the points $ P_j^\pm$, say for $x_0=P_0^+$.
So assume that the claim is wrong because for some fixed numbers
$\Lambda,M,E_0<\infty$ and $b_0<1$ there exists a number $\varepsilon_2>0$ such that
there are triples $(u_i,g_i=h_{b_i,\phi_i}^*G_{\ell_i},f_i)$ for which all the assumptions of the lemma are satisfied, but for which energy concentrates at $x_0$ in the sense
that
$$E(u_i,B_{r_i}^{g_i}(x_0))\geq \varepsilon_2$$
for a sequence of radii $r_i\to 0$.
We remark that the diffeomorphisms $h_{b,\phi}$ defined later on in section \ref{sect:diffeos} are such that
$h_{b,\phi}\equiv h_{b,\tilde \phi}$ in a neighbourhood of ${\partial C_0}$ if the parameters $\phi^\pm$ and
$\tilde\phi^\pm$ agree modulo $2\pi$.
Thus, after passing to a subsequence, the metrics converge smoothly to some
limiting metric $g=h_{b,\phi}^*G_\ell$, $\ell\geq 0$ at least in a neighbourhood $U$ of ${\partial C_0}$, compare
appendix \ref{appendix:collar}.
Away from the points $P^\pm_j$ we can apply Lemma \ref{lemma:H^2-general-boundary} to conclude that, after passing to a further subsequence, the maps converge on $U^*:=U\setminus \bigcup_{j,\pm} P_j^\pm$
in the sense that
$$u_i\to u_\infty \textit{ weakly in } H^2_{loc}(U^*) \text{ and strongly in } W^{1,p}_{loc}(U^*)\text { for every }p<\infty.$$
Furthermore, the uniform bounds on the energy
imply that the maps $u_i$ converge to $u_\infty$ weakly in $H^1$ on all of $U$ while the uniform $L^2$ bounds on $\norm{f_i}_{L^2}$
give weak $L^2$ convergence to a limit $f_\infty$ again on all of $U$.
Remark that here there is no need to specify with respect to which metrics $g_i$ the norms are computed as the metrics are uniformly equivalent.
Furthermore the traces $u_i\vert_{\partial C_0}$ converge uniformly to $u_\infty\vert_{{\partial C_0}}$ thanks to the equicontinuity obtained from the Courant-Lebesgue Lemma, so that $u_\infty$ can be extended to an element of $H^1_{\Gamma_*}(C_0)$.
We finally remark that the convergence of $(u_i,g_i,f_i)\to (u_\infty, g,f_\infty)$ implies that the differential
inequality \eqref{eq:EL-minimiser-f} is again satisfied
for $(u_\infty, g,f_\infty)$ at least for variations supported in $U^*$, see also appendix \ref{sect:Courant}.
The basic idea of the proof, working without modification only if the image of $u_i\vert_{{\partial C_0}\cap B_{r_0}^{g_i}(x_0)}$
happens to be the subset of a straight line, is now the following.
Since $u_\infty$ is an element of $H^1$, we can choose $r_0>0$ so that
$E(u_\infty,B_{2r_0}^g(x_0))$ is far smaller than $\varepsilon_2$ and thus in particular far smaller than the energy of the maps $u_i$ on this ball.
We would thus like to consider variations $u_\varepsilon$ of $u_i$ which have the form
$u_i+\varepsilon\cdot \lambda (u_\infty-u_i)$, $\lambda$ a cut-off function supported on $B_{2r_0}^{g_i}(x_0)$.
Then, if we could insert $v=\tfrac{d}{d\eps}\vert_{\eps=0} u_\varepsilon$ as test-function into \eqref{eq:E-L-eq-minimiser} we would get a contradiction since
the first term would give a
a negative contribution of roughly $-\varepsilon_2$ which could not
be compensated by the second possibly positive term. \\
Of course, having Plateau- rather than Dirichlet-boundary conditions, the maps $u_i+\varepsilon\cdot \lambda (u_\infty-u_i)$
are in general not in $H^1_\Gamma$. To obtain
an admissible variation we shall thus use the following lemma which is proved later on.
\begin{lemma}\label{lemma:make-boundary-flat}
Let $\Gamma$ be a regular closed Jordan curve in $\ensuremath{{\mathbb R}}^n$ of class at least $C^3$.
Then there exist constants $\hat r=\hat r(\Gamma)$ and $C=C(\Gamma)$ such that
for any point $p\in \Gamma$ and any $ r\in (0,\hat r(\Gamma))$
there exists a
$C^2$-diffeomorphism $\Phi:\ensuremath{{\mathbb R}}^n\to \ensuremath{{\mathbb R}}^n$ with
$\Phi=id$ outside of $B_{2r}(p)$ and $\Phi(p)=p$ which satisfies
\begin{equation} \label{est:make-boundary-flat}
r^2\norm{\Phi-id}_{L^\infty}+ r \norm{d \Phi-id}_{L^\infty}+\norm{d^2 \Phi}_{L^\infty}\leq C\end{equation}
and which
straightens out the curve $\Gamma$ in a neighbourhood of $p$ in the sense that
$$\Phi:\Gamma\cap B_{ r}(p)\to \{p\}+T_p\Gamma.$$
\end{lemma}
Returning to the proof of Lemma \ref{lemma:key-est} we let $\bar r\in (0,\hat r(\Gamma_+))$ be a fixed number that we determine later on.
Then, as the traces $u_i\vert_{{\partial C_0}}$ are equicontinuous
we can choose
$r_0>0$ small enough so that
\begin{equation} \label{cond:choice-r0}
u_i\big(B_{2r_0}^{g_i}(x_0)\cap {\partial C_0} \big)\subset B_{\bar r}(p_0)\end{equation}
where $p_0:=u_i(x_0)=\alpha^+(\theta_0)\in\ensuremath{{\mathbb R}}^n$ is prescribed by the three-point-condition.
Let now $\Phi$ be the map given by Lemma \ref{lemma:make-boundary-flat} which straightens out the boundary curve on the ball
$B_{\bar r}(p_0)\subset \ensuremath{{\mathbb R}}^n$
and let $\lambda\in C_0^\infty(B_{2r_0}^{g_i}(x_0))$ be a cut-off function which is identically $1$
on $B_{r_0}^{g_i}(x_0)\subset C_0$ and whose derivatives satisfy $\abs{\nabla_{g_i}^k \lambda_i}\leq Cr_0^{-k}$, $k=0,1,2$. Here the
constant $C$ is independent of $i$ since the metrics $g_i$ converge smoothly near the boundary.
We then define
$$u^\varepsilon_i:=\Phi^{-1}\circ \big[\Phi\circ u_i +\varepsilon\cdot \lambda_i \cdot \big( \Phi\circ u_\infty-\Phi\circ u_i\big)\big]$$
and claim that, for $\varepsilon>0$ sufficiently small, this is an admissible variation of $u_i$, i.e. that
$$u^\varepsilon_i\in H^{1}_{\Gamma,*}(C_0).$$
As $u^\varepsilon_i$ is clearly of class $H^1$, and as $u_i^\varepsilon\equiv u_i$ away from
$\supp(\lambda_i)$ it is enough to
show that
$u_i^\varepsilon(x_0)=p_0$
and that
$u_i^\varepsilon\vert_{{\partial C_0}\cap B^{g_i}_{2r_0}(x_0)}$ is a weakly monotone parametrisation of a subarc of $\Gamma$
(namely of $u_i({\partial C_0}\cap B^{g_i}_{2r_0}(x_0)\,)$).
The first claim is trivial since the three-point-condition is satisfied for all $u_i$'s and thus, since the traces converge uniformly, also for $u_\infty$.
The choice of $r_0$ and the properties of $\Phi$ imply furthermore that the restriction of $\Phi\circ u_i$ to ${\partial C_0}\cap B^{g_i}_{2r_0}(x_0)$ gives a
weakly monotone parametrisation of a segment in the tangent $t_{p_0,\Gamma}=\{p_0\}+T_{p_0}\Gamma$.
The same holds true also for
$\Phi\circ u_\infty$ and thus for any interpolation of these two maps so $u_i^\varepsilon$ is indeed an element of $H_{\Gamma,*}^1$. Thus $ w_i:=\tfrac{d}{d\eps}\vert_{\eps=0} u_i^\varepsilon\in T^+_{u_i}H_{\Gamma,*}^1$ is an admissible test function for the differential inequality
\begin{equation} \label{eq:to-contradict-EL-eq}
\int_{C_0}\langle du_i,dw_i\rangle\, dv_{g_i}+\int_{C_0}f_i\cdot w_i \,dv_{g_i}\geq 0.
\end{equation}
This will lead to
the desired contradiction
for $i$ sufficiently large provided the above construction is carried out on sufficiently small
balls $B_{r_0}^{g_i}$ and $B_{\bar r}(p_0)$.
To begin with, we remark that
\begin{equation}
w_i=\tfrac{d}{d\eps}\vert_{\eps=0} u_i^\varepsilon=\lambda_i\cdot(d\Phi^{-1})(\Phi(u_i))\cdot \big(\Phi(u_\infty )-\Phi(u_i)\big)
\end{equation}
is supported in the small ball $B_{2r_0}^{g_i}(x_0)$ and bounded
$\norm {w_i}_{L^\infty}\leq C_2$ by a constant depending only on $\Gamma^+$ and the bound $M$ imposed on the
$L^\infty$ norms of the $u_i$.
We can thus bound the second term in \eqref{eq:to-contradict-EL-eq} by
\begin{equation}\abs{ \int_{C_0} f_i\cdot w_i\, dv_{g_i}}\leq C_2\cdot \norm{f_i}_{L^1(B_{2r_0}^{g_i}(x_0))}
\leq C r_0 \norm{f_i}_{L^2(C_0)}\leq C\Lambda r_0<\tfrac{\varepsilon_2}{4} \end{equation}
provided $r_0$ is chosen sufficiently small.
The first term of \eqref{eq:to-contradict-EL-eq} on the other hand can be bounded from above by
\begin{equation}\begin{aligned} \label{est:main-term-key-lemma}
\int \langle du_i, dw_i\rangle \, dv \leq
&\int \lambda_i\cdot \bigg\langle \bigg[(d\Phi^{-1})(\Phi(u_i))\cdot d\Phi(u_\infty)\cdot du_\infty-du_i\bigg], du_i\bigg\rangle \,dv \\
&+ C \sup_{x\in B^{g_i}_{2r_0}}
\big(\abs{(d^2\Phi)(u_i(x))}\cdot \abs{\Phi(u_\infty(x))-\Phi(u_i(x)) }\big)\cdot \int \lambda_i \abs{du_i} ^2 dv \\
&+\frac{C}{r_0}\int_{\supp (d\lambda_i)}\abs{\Phi(u_\infty)-\Phi(u_i)}\cdot \abs{du_i} dv \\
&\leq-\big[1-\frac14-C\cdot\omega(r_0)\big] \cdot \int \lambda_i \cdot \abs{du_i} ^2 \,dv \\
&+C\cdot E(u_\infty, B_{2r_0}^{g_i})+C\cdot \norm{u_i-u_\infty}_{L^\infty(B_{2r_0}^{g_i}\setminus B_{r_0}^{g_i})}^2+CE(u_i,B_{2r_0}^{g_i}\setminus B_{r_0}^{g_i})
\end{aligned}\end{equation}
where all balls are to be taken with centre $x_0$, all integrals and norms are computed with respect to $g_i$ and where we set
$$\omega_i(r_0):=\sup_{x\in B_{2r_0}^{g_i}}\big(\abs{(d^2\Phi)(u_i(x))}\cdot \abs{u_\infty(x)-u_i(x) }\big).$$
Recall that $u_i\to u_\infty$ in $W_{loc}^{1,p}(U^*)$ for every $p<\infty$ and that the metrics converge.
Thus the penultimate term in \eqref{est:main-term-key-lemma} tends to zero as
$i\to \infty$, and is in particular $\leq \tfrac14\varepsilon_2$ for $i$ large. Furthermore, for $i$ large, the last term in \eqref{est:main-term-key-lemma} is bounded by
$C E(u_\infty, B_{3r_0}^g(x_0))+\tfrac14\varepsilon_2\leq \frac12 \varepsilon_2$, where
the last inequality holds provided $r_0$ is chosen sufficiently small.
Given that $\frac12\int \lambda_i \cdot \abs{du_i}^2 \,dv_{g_i}\geq E(u_i, B_{r_0}^{g_i})\geq \varepsilon_2,$
we can thus estimate (for $i$ large )
\begin{equation}\begin{aligned}
\int_{C_0}\langle du_i,dw_i\rangle\, dv_{g_i}+\int_{C_0}f_i\cdot w_i \,dv_{g_i}
&\leq -(\frac14-C\omega_i(r_0))\cdot \int \lambda_i \cdot \abs{du_i}^2 \,dv_{g_i},
\end{aligned}\end{equation}
which leads to the desired contradiction to \eqref{eq:to-contradict-EL-eq} provided we show
that $r_0>0$ can be chosen so that
\begin{equation} \label{claim:om}
C\omega_i(r_0)< \frac14 \text{ for $i$ large}.
\end{equation}
To prove this last claim we recall that
$d^2\Phi$ vanishes identically outside the ball $B_{2\bar r}(p)$.
This means that $\omega_i$ is obtained as supremum over a set on which the
oscillation of the function $u_i$ is a priori no more than $4\bar r$, for a number $\bar r$ that we can still reduce if needed.
This aspect of the construction is crucial
as we have no control on the behaviour of $u_i$ near $x_0$, so could in particular not hope for the
oscillation of $u_i$ over the full ball $B_{2r_0}^{g_i}$ to be uniformly small.
For $\bar r>0 $ sufficiently small and $i$ large, we can in particular estimate
\begin{equation}\begin{aligned}
C\omega_i(r_0)& \leq C \cdot \sup_{B_{2r_0}^{g_i}(x_0)\cap \supp(d^2\Phi\circ u_i)} \abs{u_i-p_0}
+C\cdot \sup_{B_{2r_0}^{g_i}(x_0)}\abs{u_\infty-p_0}\\
&\leq C \bar r+C\cdot \mathop{{\mathrm{osc}}}\limits_{B_{4r_0}^g(x_0)} u_\infty \leq \frac{1}{8}+C\cdot \mathop{{\mathrm{osc}}}\limits_{B_{4r_0}^g(x_0)} u_\infty.
\end{aligned}\end{equation}
We finally recall that $u_\infty$ satisfies \eqref{eq:E-L-eq-minimiser} for the function $f_\infty\in L^2$, the
limiting metric $g$ and for variations supported in $U^*$. As such Corollary \ref{lemma:cont} and Remark \ref{rem:no-lower-e-bound-2} imply that $u_\infty$ is
continuous at least in a neighbourhood of ${\partial C_0}$ and thus in particular in the points $P^\pm_j$.
Carrying out the above argument for a small enough radius $r_0$, which might depend on $u_\infty$ but is independent of $i$, we thus find that \eqref{claim:om} indeed holds.
\end{proof}
It remains to prove
\begin{proof}[Proof of Lemma \ref{lemma:make-boundary-flat}]
Let $\Gamma$ be a $C^3$ closed Jordan curve, let $p_0\in \Gamma\subset \ensuremath{{\mathbb R}}^n$, let $t_{p_0,\Gamma}=p_0+T_{p_0}\Gamma$ be the tangent to $\Gamma$ at
$p_0$ and let $\pi:\ensuremath{{\mathbb R}}^n\to t_{p_0,\Gamma}$ be the nearest point projection onto $t_{p_0,\Gamma}$.
Observe that for $\hat r>0$ chosen sufficiently small, in particular so that $\Gamma\cap B_{3\hat r}(p)$ is connected, this projection induces a $C^{2}$ bijection from
$\Gamma\cap B_{3\hat r}(p)$ to a segment in $t_{p_0,\Gamma}$.
Furthermore, after possibly reducing $\hat r$, we have that
$t_{p_0,\Gamma}\cap B_{2 r}(p)\subset \pi(\Gamma\cap B_{3 r}(p))$ for all $r\in (0,\hat r).$
We now consider
$\psi:t_{p_0,\Gamma} \cap B_{2\hat r}(p)\to \ensuremath{{\mathbb R}}^n$ defined by
$\psi=id\vert_{t_{p,\Gamma}}-\big(\pi\vert_{\Gamma\cap B_{3\hat r}(p)}\big)^{-1}$ and claim
that given any number $r>0$
\begin{equation} \label{def:Psi}
\Psi=\Psi_r:=id+ \lambda_r\cdot \psi\circ \pi=
id+\lambda_r\cdot \big[\pi-\big(\pi\vert_{\Gamma\cap B_{3\hat r}(p)}\big)^{-1}\circ \pi\big]
\end{equation}
gives the desired diffeomorphism. Here
$\lambda_r \in C^\infty_0(B_{2 r}(p),\ensuremath{{\mathbb R}})$ is given by a cut-off function which is
identically $1$ on $B_{ r}(p)$ and which satisfies the usual estimates of
$\abs{D^k\lambda }\leq C r^{-k}$, $k=0,1,2$. Since
$\pi(\Gamma\cap B_{2 r}(p))\subset
B_{2 r}(p)\cap t_{p,\Gamma}\subset \pi(\Gamma\cap B_{3\hat r}(p))$ the map $\Psi$ is well defined for any radius $r\in(0,\hat r)$.
To prove that $\Psi$ has the properties we asked for in Lemma \ref{lemma:make-boundary-flat} we first remark that
$\pi(p)=p$ and thus $\psi(p)=0$, i.e. $\Psi(p)=p$. More
generally, given any point $x\in \Gamma\cap B_{ r}(p)$ we obtain that
$$\Psi(x)=x+\lambda_r\cdot \big[\pi(x)-x\big]=\pi(x)$$
so $\Psi$ straightens the curve $\Gamma\cap B_{ r}(p)$ to a line as described in the lemma.
Since $\lambda_r$ and thus also $\Psi-id$ is supported in
$B_{2 r}(p)$, it remains to show that the estimate \eqref{est:make-boundary-flat}
claimed in the lemma holds true with a constant independent of $ r$.
Since $\Gamma$ is of class $C^3$ and since we project onto the tangent to $\Gamma$,
an estimate of the form
$\abs{\pi(x)-x}\leq C r^2$ is valid for all $x\in \Gamma\cap B_{3 r}(p)$ (recall that $ \Gamma\cap B_{3 r}(p)$ is connected).
We then use that we can write any $y\in B_{2 r}(p)\cap t_{p_0,\Gamma}$ as $y=\pi(x)$ for an $x\in \Gamma\cap B_{3 r}(p)$ to conclude that $\abs{\psi(y)}=\abs{\pi(x)-x}\leq C r^2$.
In particular
$$\norm{\Psi-id}_{L^\infty}\leq C
\sup_{B_{2 r}(p)}\abs{\psi\circ\pi}\leq C\sup_{B_{2 r(p)}\capt_{p_0,\Gamma}} \abs{\psi}\leq C r^2$$
holds true with a constant $C$ depending only on $\Gamma$.
We furthermore remark that since the derivative of the function $\psi$ (which is defined only on a line) vanishes in the point $p$
there exists a constant $C$ (again depending only on $\Gamma$) so that
$\norm{d\psi}_{L^\infty(t_{p,\Gamma}\cap B_{2r}(p))}\leq C r$. We can thus bound
\begin{equation*}\begin{aligned} \norm{d\Psi-id}_{L^\infty}&\leq \norm{d\lambda}_{L^\infty}\cdot \norm{\psi\circ \pi}_{L^\infty}+ \norm{d\psi}_{L^\infty}\cdot \norm{d\pi}_{L^\infty}\\
&\leq C r^{-1} \cdot r^2+C\norm{d\psi}_{L^\infty}\leq C r
\end{aligned}\end{equation*}
as well as
\begin{equation*}\begin{aligned} \norm{d^2\Psi}_{L^\infty}&\leq \norm{d^2\lambda}_{L^\infty}\cdot \norm{\psi\circ \pi}_{L^\infty}+ \norm{d\lambda}_{L^\infty}\cdot \norm{d\psi}_{L^\infty}
+ C\norm{d^2(\psi\circ \phi)}_{L^\infty}\leq C.
\end{aligned}\end{equation*}
\end{proof}
An important consequence we can derive from our
key Lemma \ref{lemma:key-est} is
\begin{cor}
\label{cor:H^1-strong-estimates}
Let $\Lambda, M<\infty$ and let $K$ be a compact subset of $\widetilde{\mathcal{M}}$. Then to every $\varepsilon>0$
there exists a constant $\delta>0$ such that the following holds true:
Let $u_1$ and $u_2$ be such that \eqref{eq:EL-minimiser-f} is satisfied for functions
$f_{i}$ with $\norm{f_{i}}_{L^2}\leq \Lambda$ and metrics $g_{i}\in K$.
Suppose furthermore that $E(u_i,g_i)\leq E_0$ and that $\norm{u_i}_{L^\infty}\leq M, i=1,2$.
Then if
$$
\norm{u_1-u_2}_{L^2(C_0,g)}\leq \delta$$
for some $g\in K$ then also
$$ \norm{u_1-u_2}_{H^1(C_0,g)}<\varepsilon.$$
\end{cor}
\begin{proof}
To begin with, we remark that all metrics in $K$ are uniformly equivalent since $K$ is compact. Thus it is sufficient to show the claim for norms $\norm{\cdot}_{L^2}$ and $\norm{\cdot}_{H^1}$ that are computed with respect to some fixed $g\in K$.
We then argue by contradiction. So assume there is a number $\varepsilon_1>0$ and triples $(u_1^k, g_1^k, f_1^k)_{k\in\ensuremath{{\mathbb N}}}$ as well $(u_2^k, g_2^k, f_2^k)_{k\in\ensuremath{{\mathbb N}}}$ so that all the assumptions of the lemma are satisfied (for each $k$) but for which
\begin{equation} \label{est:to-cont-cor} \norm {u_1^k-u_2^k}_{L^2(C_0)}\to 0\text{ while } \norm {u_1^k-u_2^k}_{H^1(C_0)}\geq \varepsilon.\end{equation}
Then, after passing to a subsequence
and using Proposition \ref{prop:Duzaar-Scheven} and Lemmas \ref{lemma:interior-minimiser} and \ref{lemma:H^2-general-boundary},
we find that locally on $C^*$ the maps
$u_{1,2}^k$ converge to limits $u_{1,2}$ strongly in $H^1$, where by construction these two limits must agree.
In particular, for any fixed number $r>0$ we have
\begin{equation} \label{est:difference-H1-contradict}
\norm{u_1^k-u_2^k}_{H^1\big(C_0\setminus \bigcup_{\pm,j}B_{r}^g(P_j^\pm)\big)} \to 0\end{equation}
as $k\to \infty$.
We can then choose $r>0$ so small that Lemma \ref{lemma:key-est}, combined with the equivalence of the metrics in $K$, implies that
$$\norm{u_i^k}_{H^1(B^{g}_r(x_0))}\leq C\norm{u_i^k}_{H^1(B^{g_i^k}_{Cr}(x_0))} \leq \frac{\varepsilon}{24} \text{ for all } k\in\ensuremath{{\mathbb N}}, \,i=1,2, \text{ and } x_0\in C_0.$$
Applied for the points $P_j^\pm$ and combined with \eqref{est:difference-H1-contradict}
this contradicts \eqref{est:to-cont-cor}.
\end{proof}
\subsection{Convergence of the time-discretisation scheme}
Given any initial data
$(u_0,g_0)\in H_{\Gamma,*}^1(C_0)\times \widetilde{\mathcal{M}}$ we consider
the approximate solutions of Teichm\"uller harmonic map flow
$(u_j,g_j):=(u^{h_j},g^{h_j})$, $h_j=2^{-j}$, obtained by the time discretisation scheme described in section \ref{sect:time-dicr}
We can analyse the maps $u_j$ using the results of the previous section since
$u_j(t)$ can be seen as a stationary solution of
\begin{equation}
\label{eq:E-L-time-discret}
\int \langle du_j(t), dw\rangle \, dv_{\tilde g_j(t)}
+\int D_t^{h_j} u_j(t)\cdot w \,dv_{\tilde g_j(t)} \geq 0 \text{ for }
w\in T_{u_j(t)}^+H^1_{\Gamma,*} \text{ and } t\in[0,T],
\end{equation}
where $\tilde g_j$ is piecewise constant so that $\tilde g_j(t)=g_j(t_k^{h_j})$, $t\in [t_k^{h_j}, t_{k+1}^{h_j})$.
Based on the results of the previous sections we can pass to a subsequence, still denoted by
$(u_j, g_j)$, of approximate solutions which converge to a limiting curve
of maps and metrics $(u,g)$ as described below.
To begin with, we claim that $u_j$ converges uniformly in time with respect to $L^2$ in space, i.e. that
\begin{equation} \label{est:conv-u-linfty} \sup_{t\in[0,T]}\norm{u_j(t)-u(t)}_{L^2}\to 0,\quad \text{ for } j\to \infty\end{equation}
and that $u\in C^0([0,T],L^2(C_0))$.
Here and in the following there is no need to specify with respect to which metric on $C_0$ the above convergence is to be understood as
all the considered metrics are uniformly equivalent on the interval $[0,T]$ we consider, compare \eqref{ass:inj-apriori}.
To prove this claim we let
$t\mapsto \tilde u_j(t)$ be piecewise linear with
$\tilde u_j(t_k^{h_j})=u_j(t_k^{h_j})$ for every $k$.
Then \eqref{est:energy-ineq-approx-scheme} gives
uniform $C_t^{0,\frac12}L_x^2$ estimates for $\tilde u_j(t)$,
namely
\begin{equation}\begin{aligned}
\norm{\tilde u_j(t_2)-\tilde u_j(t_1)}_{L^2}\leq \int_{t_1}^{t_2}\norm{D^{h_j}_t u_j}_{L^2} dt &\leq (t_2-t_1)^{1/2} \big(\int_{t_1}^{t_2}\norm{D^{h_j}_t u_j}_{L^2}^2 dt\big)^{1/2}\\
&\leq (2E_0)^{1/2}(t_2-t_1)^{1/2}.
\end{aligned}\end{equation}
Thus, after passing to a subsequence, $\tilde u_j$ converges in $C^0_tL^2_x$ to a limit $u$.
Furthermore, again by \eqref{est:energy-ineq-approx-scheme},
$$\sup_t\norm{u_j(t)-\tilde u_j(t)}\leq \sup_k\int_{t_k^{h_j}}^{t_{k+1}^{h_j}}\norm{D_t^{h_j} u_j}_{L^2} dt\leq h_j^{1/2}(2E_0)^{1/2}\to 0 \text{ for } j\to \infty$$
leading to \eqref{est:conv-u-linfty}.
Remark that the uniform bounds on the energy furthermore imply that the limiting map is in $L^\infty([0,T], H^1(M,g_0))$
and that the spatial derivatives converge
$$ d u_j \rightharpoonup d u \text{ weakly in } L^2([0,T]\times C_0).$$
We recall that the traces of $u_j$ on ${\partial C_0}$ are uniformly equicontinuous,
so we furthermore have that
$$u_j(t)\vert_{\partial C_0}\to u(t)\vert_{\partial C_0} \text{ uniformly on } {\partial C_0} \text{ for all } t\in [0,T],$$
where we stress that we do not claim that the rate of this convergence is uniform in time.
Additionally, the energy inequality \eqref{est:energy-ineq-approx-scheme} gives
uniform $L^2(C_0\times [0,T-h_j])$ estimates for the difference quotients
$D_t^{h_j}u_j$. Consequently, $u$ is weakly differentiable in time on
$C_0\times [0,T]$ with $D_t^{h_j}u_{j}\rightharpoonup \partial_t u$ in $L^2(C_0\times [0,T])$.
For the metric component
$g_j$ we can apply Lemma \ref{lemma:Lip-Projection} to get uniform $C^{0,1}$ estimates in time with respect to any $C^k$
metric in space since the
$L^1$ norm of the Hopf-differential is bounded in terms of the (non-increasing) energy.
We can thus get convergence of $g_j\to g$ in $C^{0,\alpha}([0,T],C^k(C_0))$, $\alpha<1$, with the
limiting curve being again of class $C^{0,1}_t C^k_x$ and thus in particular differentiable in time for almost every $t\in[0,T]$.
We shall now prove that the limit $(u,g)$ obtained in this way gives the desired solution of
Teichm\"uller harmonic map flow, namely that
\begin{prop}
\label{prop:convergence-discret}
Let $(u^{h_j},g^{h_j})$ be a sequence of approximate solutions to a fixed initial data $(u_0,g_0)$
converging as described above to some limiting curve $(u,g)$ as $h_j\to 0$.
Then the limit $(u,g)$ is a stationary weak solution of Teichm\"uller harmonic map flow
which also satisfies the energy-inequality (for a.e. $t_1<t_2$).
\end{prop}
We remark that while $g$ is clearly again an admissible curve,
we need to prove that its derivative is actually
given by the projection of the Hopf-differential of the limit.
As the projection operator is non-local,
for this part of the proof the key lemma \ref{lemma:key-est} and its Corollary \ref{cor:H^1-strong-estimates} are crucial
to get strong $H^1$ convergence for the map $u$ and thus strong $L^1$ convergence of the Hopf-differential
on all of $C_0$, in particular also near the points $P_j^\pm$
where Proposition
\ref{prop:Duzaar-Scheven} does not apply.
Conversely, the analysis of the map component can be carried out
very similarly to the work of Duzaar and Scheven \cite{D-S} and is indeed
less involved then the corresponding arguments
since the metric is well controlled and since our equation for the map is linear.
\begin{proof}[Proof of Proposition \ref{prop:convergence-discret}]
We first infer from the energy inequality \eqref{est:energy-ineq-approx-scheme}
that for any $\Lambda<\infty$ the set of times
$$A_j^\Lambda:=\{t\in [0,T] \text{ so that } \norm{D_t^{h_j}u_j(t)}_{L^2}\leq \Lambda\}$$
has measure
$\mathcal{L}^1(A_j^\Lambda)\geq T-\frac{CE_0} {\Lambda^2}$, so in particular
$\mathcal{L}^1(A^\Lambda)\geq T-\frac{CE_0} {\Lambda^2}$ also
for
$$A^\Lambda=\limsup A^\Lambda_j=\bigcap_{n=1}^\infty \bigcup_{j=n}^\infty A^\Lambda_j.$$
Recall that the maps $u_j(t)$ satisfy \eqref{eq:EL-minimiser-f} for $f=D_t^{h_j} u_j(t)$
so it is precisely bounds of the form
$\norm{D_t^{h_j}u_j(t)}_{L^2}\leq \Lambda$ that are required
in order to be able to apply the results derived in the previous section.
We begin by analysing the metric component. To prove that $g$ indeed solves \eqref{eq:evol-g} we show that it agrees with the solution $\hat g(t)\in C^{0,1}_t\widetilde{\mathcal{M}}$ of the initial value problem
$$\partial_t \hat g=P_{\hat g}^\mathcal{V}(Re((\Phi(u,\hat g)) ,\qquad \hat g(0)=g_0.$$
We will prove this claim based on Corollary \ref{cor:H^1-strong-estimates}.
So let $K$ be the set of metrics $h_{b,\phi}^*G_\ell$ satisfying \eqref{ass:inj-apriori} as well as \eqref{est:apriori-upper-ell} and let $M$ be a bound on
the $L^\infty$ norm of the initial map which therefore also serves as
bound for $\norm{u_j}_{L^\infty}$, compare Lemma \ref{lemma:ex-minimiser}.
Then given any numbers $\varepsilon>0$ and $\Lambda<\infty$ we let $\delta>0$ be the number
given by Corollary \ref{cor:H^1-strong-estimates} and select
$j_0$ so that
$$\sup_{t\in [0,T]}\norm{u_j(t)-u_k(t)}_{L^2}\leq \delta \text{ for } j,k\geq j_0.$$
This implies that
$$\norm{u_k(t)-u_j(t)}_{H^1(C_0)}\leq \varepsilon \text{ for all } t\in A_j^\Lambda\cap A_k^\Lambda \text{ and } j,k\geq j_0$$
which in turn yields the same bound for $\norm{u(t)-u_j(t)}_{H^1(C_0)}$ for $t\in A_j^\Lambda\cap A^\Lambda$.
For times in $A_j^\Lambda\cap A^\Lambda$ the
difference of the Hopf-differentials is thus controlled by
\begin{equation}\begin{aligned}
\norm{Re_{\hat g}(\Phi(u,\hat g)(t))-Re_{g_j}(\Phi(u_j,g_j)(t))}_{L^1}&\leq
C\cdot \norm{ (g_j-\hat g)(t)}_{C^0}\cdot E_0\\
&\,+C\cdot E_0^\frac{1}{2}\cdot \norm{(u_j-u)(t)}_{H^1(C_0)}\\
&\leq C\cdot \norm{ (g_j-\hat g)(t)}_{C^0}+C\cdot \varepsilon,
\end{aligned}\end{equation}
c.f. \eqref{eq:Hopf-coord}.
Here and in the following constants $C$ may depend on $E_0$, $K$ and $T$ but not on $\varepsilon$ or $j$ unless indicated otherwise.
Based on Lemma \ref{lemma:Lip-Projection}
we can thus conclude that
$$\norm{\partial_t(\hat g-g_j)(t)}_{C^k}\leq C\cdot \norm{(\hat g-g_j)(t)}_{C^k}+C\varepsilon$$
for any $t\in A_j^\Lambda\cap A^\Lambda$ and for $j\geq j_0(\Lambda, \varepsilon)$.
On the other hand, we can always bound the norms of
$\partial_t g_j$ and $\partial_t \hat g$
by $C\cdot E_0$ and we shall use these trivial bounds on the set
$[0,T]\setminus (A_{h_j}^\Lambda\cap A^\Lambda)$ on which we cannot apply any of the results of the previous section since we lack the necessary control on
the inhomogeneity of \eqref{eq:EL-minimiser-f}.
Combining these two cases, we obtain that for almost every $t$
$$\norm{\partial_t(\hat g-g_j)(t)}_{C^k}\leq C\cdot \norm{(\hat g-g_j)(t)}_{C^k}+h_j^\Lambda(t)$$
where
$$h^\Lambda_j=C\cdot \varepsilon+CE_0\chi_{[0,T]\setminus (A^\Lambda_j\cap A^\Lambda)}.$$
Based on Gronvall's Lemma, we can thus conclude that
for any $t\in [0,T]$ and $j\geq j_0(\varepsilon,\Lambda)$
$$\norm{(\hat g-g_j)(t)}_{C^k}\leq e^{CT}\cdot \int_0^t\abs{h^\Lambda_j} dt\leq C \varepsilon+\frac{C}{\Lambda^2}.$$
Choosing $\Lambda\to \infty$ and $\varepsilon\to 0$ and corresponding values of $j_0(\varepsilon,\Lambda)\to \infty$
yields the claim that
$g^j\to \hat g$ uniformly and thus that $g=\hat g$ is indeed the solution of \eqref{eq:evol-g}.
We now turn to the analysis of the map component where we follow largely the arguments of \cite{D-S}.
To begin with, we observe that for almost every time $t\in[0,T]$ there exists a number
$\Lambda<\infty$ such that $t\in A^\Lambda$. Choosing a subsequence along which
$\norm{D_t^{h_j} u_j(t)}_{L^2(C_0)}\to \liminf_{j\to \infty}\norm{D_t^{h_j} u_j(t)}_{L^2(C_0)}\ \leq \Lambda$
we conclude that
$$u_j(t)\to u(t) $$
converges not only strongly in $L^2$, but thanks to Corollary \ref{cor:H^1-strong-estimates}
indeed strongly in $H^1$ on all of $C_0$ and, thanks to Lemma \ref{lemma:H^2-general-boundary},
also weakly in
$H^2_{loc}(C^*)$ where $C^*:=C_0\setminus \bigcup \{P_j^\pm\}$. We stress that the choice of this subsequence is allowed to depend on the time $t$ we are considering.
We furthermore remark that combining the uniform $H^2$-estimates for $u_j$ valid on subsets
$\Omega\subset\subset C^*$
with the uniform convergence of the metrics $g_j\to g$ yields that also $\Delta_{g_j(t)} u_j(t)\rightharpoonup \Delta_{g(t)}u(t)$ weakly in $L^2(\Omega)$ for each such $\Omega$.
Using the $L^2$ bound on $\Delta_{g_j(t)} u_j(t)=D_t^{h_j} u_j$ valid for $t\in A^\Lambda_j$ we
can thus conclude that
$$\norm{\Delta_{g(t)} u(t)}_{L^2(\Omega)}\leq \lim_{j\to\infty}\norm{\Delta_{g_j(t)} u_j(t)}_{L^2(\Omega)}\leq \Lambda \text{ for each }
\Omega\subset\subset C^*.$$
Passing to the limit in both the Euler-Lagrange-equation and the stationarity condition, cf. appendix \ref{sect:Courant},
thus yields that $u(t)$ is a stationary solution of
\eqref{eq:EL-minimiser-f}
for a function $f(t)=\Delta_{g(t)}u(t)$ whose $L^2(C_0,g)$-norm is again bounded by
$\Lambda$, or indeed more precisely by $ \liminf_{j\to \infty}\norm{D_t^{h_j} u_j(t)}_{L^2(C_0,g_j)}$.
Repeating this argument for a.e. $t\in [0,T]$ we thus obtain a function $f:[0,T]\times C_0$
which must have bounded $L^2(C_0\times [0,T])$-norm since
$$\norm{f}_{L^2(C_0\times [0,T])}^2 \leq \int_0^T \liminf_{j\to\infty} \norm{D_t^{h_j}u_j(t)}_{L^2(C_0)}^2 dt\leq
\liminf_{j\to\infty} \norm{D_t^{h_j}u_j}_{L^2(C_0\times [0,T])}^2\leq 2 E_0$$
where the last inequality follows from \eqref{est:energy-ineq-approx-scheme}.
We now wish to show that $f$ agrees with the time derivative of $u$.
To this end, proceeding as in \cite{D-S},
we set
$$\tilde u_{j}^\Lambda(t)=\begin{cases}
u(t)&\text{ if } t \in B^\Lambda_{j}:=[0,T]\setminus A^\Lambda_{j}\\
u_j(t)& \text{ if }t\in A^\Lambda_{j}
\end{cases}
\text{ and } \tilde f_{j}^\Lambda(t)=\begin{cases}
f(t)&\text{ if } t\in B^\Lambda_{j}\\
D_t^{h_j}u^{j}(t)& \text{ if }t\in A^\Lambda_{j}
\end{cases}$$
in order to obtain a new sequence of pairs satisfying \eqref{eq:EL-minimiser-f} for $\tilde g_j(t)$
but for which the estimate
$\norm{\tilde f_j^\Lambda(t)}_{L^2(C_0)}\leq \Lambda$
is now satisfied for every $j$ and every $t\in [0,T]$.
We first claim that for any sequence $ \Lambda_j\to \infty$
$$\tilde f^{\Lambda_j}_{j}\rightharpoonup \partial_t u \text{ weakly in } L^2( C_0\times[0,T]) \text{ for }j\to \infty, $$
or, as we know that $D_t^{h_j}u_j\rightharpoonup \partial_t u$, equivalently
$\tilde f^{\Lambda_j}_j-D_t^{h_j}u_j\rightharpoonup 0$. Indeed, given any function $\phi\in L^2(C_0\times[0,T])$
we have that
\begin{equation}\begin{aligned}
\int_{[0,T]\times C_0}\phi\cdot \big(D_t^{h_j}u_j-\tilde f^{\Lambda_j}_j\big) dv_g\, dt &
\leq \norm{\phi}_{L^2(B^{\Lambda_j}_j\times C_0)}\cdot
\big(\norm{\tilde f^{\Lambda_j}_j}_{L^2([0,T]\times C_0)}+
\norm{D_t^{h_j}u_j}_{L^2([0,T]\times C_0)}\big)\\
& \leq CE_0^{\tfrac12} \norm{\phi}_{L^2(B^{\Lambda_j}_{j}\times C_0)}
\end{aligned}\end{equation}
which tends to zero as $\Lambda_j\to \infty$ since the measure $\mathcal{L}(B_j^{\Lambda_j})\to 0$.
At the same time we claim that $d\tilde u_j^{\Lambda_j}$ converges not just weakly, which would be evident from an argument just as carried out above, but indeed strongly in $L^2([0,T] \times C_0)$ to $d u$.
Indeed, let us first consider
$d u-d \tilde u_j^\Lambda$ for a fixed number $\Lambda$.
Since this difference vanishes on $B_j^\Lambda$ and since we can apply Corollary \ref{cor:H^1-strong-estimates} to $ u(t)$ and $\tilde u_j^\Lambda(t)$ for $j$ large enough (so that these
maps are $L^2$ close) we obtain that
\begin{equation} \label{est:converg-nabla-utilde1} \norm{d u-d \tilde u_j^\Lambda}_{L^2(C_0\times[0,T])}\to 0 \text{ for every fixed } \Lambda. \end{equation}
But the set of times $A_j^{\Lambda_j}\Delta A_j^\Lambda\subset B_j^{\Lambda_j}\cup B_j^\Lambda$, on which $d \tilde u_j^\Lambda$ and $d \tilde u_j^{\Lambda_j}$ do not agree, has measure no more
than $C(\Lambda^{-2}+\Lambda_j^{-2})$. Combined with the uniform bound on the energy this means that \eqref{est:converg-nabla-utilde1} suffices to conclude
that indeed $\norm{d u-d \tilde u_j^{\Lambda_j}}_{L^2(C_0\times[0,T])}\to 0$.
Thanks to the strong $H^1$ convergence we can furthermore approximate each test function $w\in L^2([0,T],T_u^+H_{\Gamma,*}^1(C_0))$ by elements
$w_i\in L^2([0,T],T_{u_i}^+H_{\Gamma,*}^1(C_0))$ in the sense that $\norm{dw-dw_i}_{L^2(C_0\times [0,T])}\to 0$, compare appendix \ref{sect:Courant}. We thus conclude that
$$\int_0^T\int_{C_0}du\cdot dw +\partial_t u\cdot w dv_g dt\geq 0$$
for all such $w$. As we have already shown that $g$ satisfies \eqref{eq:evol-g}, we thus obtain that $(u,g)$ is indeed a weak solution of Teichm\"uller harmonic map flow.
Knowing that $d u_j^{ \Lambda_j}$ converges strongly and not just weakly furthermore implies convergence of the Hopf-differentials $\Phi(u_j,g_j)$ to $ \Phi(u,g)$
in $L^1$, allowing us to pass to the limit in the stationarity condition to conclude that \eqref{eq:stationar} holds true for almost every time $t$.
It remains to show that the energy inequality holds true (for almost every pair of times $t_1$, $t_2$.)
So let $t_1\in [0,T]$ be any time which is contained in one of the sets $A^\Lambda$, $\Lambda<\infty$, i.e. for which there exists a sequence of $h_j$ so that
$D_t^{h_j}u_j(t_1)$ is bounded in $L^2$. For this subsequence of $u_j(t_1)$, Corollary \ref{cor:H^1-strong-estimates} implies strong $H^1$ convergence
and thus in particular that $E((u_j,g_j)(t_1))\to E((u,g)(t_1))$.
We then recall that $u_j$ satisfies the energy inequality
$$E(u_j,g_j)(t_1)-E(u_j,g_j)(t)\geq \int_{t_1}^t \norm{\partial_t g_j}_{L^2(C_0,g_j)}^2 dt+{\tfrac12} \int_{t_1}^t\norm{\partial_t u_j}_{L^2(M,\tilde g_j(t))}^2 dt
$$
for all times $t\in [0,T]$, $t>t_1$ and $u_j(t)$ converges at least weakly in $H^1$ to
$u(t)$ (a further strongly convergent subsequence could be found for almost every $t$ but is not needed).
We thus obtain that for every $t\geq t_1$
$$E(u,g)(t_1)-E(u,g)(t)\geq \int_{t_1}^t \norm{\partial_t g}_{L^2(C_0,g)}^2 dt+{\tfrac12} \int_{t_1}^t\norm{\partial_t u}_{L^2(C_0,g)}^2 dt.$$
\end{proof}
\section{Long time existence}\label{sect:long-time}
\subsection{A priori estimates for the metric component}
Before we can analyse admissible curves of metrics in more detail we finally need to decide how to select the family of diffeomorphisms $h_{b,\phi}$
which we use to compensate for the lost degrees of freedom of the three-point-condition.
Rather than just writing down a possible family, we shall first describe which properties we require in the present context
of flowing to minimal surfaces.
We will then later give an example of such a family but do not claim that this choice is in any way unique.
To begin with, in order to obtain solutions of the flow that exist for all times we need the following $L^2$-completeness property:
\begin{lemma}\label{lemma:diffeo-properties-1}
Let $(h_{b,\phi})$ be the family of diffeomorphisms defined in \eqref{def:diffeos}. Assume that
$g(t)=h_{b(t),\phi(t)}^*G_{\ell(t)}$, $t\in [0,T)$, is such that
the diffeomorphisms $h_{b,\phi}$ become singular as $t\to T$, i.e. so that (at least) one of the values $\abs{b^\pm}\to 1$ or $\abs{\phi^\pm}\to\infty$ as $t\to T$.
Then
$$ \int_0^T\norm{\partial_t g}_{L^2(C_0,g)} dt =\infty.$$
\end{lemma}
A further requirement we want to impose in preparation for the asymptotic analysis carried out later in section \ref{sect:asympt} is
\begin{lemma}\label{lemma:diffeo-properties-2}
Let $(h_{b,\phi})$ be the family of diffeomorphisms defined in \eqref{def:diffeos} and let
$\chi(b,\phi)$ the space of generating vectorfields of $h_{b,\phi}$. Then
$$\Gamma(TC_0)=\Gamma(TC_0)_*\oplus h_{b,\phi}^*\chi(b,\phi)$$
for all $(b,\phi)\in \Omega_{h}:=D_1(0)^2\times \ensuremath{{\mathbb R}}^2$.
\end{lemma}
Here and in the following $\chi(b,\phi)\subset \Gamma(TC_0)$ is the 6 dimensional vectorspace spanned by
the vectorfields generating the diffeomorphisms $h_{b,\phi}$, i.e. by $Y_{\phi^\pm}(b,\phi)$
characterised by
\begin{equation} \label{eq:generating-vf}
\tfrac{d}{d\phi^\pm}h_{b,\phi}(s,\theta)=Y_{\phi^\pm}(b,\phi)(h_{b,\phi}(s,\theta)),\end{equation}
together with the vectorfields $Y_{Re(b^{\pm})}(b,\phi)$ and $Y_{Im(b^\pm)}(b,\phi)$ defined by the analogue of \eqref{eq:generating-vf}
or, if $b^\pm\neq 0$, equivalently
together with the vectorfields $Y_{\abs{b^{\pm}}}(b,\phi)$, $Y_{Arg(b^\pm)}(b,\phi)$ corresponding to variations of the absolute value respectively the argument of $b^\pm$.
The final property we shall ask of the diffeomorphisms $h_{b,\phi}$ is that their support is disjoint from the middle geodesic.
This will have the advantage that the modification by these diffeomorphisms does not interfere with the analysis of a possible collapse of the central geodesic, cf. Lemma \ref{lemma:inj}.
It will furthermore prove to be useful to choose the $h_{b,\phi}$ so that the
support of the induced variations of the metrics with respect to the parameters $\phi^\pm$ on the one hand and $b^\pm$ on the other hand
are
disjoint.
\subsubsection{Choice of diffeomorphisms}\label{sect:diffeos}
A simple way of assuring that our diffeomorphisms satisfy Lemma \ref{lemma:diffeo-properties-2} is to choose them as restrictions of M\"obiustransforms on the boundary of $C_0$.
Given numbers $ \phi^\pm\in\ensuremath{{\mathbb R}}$ and $ b^\pm\in\ensuremath{{\mathbb C}}$ with $\abs{b^\pm}<1$ we consider
the functions $ f_{b^\pm,\phi^\pm}:\ensuremath{{\mathbb R}}\to \ensuremath{{\mathbb R}}$ which are induced by the M\"obiustransforms $M_{b^\pm,\phi^\pm}$,
i.e. chosen so that $f_{0,0}=id$ and $e^{i f_{b^\pm, \phi^\pm}(\theta)}=M_{b^\pm, \phi^\pm}(e^{i\theta})$, where
$$M_{b,\phi}(z):=e^{i\phi} \frac{z+b}{1+\bar b\cdot z}, \text{ for } b\in D_1(0)\subset \ensuremath{{\mathbb C}}, \phi\in \ensuremath{{\mathbb R}}, \, z\in\overline{D_1(0)}\subset \ensuremath{{\mathbb C}}.$$
We then extend the maps induced by $f_{b^\pm,\phi^\pm}$ on ${\partial C_0}$
to a suitable diffeomorphim $h_{b,\phi}$ on the whole cylinder. Namely, we choose $\lambda_{1,2}$ as smooth cut-off functions such that
$\lambda_1\equiv 0$ on $[-1,\tfrac34]$ with $\lambda_1\equiv 1$ on $[\tfrac78,1]$ while $\lambda_2\equiv 0$ on $[-1,\tfrac12]$ with $\lambda_2\equiv 1$ on $[\tfrac58,1]$.
We then define $h_{b,\phi}:C_0\to C_0$, $b=(b^-,b^+), \phi=(\phi^-,\phi^+)$ through
\begin{equation}
\label{def:diffeos}
h_{b,\phi}(s,\theta)= \big(s, \lambda_1(s)\cdot f_{b^+,\phi^+}(\theta)
+\big(1-\lambda_1(s)\big)\cdot \big(\theta+\lambda_2(s)\cdot \phi^+)\big)\end{equation}
if $s \geq 0$ respectively by the analogue formula, replacing $(b^+,\phi^+)$ with $(b^-,\phi^-)$ and $s$ by $-s$, if $s\leq 0$.
Since $f_{b,\phi}(\theta)=f_{b,0}(\theta)+\phi$ this formula reduces to
$h_{b,\phi}(s,\theta)=(s,\theta+\lambda_1\cdot (f_b(\theta)-\theta)+\lambda_2\cdot\phi_+)$
where we write for short $f_b$ for $f_{b,\phi=0}$.
In order to show that this family of diffeomorphisms satisfies Lemma \ref{lemma:diffeo-properties-1}
we observe that a change of one of the parameters, say of $\abs{b^+}$, induces
a change of the metric of
$$\tfrac{d}{d\abs{b^+}} \big(h_{b,\phi}^*G\big)=h_{b,\phi}^*L_{Y_{\abs{b^+}}}G=L_{h_{b,\phi}^*Y_{\abs{b^+}}}g$$
for $g=h_{b,\phi}^*G$
and that the resulting Lie derivatives of the collar metrics $G$ satisfy the following estimates
\begin{lemma}\label{lemma:variations-parameter}
Let $(h_{b,\phi})$ be the family of diffeomorphisms defined in \eqref{def:diffeos},
let $Y_{\abs{b^\pm}}$, $Y_{Arg(b^\pm)}$ and $Y_{\phi^\pm}$ be its generating vectorfields
and let $(G_\ell)$ be the family of metrics defined in Lemma \ref{lemma:horizontal-family} for some fixed number $\eta>0$.
Then to any number $L_0<\infty$ there exist constants $C_{1,2,3,4}\in \ensuremath{{\mathbb R}}^+$
(depending only on $L_0$ and $\eta$) such that the following estimates hold true for any metric $G=G_\ell$ with $\ell<L_0$ and any $(b,\phi)$ (where we assume that $b^+\neq 0$ for the first estimate)
\begin{equation} \label{est:variations-parameters}
\norm{L_{Y_{\abs{b^+}}}G}_{L^2(C_0,G)}\geq \frac{C_1}{1-\abs{b^+}}-C_2 \quad
\text{and} \quad
C_3\leq \norm{L_{Y_{\phi^+}}G}_{L^2(C_0,G)}\leq C_4
\end{equation}
Furthermore
$$L_{Y_{\abs{b^+}}}G, \, L_{Y_{Arg(b^+)}}G \text{ and } L_{Y_{\phi^+}}G $$
are $L^2(C_0,G)$-orthogonal to each other.
\end{lemma}
The claims made above for variations with respect to
$\phi^+$ and $b^+$ are of course valid also for variations with respect to $b^-$ and $\phi^-$ and from the construction it is evident that variations with respect to
$(\phi^+,b^+)$ on the one hand and
$(\phi^-,b^-)$ on the other hand have disjoint support so result in Lie-derivatives that are trivially orthogonal.
With regards to the proof of this lemma,
we observe that the
orthogonality of $ L_{Y_{\phi^+}}G$ to the variations with respect to $b^+$ follows since $Y_{\phi^+}$
is given by the Killing field
$\tfrac{\partial}{\partial\theta }$ on the support of $Y_{\abs{b^+}}$ and $Y_{Arg(b^+)}$.
The orthogonality of $L_{Y_{\abs{b^+}}}G $ and $ L_{Y_{Arg(b^+)}}G $ on the other hand will follows from the different symmetry properties of these two tensors
The proof of this last part and of the estimates claimed in the lemma is not difficult though a bit technical so we include it in the appendix \ref{appendix:diffeos}.
As a consequence of Lemma \ref{lemma:variations-parameter}
we can now prove Lemma \ref{lemma:diffeo-properties-1} for this particular choice of diffeomorphism
\begin{proof}[Proof of Lemma \ref{lemma:diffeo-properties-1}]
Let $g(\cdot)=h_{b(\cdot),\phi(\cdot)}^*G_{\ell(\cdot)}$ be an admissible curve of metrics
with $ L_{L^2}g(\cdot)=\int_0^T\norm{\partial_t g}_{L^2(C_0,g)} dt<\infty.$
We first recall that $Re({\cal H}} %{\Upsilon(g))$ is orthogonal to $\{L_Xg\}$ so that
both $\norm{\partial_t G}_{L^2(C_0,G)}\leq \norm{\partial_t g(\cdot)}_{L^2(C_0,g)}$ and $\norm{\tfrac{d}{d\eps}\vert_{\eps=0} h_{b(\cdot+\varepsilon),\phi(\cdot+\varepsilon)}G(\cdot)}_{L^2(C_0,g)}\leq \norm{\partial_t g(\cdot)}_{L^2(C_0,g)}$
must have finite integral over $[0,T)$.
On the one hand, this implies that
$\ell(t)$ is bounded from above by a constant $\bar L$ depending only on the initial metric and $L_{L^2}g(\cdot)$, compare \eqref{est:evol-l-large}.
Using the orthogonality of $L_{Y_{\phi^+}}G$ to the variations generated by a change of any of the other parameters,
as well as estimate \eqref{est:variations-parameters}, we know furthermore that
$$\norm{\partial_t g}_{L^2(C_0,g)}\geq \abs{\tfrac{d}{dt} \phi^+}\cdot \norm{\tfrac{d}{d\phi^+} h_{b,\phi}^* G}_{L^2(C_0, h_{b,\phi}^* G)}
=\abs{\tfrac{d}{dt} \phi^+}\cdot \norm{L_{Y_{\phi^+}}G}_{L^2(C_0,G)}\geq C_3\cdot \abs{\tfrac{d}{dt}\phi^+}\,$$
where $C_3$ depends only on the upper bound on $\ell$ obtained above.
This implies that $\phi^+$, and by the same argument also $\phi^-$, remains bounded.
So consider instead the behaviour of $b^{\pm}$, say of $b^+$.
The orthogonality relations of Lemma \ref{lemma:variations-parameter}
combined with \eqref{est:variations-parameters} imply
\begin{equation}\begin{aligned}
\norm{\partial_t g}_{L^2(C_0,g)}^2&\geq
\norm{\tfrac{d}{dt}{ \abs{b^+}}\cdot L_{Y_{\abs{b^+}}}G}_{L^2(C_0,G)}^2
\geq \big[\tfrac{C_1}{1-\abs{b^+}}-C_2\big]^2
\abs{\tfrac{d}{dt}{ \abs{b^+}}}^2.
\end{aligned}\end{equation}
In particular, for $\abs{b^+}$ sufficiently close to $1$, an estimate of the form
$$\abs{\tfrac{d}{dt} \log(1-\abs{b^+})}\leq C \norm {\partial_t g}_{L^2(M,g)}$$
holds true which prevents $b^+$ from reaching
$\partial D_1(0)$ if the curve $g$ has finite $L^2$ length.
\end{proof}
We remark that the Teichm\"uller space of the cylinder equipped with the metric that results from representing conformal structures by hyperbolic metrics $f^*G_\ell^\eta$ as described in
Lemmas \ref{lemma:conf}, \ref{lemma:metric-cyl-1} and \ref{lemma:splitting} is not complete. Indeed, as explained in appendix \ref{appendix:collar},
for general curves in $\widetilde{\mathcal{M}}$ and $\ell$ small we can only bound
$$\abs{\frac{d\ell}{dt} }\leq C\cdot \norm{\partial_t g}_{L^2}\cdot \ell^{1/2}$$
so that the possibility that $\ell\to 0$ is not excluded for curves of finite length.
Nonetheless, for Teichm\"uller harmonic map flow a degeneration of the metric in finite time is excluded since we can prove
\begin{lemma}\label{lemma:inj}
To any numbers $\ell_1>0$ and $M,T,E_0<\infty$ there exist constant $C<\infty$ and $\varepsilon_0>0$ such that the following holds true.
Let $(u_0,g_0)\in H_{\Gamma,*}^1(C_0)\times \widetilde{\mathcal{M}}$ be any initial data so that
$E(u_0,g_0)\leq E_0$, $\norm{u_0}_{L^\infty}\leq M$
and $\inj(C_0,g_0)\geq 2\ell_1$ and let $(u,g)$ be the corresponding stationary weak solution of Teichm\"uller
harmonic map flow whose existence on some interval $[0,T_1)$ is assured by Proposition \ref{prop:convergence-discret}.
Then the weighted energy is bounded by
\newcommand{{\mathcal{I}}}{{\mathcal{I}}}
\begin{equation} \label{def:I}
I(t):=\int_{C_0} e(u(t),g(t))\rho^{-2}(t) dv_{g(t)} \leq C,\end{equation}
and the injectivity radius by
$$\inj(C_0,g)\geq \varepsilon_0$$
for every $t\in [0,\min(T,T_1))$.
Here $e(u,g)={\tfrac12} \abs{du}_{g}^2$ is the energy density while
$\rho(t)(x,\theta)=\rho_{\ell(t)}(s_{\ell(t)}(x))$ is the conformal factor of the hyperbolic collar.
\end{lemma}
This result is essentially a consequence of results proven in \cite{RT2} for Teichm\"uller harmonic map flow from closed surfaces into non-positively curved targets because the action of the diffeomorphisms $h_{b,\phi}$ does not affect the region near the central geodesic.
We also recall that while the
metric component $g$ is in general not smooth, it is Lipschitz continuous in time with respect to any metric in space. So while
$u$ might not be smooth in the interior of $C_0$ it also satisfies such $C^{0,1}_tC^k_x$ bounds, at least away from time $t=0$ which is enough to apply the arguments of \cite{RT2}
on almost every time-slice.
\begin{proof}
We first explain why a bound on the weighted energy $I$ results in a bound on the injectivity radius.
We recall that the evolution of $g(t)=h_{(b,\phi)(t)}^*G(t)$, $G(t)=G_{\ell(t)}$,
splits $L^2$-orthogonally
into the projection of the Hopf-differential onto the subspace $\{L_{h_{b,\phi}^*X}g, \, X\in \chi(b,\phi)\}$ and into the projection onto $Re({\cal H}} %{\Upsilon(g))$ and thus that
$\partial_t G=Re(P^{{\cal H}} %{\Upsilon(G)}(\Phi))$.
For the cylinder the space ${\cal H}} %{\Upsilon(G)$ consists only of tensors that can be written as
$$a_0\cdot dz^2 , a_0\in\ensuremath{{\mathbb R}}$$
with respect to collar coordinates $z=s+i\theta, (s,\theta)\in[-Y(\ell),Y(\ell)]\times S^1$, so
\begin{equation}\label{eq:ptG} \partial_t G=Re\bigg( \langle \Phi, \frac{dz^2}{\norm{dz^2}_{L^2}^2}\rangle_{L^2} dz^2\bigg).\end{equation}
We furthermore recall that if $\partial_tg =a_0dz^2$, then the length of the central geodesic evolves according to $\frac{d\ell}{dt}=-\frac{2\pi^2}{\ell} a_0$,
compare \eqref{eq:ddl}.
For small values of $\ell$, say $\ell\in (0,\ell_0)$, the norm
$\norm{dz^2}_{L^2}^2$ is given by \eqref{eq:dz-small-l}
which, once combined with \eqref{eq:ptG} and \eqref{eq:ddl}
and \eqref{eq:Linfty-dz}, implies that
\begin{equation*}
\abs{\frac{d}{dt}\log\ell +\frac{1}{16\pi^3}\cdot \ell \int_{\ensuremath{{\cal C}}(\ell)} (\abs{u_s}^2-\abs{u_\theta}^2)\rho^{-2} dsd\theta}\leq C\ell\norm{\Phi(u,g)}_{L^1}, \end{equation*}
$\ensuremath{{\cal C}}(\ell)=[-Y(\ell),Y(\ell)]\times S^1$, compare also section 5 of \cite{RT2}.
In particular $\abs{\tfrac{d}{dt} \log(\ell)}\leq C\ell+\ell I(t)$ is bounded if $I(t)$ is bounded, resulting in the desired lower bound
on $\ell=2\inj(C_0,G_\ell)$.
For the proof of \eqref{def:I} we can use results derived in sections 3 and 5 of \cite{RT2}. Namely,
the results of \cite[section 3]{RT2}, in particular Proposition 3.6, give angular energy estimates for maps from hyperbolic collars
into compact non-positively curved targets. Since we know that $\norm{u}_{L^\infty}\leq M$ these results apply without change also to the present situation.
As in section 5 of \cite{RT2} we consider a cut-off version of the weighted energy given by
\newcommand{{\mathcal{I}}}{{\mathcal{I}}}
\begin{equation} \label{I-zeppelin}
{\mathcal{I}}(t):=\int_{\ensuremath{{\cal C}}(\ell(t))} e(u(t),g(t))\rho^{-2}(t)\varphi(\rho(t))dv_{g(t)},\end{equation}
where $\varphi\in C_0^\infty([0,2\delta),[0,1])$ is a cut-off function with $\varphi\equiv 1$ on $[0,\delta]$, and where $\delta>0$ can be chosen to be any fixed number.
We remark that we can choose $\delta$ sufficiently small, so that
the diffeomorphism
$h_{b,\phi}$ agree with the identity on the support of $\varphi\circ \rho$, compare \eqref{eq:Xde} and the subsequent comments.
For such a choice of $\delta$ we conclude that the evolution of the metric reduces to
$\partial_t g=Re(c(t)dz^2)$ on the relevant region, i.e. on the support of $\varphi\circ \rho$.
Consequently the Bochner formula for the energy density given in Lemma 5.2 of \cite{RT2} and the evolution equation for the conformal factor described in Lemma 5.4 of \cite{RT2}
apply without change and could indeed be further simplified
as $\partial_t g$ evolves not just by any holomorphic quadratic differential but by $c_0dz^2$.
Then arguing precisely as in the proof of Lemma 5.1 in \cite{RT2} we obtain that
\begin{equation} \label{est:I-zeppelin}
\bigg|\tfrac{d}{dt} \log(1+{\mathcal{I}})\bigg| \leq C\left(1+\norm{\Delta_g u}_{L^2(C_0,g)}^2\right)\leq C\left(1+\norm{\partial_t u}_{L^2(C_0,g)}^2\right)\end{equation}
with $C$ depending only on $M$, the initial energy and the choice of $\delta$.
Thus ${\mathcal{I}}$ and consequently also $I\leq {\mathcal{I}}+C_\delta E_0$ is bounded uniformly on every compact time interval as claimed in the lemma.
\end{proof}
From Lemma \ref{lemma:inj} we thus conclude that for arbitrary initial data $(u_0,g_0)\in H^1_\Gamma\times \ensuremath{{\mathcal M}}_{-1}$ solutions to Teichm\"uller harmonic map flow from the cylinder indeed extist for all times as claimed in Theorem
\ref{thm1}.
\section{Asymptotics of global solutions}\label{sect:asympt}
We now turn to the proof of the second main result of the paper, the asymptotic convergence for the global weak solutions whose existence we have just proven.
In the present work we analyse the asymptotics in case that the three-point-condition does not
degenerate as $t\to \infty$, i.e. for solutions for which the parameters $b^\pm$ remain bounded away from $\partial D_1$ (at least for a subsequence $t_j\to \infty$). The
remaining
case of the asymptotics will be analysed in future work.
So let $(u,g)$ be a global stationary weak solution of Teichm\"uller harmonic map flow which satisfies the energy inequality.
We then choose $t_i\to \infty$ such that the stationarity condition is satisfied for the times $t_i$ and so that
\begin{equation}
\label{eq:asympt-subseq-1}
\norm{\Delta_{g(t_i)} u(t_i)}_{L^2(C_0,g(t_i))}\to 0 ,
\end{equation}
\begin{equation} \label{eq:asympt-subseq-2}
\norm{P_g^{\mathcal{V}}(Re(\Phi(u,g)(t_i))}_{L^2(C_0,g(t_i))} \to 0
\end{equation}
and $\abs{b^{\pm}(t_i)}\nrightarrow 1$
as $i\to \infty$.
We can thus pass to a subsequence to achieve that
\begin{equation} \label{est:conv-parameters}
b_i^\pm=b^\pm(t_i)\to b_\infty^\pm\in D_1(0)\subset \ensuremath{{\mathbb C}} \text{ and }
\tilde \phi_i^\pm:= \phi_i^\pm-n_i^\pm \cdot 2\pi\to \phi_\infty^\pm,\end{equation}
as $i\to\infty$ where $n_i=\lfloor\tfrac{\phi(t_i)^\pm}{2\pi}\rfloor$.
We then pull-back the map and metric by the diffeomorphisms
$f_i:=h_{0,2\pi n_i}$. Remark that since $f_i= h_{b_i,\phi_i}^{-1}\circ h_{b_i,\tilde \phi_i}$ the resulting metrics
$g_i:=f_i^*g(t_i)$ are simply given by
$g_i=h_{b_i,\tilde \phi_i}^*G_{\ell(t_i)}$.
We furthermore recall that $f_i$ agrees with the identity in a neighbourhood of the boundary so that the pulled-back maps $u_i=
u(t_i)\circ f_i$ still satisfy the three-point-condition, i.e. are again elements of $H^1_{\Gamma,*}(C_0)$.
Remark that \eqref{eq:asympt-subseq-1} and \eqref{eq:asympt-subseq-2}
are satisfied also for $(u_i,g_i)$ and both the differential inequality
\begin{equation} \label{eq:diff-ineq-for-asympt}
\int \langle du_i, dw\rangle\, dv_{g_i}+\int \Delta_{g_i}u_i\cdot w \, dv_{g_i}\geq 0 \text{ for all } w\in T^+_{u_i}H^1_{\Gamma,*}\end{equation}
and the stationarity equation
\begin{equation} \label{eq:stationarity-for-asympt}
\int Re(\Phi(u_i,g_i))\cdot L_Xg_i +\Delta_{g_i}u_i\cdot d u_i(X) \,dv_g=0 \text{ for all } X\in \Gamma_*(TC_0)\end{equation}
hold true.
To prove convergence of $(u_i,g_i)$ to a critical point of area as described in Theorem \ref{thm2} we now distinguish between the
non-degenerate case, $\ell(t_i)\nrightarrow 0$, in which we will obtain a (branched) minimal immersion parametrised over a cylinder, and the degenerate case $\ell(t_i)\to 0$ in which the surface splits into
two minimal discs.
We begin with
\begin{proof}[Proof of Theorem \ref{thm2} part (i): The non-degenerate case]
After possibly passing to a further subsequence we can assume that
$\ell_i=\ell(t_i)\to \ell_\infty>0$ which implies that the metrics converge $g_i\to g_\infty=h_{b_\infty,\phi_\infty}^*
G_{\ell_\infty}$ smoothly on $C_0$.
Furthermore, as $u_i$ is a solution of \eqref{eq:diff-ineq-for-asympt} for which $\norm{\Delta_{g_i} u_i}_{L^2}$ is bounded,
we can
apply the $H^2$-estimates of Lemma \ref{lemma:H^2-general-boundary} away from $P_j^\pm$ as well as the $H^1$ estimates of
Lemma \ref{lemma:key-est} and Corollary \ref{cor:H^1-strong-estimates} on the whole of $C_0$. We conclude that
a subsequence of the $u_i$ converges to a limit $u_\infty\in H^2_{loc}(C^*)\cap H^1(C_0)$ where the obtained convergence is
weak $H^2_{\text{loc}}$ and strong $W^{1,p}_{\text{loc}}$ convergence on $C^*:=C_0\setminus \bigcup P_j^\pm$ as well as strong $H^1$ convergence on all of $C_0$. Furthermore, Corollary \ref{lemma:cont} implies that the maps $u_i$ are equicontinuous
near the boundary, and thus by the $H^2$ estimates on all of $C_0$, so that the $u_i$ converge uniformly on $C_0$. In particular, $u_\infty\in C^0(C_0)$.
The above convergence implies not only that
\begin{equation} \label{eq:limit-harmonic}
\Delta_{g_\infty} u_\infty\equiv 0 \text{ on } C_0
\end{equation}
and consequently that the Hopf-differential of the limit is holomorphic, but furthermore that
the Hopf-differentials $\Phi(u_i,g_i)\to \Phi(u_\infty,g_\infty)$ converge in $L^1$ on the whole cylinder $C_0$.
From Lemma \ref{lemma:Lip-Projection} we thus obtain that
\begin{equation} \label{eq:proj-Hopf-diff-limit}
P_{g_\infty}^\mathcal{V}(\Phi(u_\infty,g_\infty))=\lim_{i\to \infty } P_{g_i}^\mathcal{V}(\Phi(u_i,g_i))=0.\end{equation}
On the one hand, this implies that
\begin{equation} \label{eq:Phi-infty}
\int Re(\Phi(u_\infty,g_\infty))\cdot L_Y g_\infty dv_{g_\infty} = 0\end{equation}
holds true for the vectorfields
$Y\in h_{b_\infty,\tilde\phi_\infty}^*\chi(b_\infty, \tilde \phi_\infty)$ generating the diffeomorphisms $h_{b,\phi}$.
On the other hand, the convergence of the Hopf-differential allows us to pass to the limit in the stationarity condition to conclude that
\eqref{eq:Phi-infty}
holds true also for all vectorfields $X\in \Gamma(TC_0)_*$. Thus, by Lemma \ref{lemma:diffeo-properties-2}, we find that
\eqref{eq:Phi-infty} is indeed true for any smooth vectorfield on $C_0$ which is tangential to $\partial C_0$ on $\partial C_0$.
We now show that this forces $\Phi_\infty=\Phi(u_\infty, g_\infty)$ to be of the form $cdz^2$ for some $c\in\ensuremath{{\mathbb R}}$.
Remark that if $\Phi_\infty$ were smooth (or even just $W^{1,1}$) upto the boundary,
we could directly combine \eqref{eq:Phi-infty} with Stokes theorem to conclude that
$\Phi_\infty$ is real on the boundary and then to conclude that $\Phi_\infty=cdz^2, c\in \ensuremath{{\mathbb R}}$.
However, while $\Phi_\infty$ is holomorphic
and thus smooth in the interior as well as in $W^{1,p}$, $p<2$ in a neighbourhood of general boundary points, near the points $P_j^\pm$ we know a priori only that $\Phi_\infty$ is in $L^1$.
Thus $\Phi_\infty$ could have a pole at such a point
and we need to proceed with more care.
Given any fixed $X\in \Gamma(TC_0)$
we use that \eqref{eq:Phi-infty}
implies that
\begin{equation} \label{est:int-X-infty}
\abs{\int_{[-1+\varepsilon,1-\varepsilon]\times S^1} L_Xg_\infty \cdot Re(\Phi_\infty) \,dv_{g_\infty}}\to 0 \text{ as }\varepsilon\to 0\end{equation}
and we initially work on such subcylinders where $\Phi_\infty$ is smooth.
Recall that $L_Xg$ can be identified with $-\delta_g^*X$, where $\delta_g^*$ is the $L^2$-adjoint of the divergence operator and that the real part
of a holomorphic quadratic differential is divergence free.
So, switching to collar coordinates $(s,\theta)\in [-Y_\infty,Y_\infty]\times S^1$, $Y_\infty=Y(\ell_\infty)$,
and applying Stokes theorem to
\eqref{est:int-X-infty} yields
\begin{equation} \label{est:real-on-boundary-eps}
\abs{\int_{\{Y_\infty-\tilde\varepsilon\}\times
S^1} Re(\Phi_\infty)(\tfrac{\partial}{\partial s},X) \rho^{-2}d\theta
-\int_{\{-Y_\infty+\tilde\varepsilon\}\times S^1} Re(\Phi_\infty)(\tfrac{\partial}{\partial s},X) \rho^{-2}d\theta}\to 0 \text{ as } \tilde \varepsilon\to 0\end{equation}
where $\rho=\rho_{\ell_\infty}(s)$.
Away from the boundary of $ [-Y,Y]\times S^1$
we now represent $\Phi$ by its Fourier expansion $\Phi_\infty=\sum_{n\in\ensuremath{{\mathbb Z}}} (a_n+ib_n) e^{ns}e^{ni\theta}$, $a_n,b_n\in \ensuremath{{\mathbb R}}$
and apply \eqref{est:real-on-boundary-eps} for vectorfields of the form
$X=\lambda^{\pm}(s) \cos(m\theta)\cdot\frac{\partial}{\partial \theta}$ and $X=\lambda^{\pm}(s) \sin(m\theta)\cdot\frac{\partial}{\partial \theta}$, $m\in\ensuremath{{\mathbb N}}$,
where
$\lambda^\pm$ are cut-off functions that are identically one
in a neighbourhood of $\pm 1$ and that vanish say on
$\{ \pm s\leq {\tfrac12}\}$.
Passing to the limit $\tilde\varepsilon\to 0$ in \eqref{est:real-on-boundary-eps}
yields
$$b_me^{m Y_\infty}=b_{-m}e^{- mY_\infty} \text{ and }
b_me^{-m Y_\infty}=b_{m}e^{mY_\infty} $$
as well as
$$a_me^{ m Y_\infty}=-a_{-m}e^{-m\cdot Y_\infty} \text{ and }
a_me^{ -m Y_\infty}=-a_{-m}e^{m\cdot Y_\infty}. $$
so that all Fourier coefficients except for $c_0=a_0+ib_0$ need to be zero. Of course, testing with $X=\lambda^\pm\cdot \frac{\partial}{\partial \theta}$ furthermore gives that $b_0=0$ and thus
that $ \Phi_\infty=a_0dz^2$ is indeed an element of ${\cal H}} %{\Upsilon(C_0)$.
But \eqref{eq:proj-Hopf-diff-limit} also implies that
the projection of $\Phi_\infty$
onto ${\cal H}} %{\Upsilon(g_\infty)=\{cdz^2, c\in\ensuremath{{\mathbb R}}\}$
vanishes so $\Phi_\infty$ must vanish meaning that $u_\infty$ must be (weakly) conformal.
Thus $u_\infty$ is a weakly conformal and harmonic map which spans $\Gamma$ and can thus in particular not be constant so
must be a (possibly branched) minimal immersion \cite{GOR}.
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm2} part (ii): The degenerate case:]
Let $(u_i,g_i)$ be as above and assume now that $\ell_i\to 0$.
We let $C^+=(0,1]\times S^1$ and $C^-=[-1,0)\times S^1$
and observe that the subcylinders $(C^\pm, g_i)$ are isometric to
$$\big([0,Y_i)\times S^1, \rho_\ell^2(Y_i-s)\cdot (ds^2+d\theta^2)\big),\quad Y_i=Y(\ell_i)
$$
with an isometry given by
$\tilde f_{\ell_i}^\pm:(x,\theta)\mapsto
(Y_i\mp s_{\ell_i}(x),\theta).$
We remark that
$\rho_\ell(Y(\ell)-s)\to \frac{1}{2\pi s+ \eta}$ as
$ \ell\to 0$ locally smoothly on $[0,\infty)\times S^1.$
At the same time
$\tilde f_\ell^ \pm$
converges locally to the diffeomorphism
$\tilde f_\infty^\pm:C^\pm\to [0,\infty)\times S^1$ given by
$f_\infty^\pm(x,\theta)=
(\frac{2\pi}{\ell_0}\cdot \tan\big(\frac\pi2\mp \frac{\ell_0 x}{2\pi}\big)-2\pi\eta,\theta)$.
Thus the metrics $g_i$ converge smoothly locally to a metric $g_\infty$ that is isometric to
the hyperbolic cusp
$$([0,\infty)\times S^1, \rho_0^2(s)\cdot (ds^2+d\theta^2))$$
described in the theorem.
At the same time, we get subconvergence for the maps $u_i=u(t_i)\circ h_{0,2\pi n_i}\to u_\infty$
as described in the theorem since the bounds on $\abs{b_i^\pm}$ allow us to apply the $H^2$-estimates of Lemma \ref{lemma:interior-minimiser} and Lemma \ref{lemma:H^2-general-boundary} as well as the $H^1$ estimate of Lemma \ref{lemma:key-est} and the equicontinuity result of Corollary \ref{lemma:cont} on every compact subset of $C^\pm$,
see also Remarks \ref{rem:no-lower-e-bound} and \ref{rem:no-lower-e-bound-2}.
The above convergence of the maps and metrics implies in particular that
$\Delta u_\infty=0$ and thus that
$\Phi_\infty=\Phi(u_\infty, g_\infty)$ is holomorphic on $C^\pm$.
We then observe that
the local convergence of the map and metric on $C^\pm$
allows us to pass to the limit in the stationarity condition to conclude that
\begin{equation} \label{eq:Phi-infty-deg}
\int_{C^\pm}L_Xg_\infty \cdot Re(\Phi_\infty)dv_{g_\infty}=0
\end{equation}
provided we only consider vectorfields $X\in \Gamma(TC_0)_*$ whose support is contained in one of the subcylinders $C^\pm$.
We recall that
also the support of the
vectorfields $Y_{\phi^\pm}$, $Y_{Re(b^\pm)}$ and $Y_{Im(b^\pm)}$
generating the diffeomorphisms $h_{b,\phi}$ is contained in
$C^\pm$ and that the corresponding projection of $\Phi$ tends to zero, compare \eqref{eq:asympt-subseq-2}. So local strong convergence of $u_i$ in $H^1$ and consequently of $\Phi_i$ in $L^1$ implies that
\eqref{eq:Phi-infty-deg} holds true also for these particular vectorfields, and thus,
by Lemma \ref{lemma:diffeo-properties-2}, indeed for
arbitrary vectorfields $X\in \Gamma(TC_0)$ whose support
is contained in either $C^+$ or $C^-$.
As the Fourier expansion of $\Phi(u_\infty,g_\infty)$ on $(C^\pm,g_\infty)\simeq [0,\infty)\times S^1$ cannot have any exponentially growing terms (since $\norm{\Phi_\infty}_{L^1}<CE_0<\infty$)
we can then argue as in the previous proof to conclude that in collar coordinates
$(s,\theta)\in [0,\infty)\times S^1$
$$\Phi_\infty=c^\pm (ds+id\theta)^2 \quad \text{ for some } c^\pm\in \ensuremath{{\mathbb R}}.$$
Pulling $\Phi_\infty$ back to the punctured disc $D^*=D_1(0)\setminus\{0\}$
through a conformal diffeomorphism
$f:(D^*,g_{eucl})\to ([0,\infty)\times S^1, ds^2+d\theta^2)$
we thus find that the Hopf-differential of $u_\infty$ is represented by
$c^\pm z^{-2} dz^2$, $c^\pm\in \ensuremath{{\mathbb R}}$ for $z\in D^*$.
But the limiting map $u_\infty$ can be seen as a harmonic map from the punctured disc $(D^*,g_{eucl})$
whose energy is finite (since the energy is conformally invariant) which implies that $u_\infty$ can be continued smoothly across the puncture, compare \cite{SU}. Thus
$\Phi_\infty$ must be smooth on all of $D_1(0)$
and must thus vanish identically.
This proves that the maps $u_\infty^\pm=u_\infty\vert_{C^\pm}
$ extend to
weakly conformal harmonic maps from the disc and thus give two (possibly branched) minimal immersions with
each of them spanning one of the boundary curves $\Gamma^\pm$.
\end{proof}
|
\section{Introduction}
Initiated by K.~Walker \cite{Wa},
the classification of polygon spaces with $n$ edges (see \cite{HR} and \secref{S.polyg} hereafter)
involves chambers delimited by a hyperplane arrangement in $({\mathbb R}_{>0})^n$ and
so-called virtual genetic codes. The number $c(n)$ of chambers
modulo coordinate permutations and the number $v(n)$ of virtual genetic codes were computed by several authors
(see \cite{HRweb}) and the currently known figures are as follows
\sk{4}
\hskip -0mm
\begin{minipage}{110mm}
\begin{tabular}{c|cccccccccccc} \small
$n$ & \footnotesize 3&\footnotesize 4&\footnotesize 5&\
\footnotesize 6 &\footnotesize 7&\footnotesize 8&\footnotesize 9 &\footnotesize 10 &\footnotesize 11
\\[1mm]\hline \rule{0mm}{4mm}
\small $c(n)$ &\footnotesize 2 &\footnotesize 3&
\footnotesize 7&\footnotesize \kern 2.9pt 21&\footnotesize 135&
\footnotesize 2,470& \footnotesize 175,428 &\footnotesize 52,980,624 &\footnotesize ?
\\[1mm]\hline \rule{0mm}{4mm}
\small $v(n)$ &\footnotesize 2 &\footnotesize 3&
\footnotesize 7&\footnotesize \kern 2.9pt 21&\footnotesize 135&
\footnotesize 2,470& \footnotesize 319,124 & \footnotesize 1,214,554,343 &
\footnotesize $\sim 1.7\cdot 10^{15}$
\end{tabular}
\end{minipage}
\sk{3}\noindent
(more precisely: $v(11)=1,\!706,\!241,214,185,942$, computed by Minfeng Wang: see \cite{HRweb}).
According to the {\em On-Line Encyclopedia of Integer Sequences (OEIS)}, these numbers occur in other sequences:
\begin{ccote}\label{OLEIS2} \rm
The numbers $c(n)$ of chambers up to permutation coincide with the
{\em Numbers of self-dual equivalence classes of threshold functions of $n$ or fewer variables}, or the
{\em numbers of majority (i.e., decisive and weighted) games with $n$ players}, listed in \cite{OLEISsdth}.
\end{ccote}
\begin{ccote}\label{OLEIS1} \rm
The numbers $v(n)$ of virtual genetic codes coincide with the
{\em numbers of Boolean functions of $n$ variables that are self-dual and regular},
listed for $n\leq 10$ in \cite{OLEISsdr}.
\end{ccote}
The aim of this note is to explain these numerical coincidences by constructing natural bijections between
the sets under consideration. In particular, the above mentioned precise value $v(11)$
may be added in \cite{OLEISsdth}. The principal results are Propositions~\ref{Pa}, \ref{P.games} and~\ref{P2}
The paper is organized as follows. \secref{S.caltn} presents the transformation group used in various equivalence
relations. \secref{S.polyg} recalls the notations and the classification's result for polygon spaces.
In Sections~\ref{S.sptri} and~\ref{S.games}, we introduce threshold functions and majority games and prove
the bijections involved in~\ref{OLEIS2}, while \secref{S.sdf} concerns the case of~\ref{OLEIS1}.
Finally, we treat in \secref{S.ngen} the case of non-generic polygon spaces, giving rise to an apparently unknown
integer sequence.
\sk{1}
I thank Matthias Franz for drawing my attention to this problem and for useful conversations.
\section{The transformation group ${\mathcal T}_n$}\label{S.caltn}
In this section, we define the transformation group ${\mathcal T}_n$, responsible for several equivalence relations
occurring in this paper. Incidentally, a few notation are introduced, which are used throughout the next sections
Fix a positive integer $n$.
If $X$ is a set, the symmetric group $\sym{n}$ acts on $X^n$ by permuting the components.
This is a right action:
an element $x\in X^n$ is formally a map $x:\{1,\dots,n\}\to X$ ($x_i=x(i)$)
and $\sigma\in\sym{n}$ acts by pre-composition, i.e.
$x^\sigma=x\kern .7pt {\scriptstyle \circ} \kern 1pt\sigma$. Note that right actions are most often denoted exponentially in this paper.
Let ${\mathcal A}_n=({\mathbb Z}_2)^n$, the elementary abelian group of rank $n$ denoted additively.
The $\sym{n}$-action on ${\mathcal A}_n$ gives rise to the semi-direct product
\begin{equation}\label{EdefTn}
{\mathcal T}_n = {\mathcal A}_n \rtimes \sym{n} \, .
\end{equation}
Recall that, as a set, ${\mathcal T}_n$ coincides with ${\mathcal A}_n\times\sym{n}$.
We we may use the short notations $\nu=(\nu,{\rm id})$
and $\sigma=(0,\sigma)$ (which enables us to consider $\sym{n}$ as a subgroup of ${\mathcal T}_n$).
The group ${\mathcal T}_n$ is thus generated by $\nu\in{\mathcal A}_n$ and $\sigma\in\sym{n}$, subject
to the relations $\sigma^{-1}\nu\sigma=\nu^\sigma$. Note the formulae
$(\nu,\sigma)(\mu,\tau)=(\nu\mu^{\sigma^{-1}},\sigma\tau)$ and $(\nu,\sigma)^{-1}=(\nu^\sigma,\sigma^{-1})$.
The group ${\mathcal T}_n$ will act on several sets. We finish this section with a few examples.
\begin{Example}\label{EactRn} The action of ${\mathcal T}_n$ on ${\mathbb R}^n$ \rm is
defined as follows: if $z=(z_1,\dots,z_n)\in{\mathbb R}^n$,
the $i$-th component of $z^{(\nu,\sigma)}$ is
\begin{equation}\label{EDacrRn}
(z^{(\nu,\sigma)})_i = (-1)^{\nu_{\sigma(i)}} z_{\sigma(i)} \, .
\end{equation}
In particular, $z^\nu=\big((-1)^{\nu_1}z_1,\dots,(-1)^{\nu_n}z_n\big)$ and $z^\sigma=(z_{\sigma(1)},\dots,z_{\sigma(n)})$.
The following lemma will be useful.
\end{Example}
\begin{Lemma}\label{Ld1}
The inclusion $({\mathbb R}_{\geq 0})^n\hookrightarrow {\mathbb R}^n$ induces a bijection on the orbit sets
$$
({\mathbb R}_{\geq 0})^n\big/ \sym{n} \stackrel{\approx}{\longrightarrow} {\mathbb R}^n\big/{\mathcal T}_n \, .
$$
\end{Lemma}
\begin{proof}
Suppose that $z'=z^{(\nu,\sigma)}$.
If $z_i\geq 0$ and $z_i'\geq 0$, Formula~\eqref{EDacrRn} implies that $\nu_i=1$ only if $z'_i=0=(-1)^{\nu_i}z_{\sigma(i)}$,
in which case $\nu_i$ may be replaced by $0$ without changing $z'$.
Hence, $z'=z^{(\nu,\sigma)}=z^\sigma$, which implies that our map is injective.
By Formula~\eqref{EDacrRn} again, each ${\mathcal T}_n$-orbit contains an element $a$ with $a_i\geq 0$, so the map is also surjective.
\end{proof}
\begin{Example}\label{EactAn} The action of ${\mathcal T}_n$ on Boolean\ vectors. \ \rm
We consider another copy of ${\mathbb Z}_2^n$ called ${\mathcal B}_n$, the set of $n$-tuples $(x_1,\dots,x_n)$ of Boolean\ variables.
The set ${\mathcal B}_1$ is thus $\{true,false\}$, with its usual numerisation $true=1$, $false=0$, ${\rm xor=+}$, {\it etc}.
We sometimes use binary strings, e.g. $1010$ for $(1,0,1,0)$.
The addition law of ${\mathbb Z}_2^n$ produces a right action ${\mathcal B}_n\times{\mathcal A}_n\to{\mathcal B}_n$ of ${\mathcal A}_n$ on ${\mathcal B}_n$.
Note that the action of $1$ on $x_i$ is $(x_i+1)_{{\rm mod}\, 2} = \bar x_i$, the \dfn{negation} of $x_i$ ($\bar 0=1$ and $\bar 1=0$).
This is the reason for which an element of ${\mathcal A}_n$ is, in this paper, denoted by $\nu=(\nu_1,\dots,\nu_n)$,
the letter $\nu$ standing for {\it negation}.
Another useful equality is $\bar x_i=1-x_i$ (viewing $\{0,1\}\subset{\mathbb R}$).
This action extends to an action of ${\mathcal T}_n$ on ${\mathcal B}_n$ by the formula
$$
\big(x^{(\nu,\sigma)}\big)_i = x_{\sigma(i)}+ \nu_{\sigma(i)} \, .
$$
\end{Example}
\begin{Example}\label{EactPn}
The action of ${\mathcal T}_n$ on ${\mathcal P}(\underline{n})$, \rm where ${\mathcal P}(\underline{n})$ is the set of
subsets of~$\underline{n}$. We use the bijection $\chi\:{\mathcal P}(\underline{n})\to{\mathcal B}_n$
associating to $J\subset\underline{n}$ its \dfn{characteristic} $n$-tuple $\chi(J)$, whose $i$-th component is
$$
\chi(J)_i = \chi(J)(i) = {\rm truth}(i\in J)
$$
(i.e. $\chi(J)(i)=1$ if and only if $i\in J$). Note that $\chi(J)+\chi(K)=\chi(J{\scriptstyle\bigtriangleup} K)$,
where $\scriptstyle{\bigtriangleup}$ denotes the symmetric difference.
The ${\mathcal T}_n$-action on ${\mathcal P}(\underline{n})$ is defined so that
$\chi$ is equivariant, using the ${\mathcal T}_n$-action of \exref{EactAn}: $\chi(J^{(\nu,\sigma)}) = \chi(J)^{(\nu,\sigma)}$.
This amounts to the formulae $J^\nu=J{\scriptstyle\bigtriangleup}\chi^{-1}(\nu)$, $J^\sigma=\sigma^{-1}(J)$ and thus
$$
J^{(\nu,\sigma)} = \sigma^{-1}(J{\scriptstyle\bigtriangleup}\chi^{-1}(\nu)) \, .
$$
\end{Example}
\section{Polygon spaces}\label{S.polyg}
\setcounter{equation}{0}
In this section, we recall the notations for polygon spaces and their classification (see \cite{HR} or \cite[\S~10.3]{Ha}).
Fix two integers $n$ and $d$ and set $\underline{n}=\{1,2\dots,n\}$.
For $a=(a_1,\dots,a_n)\in{\mathbb R}^n$, the \dfn{polygon space} $\nua{n}{d}(a)$ is defined by
\begin{equation}\label{E.defpolsp}
\nua{n}{d}(a) = \Big\{z\in (S^{d-1})^n\,\big|\,\llangle{a}{z}=0\Big\} \bigg/SO(d) \, ,
\end{equation}
where $\llangle{\kern 1pt}{}$ denotes the standard scalar product in ${\mathbb R}^n$.
Classically, this definition is restricted to $a\in({\mathbb R}_{>0})^n$, in which case an
element of $\nua{n}{d}(a)$ may be visualized
as a configuration of $n$ successive segments in ${\mathbb R}^d$, of length
$a_1,\dots,a_{n}$, starting and ending at the origin.
The vector $a$ is thus called the \dfn{length vector}.
Following some recent works (see e.g. \cite{Fr}), we take advantage of
Definition~\eqref{E.defpolsp} making sense for $a\in{\mathbb R}^n$.
In most of the cases, this extension does not create
new polygon spaces up to homeomorphism (see \remref{R.newSp}).
The classification of polygon spaces up to homeomorphism is based on the stratification induced
by the \dfn{tie hyperplane arrangement} (or just \dfn{tie arrangement}) ${\mathcal H}({\mathbb R}^n)$ in ${\mathbb R}^n$
$$
{\mathcal H}({\mathbb R}^n)=\{{\mathcal H}_J\mid J\subset \underline{n}\} \, ,
$$
where the \dfn{$J$-tie hyperplane} ${\mathcal H}_J$ is defined by
$$
{\mathcal H}_J:=\Big\{(a_1,\dots,a_n)\in{\mathbb R}^n \Bigm|
\sum_{i\in J}a_i=\sum_{i\notin J}a_i\Big\}.
$$
A tie hyperplane is often called just a \dfn{wall}.
The stratification associated to ${\mathcal H}={\mathcal H}({\mathbb R}^n)$ is defined by the filtration
$$\{0\} = {\mathcal H}^{(0)}\subset
{\mathcal H}^{(1)}\subset\cdots\subset{\mathcal H}^{(n)}={\mathbb R}^n,
$$
with ${\mathcal H}^{(k)}$ being
the subset of those $a\in{\mathbb R}^n$ which belong to at least
$n-k$ distinct walls ${\mathcal H}_J$. A {\it stratum} of dimension
$k$ is a connected component of ${\mathcal H}^{(k)}-{\mathcal H}^{(k-1)}$.
Note that a stratum of dimension $k\geq 1$ is an open convex cone
in a $k$-plane of ${\mathbb R}^n$.
Strata of dimension $n$ are called {\it chambers} and their elements are called \dfn{generic}.
Note that the tie arrangement ${\mathcal H}({\mathbb R}^n)$ is invariant under the action of ${\mathcal T}_n$.
Indeed, using the tools of \exref{EactPn}, one checks
that $({\mathcal H}_J)^{(\nu,\sigma)}= {\mathcal H}_K$ with $K=\sigma\big(J{\scriptscriptstyle\bigtriangleup}\chi^{-1}(\nu)\big)$.
We may restrict the stratification ${\mathcal H}$ to $V=({\mathbb R}_{>0})^n$, $({\mathbb R}_{\neq 0})^n$, $({\mathbb R}_{\geq0})^n$ or ${\mathbb R}^n$.
Each of these choices for V gives rise to a set
$\mathbf{Ch}(V)$ of chambers, contained in the set $\mathbf{Str}(V)$ of corresponding strata.
\begin{Proposition}\label{PstratCH}
The inclusions $({\mathbb R}_{>0})^n\hookrightarrow ({\mathbb R}_{\geq0})^n \hookrightarrow{\mathbb R}^n$ descend to bijections
$$
\mathbf{Ch}(({\mathbb R}_{>0})^n)/\sym{n} \stackrel{\approx}{\longrightarrow}
\mathbf{Ch}(({\mathbb R}_{\geq0})^n)/\sym{n}
\stackrel{\approx}{\longrightarrow} \mathbf{Ch}({\mathbb R}^n)/{\mathcal T}_n
$$
and
$$
\mathbf{Str}(({\mathbb R}_{\geq0})^n)/\sym{n}
\stackrel{\approx}{\longrightarrow} \mathbf{Str}({\mathbb R}^n)/{\mathcal T}_n \, .
$$
\end{Proposition}
\begin{proof}
These bijections are direct consequences of \lemref{Ld1}, except for the first one
$\mathbf{Ch}(({\mathbb R}_{>0})^n)/\sym{n}\to\mathbf{Ch}(({\mathbb R}_{\geq0})^n)/\sym{n}$. The latter is obviously
injective. For the surjectivity, we use that if $\varepsilon>0$ is small enough and $a$ is generic, we can
replace the zero components of $a$ by $\varepsilon$ without leaving the chamber of $a$ (see e.g. \cite[\S\,2.1]{HauGDPS}).
\end{proof}
\begin{Remarks}\label{R.newstr}\rm
(a) The map $\mathbf{Str}(({\mathbb R}_{>0})^n)/\sym{n}\to \mathbf{Str}(({\mathbb R}_{\geq0})^n)/\sym{n}$ is clearly injective
but it is not surjective. The stratum $\{(0,\dots,0)\}$ is of course not in the image but also other less degenerate
strata, such as the intersection of the two walls ${\mathcal H}_{\{2\}}\cap{\mathcal H}_{\{3\}}$ in $\mathbf{Str}(({\mathbb R}_{\geq0})^3)$.
Indeed, the equations $a_2=a_1+a_3$ and $a_3=a_1+a_2$ imply that $a_1=0$. See also~\ref{comptStr}
(b) In \cite{HR} and in the lists of \cite{HRweb}, \dfn{conventional representatives} for classes in
$\mathbf{Ch}(({\mathbb R}_{>0})^n)/\sym{n}$ are used, allowing zero components in length-vectors.
These zeros stand there for any small enough positive numbers (e.g. $(0,1,1,1)$ means $(\varepsilon,1,1,1)$).
This actually uses the first bijection
of \proref{PstratCH} (see also its proof). Thus, in our new setting, these conventional representatives are bona fide
representative of $\mathbf{Ch}(({\mathbb R}_{\geq0})^n)/\sym{n}$.
However, using zero length is unsuitable for the symplectic geometry of spatial polygon spaces (see e.g. \cite{HK}).
\end{Remarks}
The main theorem for the classification of polygon spaces \cite[Theorem~1.1]{HR} generalizes,
with the same proof, in the following statement.
\begin{Theorem}\label{T.polcla}
Let $a,a'\in{\mathbb R}^n$. If $a$ and $a'$ are two representatives of the same class in $\mathbf{Str}({\mathbb R}^n)/{\mathcal T}_n$
then $\nua{n}{d}(a)$ and $\nua{n}{d}(a')$ are homeomorphic. \hfill \ensuremath{\Box}
\end{Theorem}
\begin{Remark}\rm
For generic $a$ and certain $n$ and $d$, the converse of \thref{T.polcla} is true: if
$\nua{n}{d}(a)$ and $\nua{n}{d}(a')$ are homeomorphic, then
$a$ and $a'$ represent the same class in $\mathbf{Ch}({\mathbb R}^n)/{\mathcal T}_n$
(see, e.g. \cite{FHS,Sc,Sc2}).
\end{Remark}
\begin{Remark}\label{R.newSp}\rm
By \proref{PstratCH} and \thref{T.polcla}, taking generic length vectors in ${\mathbb R}^n$
instead of $({\mathbb R}_{>0})^n$ does not produce new polygon spaces up to homeomorphism.
This is the same for non-generic length vectors, provided that they are taken in $({\mathbb R}_{\neq 0})^n$
(if $a\in({\mathbb R}_{\neq 0})^n$, then $b=a^\nu\in({\mathbb R}_{>0})^n$ for some $\nu\in{\mathcal A}_n$
and the map $x\mapsto x^\nu$ gives a homeomorphism from $\nua{n}{d}(b)$ to $\nua{n}{d}(a)$).
But the non-generic strata of
${\mathcal H}({\mathbb R}_{\geq 0})^n)$ (see \remref{R.newstr}.(a)) produce, in general, new polygon spaces.
For example, for $(0,1,1)\in{\mathcal H}_{\{2\}}\cap{\mathcal H}_{\{3\}}\in\mathbf{Str}(({\mathbb R}_{\geq0})^3)$, one has
$$
\begin{array}{rcll}
\nua{3}{3}(0,1,1) &=& \{(x,y,z)\in(S^2)^3\mid y+z=0\}/SO(3) \\[2mm] &\approx &
(S^2\times S^2)/SO(3) \\[2mm] &\approx &
pt \times S^2/SO(2) \approx [-1,1] \, ,
\end{array}
$$
which is not homeomorphic to a polygon space $\nua{3}{3}(a)$ for $a\in ({\mathbb R}_{>0})^3$.
Indeed, $\mathbf{Str}(({\mathbb R}_{>0})^3)/\sym{3}$ contains $3$ strata, giving
$\nua{3}{3}(0,0,1)=\emptyset$, $\nua{3}{3}(1,1,2)=pt$ and $\nua{3}{3}(1,1,1)=pt$.
\end{Remark}
We now restrict ourselves to to the generic case. If $a\in{\mathbb R}^n$ is generic then
$\sum_{i\in J}a_i\neq\sum_{i\notin J}a_i$ for all $J\subset\underline{n}$.
When $\sum_{i\in J}a_i<\sum_{i\notin J}a_i$,
the set $J$ is called \dfn{$a$-short} (or just \dfn{short}) and
its complement is \dfn{$a$-long} (or just \dfn{long}).
Short subsets form a subset ${\rm Sh}(a)$ of ${\mathcal P}(\underline{n})$.
Define ${\rm Sh}_n(a)=\{J\in\{1,\dots,n-1\}\mid J\cup\{n\} \in {\rm Sh}(a)\}$.
As $J$ is short if and only if $\bar J=\underline{n}-J$ is long,
the set ${\rm Sh}_n(a)$ determines ${\rm Sh}(a)$.
Indeed, either $n\in J$ or $n\in\bar J$, and thus
${\rm Sh}_n(a)$ tells us whether $J\in{\rm Sh}(a)$ or $\bar J\in{\rm Sh}(a)$.
The chamber of $a$ is obviously determined by ${\rm Sh}(a)$ and thus by ${\rm Sh}_n(a)$.
This permits us to characterize $\alpha\in\mathbf{Ch}(({\mathbb R}_{\geq0})^n)/{\mathcal T}_n$ by
a subset ${\rm gc}(\alpha)$ of ${\mathcal P}(\underline{n})$,
called the \dfn{genetic code} of $\alpha$.
For this, we consider, using \proref{PstratCH}, the only representative $a$ of $\alpha$ such that
$0\geq a_1 \geq\cdots\geq a_n$.
Define a partial order ``\ensuremath{\hookrightarrow}'' on
${\mathcal P}(\underline{m})$ by saying that $A\ensuremath{\hookrightarrow} B$ if and only if there exits
a non-decreasing map $\varphi:A\to B$ such that $\varphi(x)\geq x$.
Note that, if $B$ is short, so is $A$ if $A\ensuremath{\hookrightarrow} B$.
The {\it genetic code} ${\rm gc}(\alpha)$ of $\alpha$ is the set of
elements $A_1,\dots,A_k$ of ${\rm Sh}_n(a)$ which are
maximal with respect to the order ``\ensuremath{\hookrightarrow}''. The chamber of $a$ (and thus $\alpha$)
is determined by ${\rm gc}(\alpha)$.
We also use the notation ${\rm gc}(a)$ for ${\rm gc}(\alpha)$.
For instance ${\rm gc}((0,0,1))=\{\emptyset\}$, ${\rm gc}((0,1,1,1))= \{\{4,1\}\}$,
${\rm gc}((1,1,2,3,3,5))=\{\{6,2,1\},\{6,3\}\}$, etc (see \cite{HR}).
An algorithm is designed in \cite{HR} to list all the possible genetic code.
A first step is
to observe that all $A,B\in{\rm gc}(\alpha)$ satisfy
\begin{enumerate}\renewcommand{\labelenumi}{(\alph{enumi})}
\item $A\not\ensuremath{\hookrightarrow} B$ if $A\neq B$ and
\item $\bar A\not\ensuremath{\hookrightarrow} A$.
\end{enumerate}
Indeed, Condition (a) holds true by maximality of $A$ and,
if $\bar A\ensuremath{\hookrightarrow} A$, then $A$ would be both short and long which is impossible.
A finite set $\{A_1,\dots,A_k\}\subset {\mathcal P}_m(\underline{n})$,
satisfying Conditions (a) and (b) is called a {\it virtual genetic code} (of type~$n$).
\section{Self-dual threshold functions}\label{S.sptri}
\setcounter{equation}{0}
Fix a positive integer $n$.
We use the set ${\mathcal B}_n$ of Boolean\ vectors with its ${\mathcal T}_n$-action introduced in \exref{EactAn}.
A Boolean\ function on $n$ variables is a map ${\mathcal B}_n\to {\mathcal B}_1={\mathbb Z}_2$.
The group ${\mathcal T}_n$ acts on the right on the set of
Boolean\ functions by $f^{(\nu,\sigma)}(x)= f(x^{(\nu,\sigma)^{-1}})$, which gives the formula
\begin{equation}\label{E.aonfns}
f^{(\nu,\sigma)}(x_1,\dots,x_n) = (x_{\sigma^{-1}(1)}+\nu_1,\dots,x_{\sigma^{-1}(n)}+\nu_n) \, .
\end{equation}
This ${\mathcal T}_n$-action on the set of Boolean\ functions produces the equivalence relation used in \ref{OLEIS1}.
A Boolean\ function $f$ is called \dfn{self-dual} if $f(\bar x)=\overline{f(x)}$.
\begin{Lemma}\label{La}
Self-duality is preserved by the action of ${\mathcal T}_n$.
\end{Lemma}
\begin{proof}
Let $f\:{\mathcal B}_n\to{\mathcal B}_1$ be a Boolean\ function which is self-dual and let $(\nu,\sigma)\in{\mathcal T}_n$.
Note that $\bar x = x^{({\mathbf 1},{\rm id})}$ where ${\mathbf 1}=(1,\dots,1)$. Then,
$$
\begin{array}{rcll}
f^{(\nu,\sigma)}(\bar x) &=& f^{(\nu,\sigma)}(x^{({\mathbf 1},{\rm id})}) \\[2mm] &=&
f(x^{({\mathbf 1},{\rm id})(\nu,\sigma)^{-1}}) \\[2mm] &=&
f(x^{(\nu,\sigma)^{-1}({\mathbf 1},{\rm id})}) & \comeq{since $({\mathbf 1},{\rm id})$ is in the center of ${\mathcal T}_n$}\\[2mm] &=&
f\big(\overline{x^{(\nu,\sigma)^{-1}}}\big) \\[2mm] &=&
\overline{f(x^{(\nu,\sigma)^{-1}})} & \comeq{since $f$ is self-dual} \\[2mm] &=&
\overline{f^{(\nu,\sigma)}(x)} \, ,
\end{array}
$$
which proves that $f^{(\nu,\sigma)}$ is self-dual.
\end{proof}
Let $w=(w_1,\dots,w_n)\in{\mathbb R}^n$ and $t\in {\mathbb R}$. We consider the Boolean\ function
$$
f_{(w,t)}(x) = {\rm truth}\big(\llangle{x}{w} \geq t \big) =
\left\{
\begin{array}{lll}
1 & \hbox{if } \llangle{x}{w} \geq t \\
0 & \hbox{otherwise,}
\end{array}\right.
$$
where $\llangle{\kern 1pt}{}$ denotes the standard scalar product in ${\mathbb R}^n$. The function $f_{(w,t)}$
is called the \dfn{threshold function} with \dfn{weights} $w_1,w_2,\dots,w_n$ and \dfn{threshold} $t$
(see \cite{Hn} and \remref{rkA} below). Reference \cite{Hn} emphasize the importance of threshold functions
in neuron-like systems. The few subsequent lemmas gather several properties of threshold functions.
\begin{Lemma}\label{Lb}
Threshold functions are preserved by the action of ${\mathcal T}_n$. More precisely,
let $(w,t)\in{\mathbb R}^n\times{\mathbb R}$ and let $(\nu,\sigma)\in{\mathcal T}_n$.
Then $f_{(w,t)}^{(\nu,\sigma)}=f_{(w',t')}$ where
$$
w'=w^{(\nu,\sigma)} \ \hbox{ and } \ t'=t - \llangle{\nu}{w^{\sigma}} \, .
$$
\end{Lemma}
\begin{proof}
As we are dealing with an action of ${\mathcal T}_n$, it is enough to prove the lemma for $\sigma=(0,\sigma)$ and
for $\nu=(\nu,{\rm id})$.
In the first case, one has
$$
f_{(w,t)}^\sigma(x) = {\rm truth}\big(\llangle{x^{\sigma^{-1}}}{w} \geq t \big)
= {\rm truth}\big(\llangle{x}{w^\sigma} \geq t \big) = f_{(w^\sigma,t)} \, .
$$
In the other case, one first prove, using the truth tables of $x_i$ and $\nu_i$, that
$$
(x_i)^{\nu_i} w_i = x_i (-1)^{\nu_i}w_i + \nu_iw_i \, .
$$
This implies that
$\llangle{x^\nu}{w}=\llangle{x}{w^\nu}+\llangle{\nu}{w}$.
Therefore,
$$
\begin{array}{rcll}
f_{(w,t)}^\nu(x) &=& f_{(w,t)}(x^\nu) \\[2mm] &=&
{\rm truth}\big(\llangle{x^{\nu}}{w}\geq t \big) \\[2mm] &=&
{\rm truth}\big(\llangle{x}{w^\nu}\geq t - \llangle{\nu}{w}\big) \\[2mm] &=&
f_{(w',t')}(x) \, .
\end{array}
$$
\end{proof}
For $w\in{\mathbb R}^n$, define $\peri{w}= \frac{1}{2}\sum_{i=1}^n w_i$.
The relationship between threshold and self-dual functions is the following.
\begin{Lemma}\label{Lsd}
A threshold function $f_{(w,t)}$ is self-dual if and only if the following two conditions hold.
\begin{itemize}
\item[(a)] $f_{(w,t)}=f_{(w,\peri{w})}$ and
\item[(b)] $\llangle{x}{w}\neq \peri{w}$ for any $x\in{\mathcal B}_n$.
\end{itemize}
\end{Lemma}
\begin{proof}
Conditions (a) and (b) are clearly sufficient for $f_{(w,t)}$ being self-dual. Conversely, if $f=f_{(w,t)}$ is self-dual, then
$$
\llangle{x}{w} \geq t \ \Leftrightarrow \
f(x)=1 \ \Leftrightarrow \ f(\bar x)=0 \ \Leftrightarrow \ \llangle{\bar x}{w} < t \, .
$$
This, together with the same argument exchanging $x$ and $\bar x$, proves that
$$
\llangle{x}{w} < \llangle{\bar x}{w} \ \hbox{ or } \ \llangle{x}{w} > \llangle{\bar x}{w}
$$
for all $x\in{\mathcal B}_n$. As $\llangle{x}{w} + \llangle{\bar x}{w} = 2\peri{w}$, this proves (a) and (b).
\end{proof}
\begin{Lemma}\label{Lperi}
Let $w\in{\mathbb R}^n$ and $(\nu,\sigma)\in{\mathcal T}_n$. Then $f_{(w,\peri{w})}^{(\nu,\sigma)} = f_{(w',\peri{w'})}$
(where $w'$ is given by \lemref{Lb}).
\end{Lemma}
\begin{proof}
As we are dealing with an action of ${\mathcal T}_n$, it is enough to prove the lemma for $(0,\sigma)$ and $(\nu,{\rm id})$.
The case $(0,\sigma)$ is straightforward by \lemref{Lb} and the equality $\peri{w^{\sigma^{-1}}}=\peri{w}$.
For $(\nu,{\rm id})$, \lemref{Lb} again tells us that $f_{(w,\peri{w})}^{(\nu,{\rm id})} = f_{(w',t')}$ with
$w'=w^{\nu}$ and
$$
t' = \peri{w} -\llangle{\nu}{w} = \frac{1}{2}\big(\sum w^i -2\llangle{\nu}{w}\big) = \peri{w^\nu} \, . \qedhere
$$
\end{proof}
We are ready to prove the main result of this section.
To a generic $a\in{\mathbb R}^n$, we associate the threshold function $f_{(a,\peri{a})}$,
which is self dual by \lemref{Lsd}. If $a$ and $a'$ belong to the same chamber of ${\mathcal H}({\mathbb R}^n)$,
then
$$
f_{(a,\peri{a})}^{-1}(\{0\}) = {\rm Sh}(a) = {\rm Sh}(a') = f_{(a',\peri{a'})}^{-1}(\{0\}) \, ,
$$
which proves that $f_{(a,\peri{a})}=f_{(a',\peri{a'})}$. We thus get a map
$$
\tilde\Xi:\mathbf{Ch}({\mathbb R}^n) \to \mathbf{SDT}(n) \, ,
$$
where $\mathbf{SDT}(n)$ denotes the set of self-dual threshold functions on ${\mathcal B}_n$.
\begin{Proposition}\label{Pa}
The above map $\tilde\Xi$ descends to a bijection
$$
\Xi:\mathbf{Ch}({\mathbb R}^n)/{\mathcal T}_n \stackrel{\approx}{\longrightarrow} \mathbf{SDT}(n)/{\mathcal T}_n \, .
$$
\end{Proposition}
\begin{proof}
By \lemref{Lperi}, the map $\tilde\Xi$ is ${\mathcal T}_n$-equivariant, so the orbit map $\Xi$ is well defined.
It is surjective by \lemref{Lsd}.
For the injectivity, let $\alpha$ and $\beta$ be two chambers, represented by length vectors $a$ and $b$.
If $\Xi(\alpha)=\Xi(\beta)$, then $f_{(b,\peri{b})}=f_{(a,\peri{a})}^{(\nu,\sigma)}$
for some $(\nu,\sigma)\in{\mathcal T}_n$. By \lemref{Lperi}, one has
$f_{(b,\peri{b})}=f_{(a^{(\nu,\sigma)},\peri{a^{(\nu,\sigma)}})}$. Therefore
$$
{\rm Sh}\big(a^{(\nu,\sigma)}\big) =
f_{(a^{(\nu,\sigma)},\peri{a^{(\nu,\sigma)}})}^{-1}(\{0\}) =
f_{(b,\peri{b})}^{-1}(\{0\}) = {\rm Sh}(b) \, ,
$$
which implies that $\beta=\alpha^{(\nu,\sigma)}$.
\end{proof}
\begin{Remarks}\label{rkA}\rm
(a) There are several variants in the literature for the definition of a threshold function, for instance
by requiring that $w_i$ and/or $t$ be integers (see e.g. \cite[p.~75]{Kn}).
As chambers contain integral representative, \proref{Pa} holds true as well for these versions.
(b) Variables corresponding to zero weights are idle for $f_{(a,\peri{a})}$, so the latter
depends on fewer than $n$ variables. This is the reason of the words ``$n$ or fewer variables'' in \ref{OLEIS2}.
(c) Composing the bijection $\Xi$ of \proref{Pa} with the bijection\\
$\mathbf{Ch}(({\mathbb R}_{>0})^n)/\sym{n} \stackrel{\approx}{\longrightarrow} \mathbf{Ch}({\mathbb R}^n)/{\mathcal T}_n$ of \proref{PstratCH},
produces a bijection
\begin{equation}\label{E.bij1}
\mathbf{Ch}(({\mathbb R}_{>0})^n)/\sym{n} \stackrel{\approx}{\longrightarrow} \mathbf{SDT}(n)/{\mathcal T}_n \, .
\end{equation}
This explains why the numbers of the first line of the table in the introduction are equal to those of~\cite{OLEISsdth}.
\end{Remarks}
By \proref{PstratCH}, the bijection of \eqref{E.bij1} factors through the bijection\\
$\mathbf{Ch}(({\mathbb R}_{>0})^n)/\sym{n} \stackrel{\approx}{\longrightarrow} \mathbf{SDT}(n)/{\mathcal T}_n$.
A direct consequence is the following lemma, which will be useful later.
\begin{Lemma}\label{Ld}
Let $(w,t)$ and $(w',t')$ be two elements of ${\mathbb R}^n\times{\mathbb R}$. Suppose that $w$ and $w'$ are generic and that
$f_{(w,t)}$ and $f_{(w',t')}$ are in the same ${\mathcal T}_n$-orbit.
If $w_i\geq 0$ and $w_i'\geq 0$ for all $i\in\underline{n}$, then $w$ and $w'$ are in the same $\sym{n}$-orbit.
\end{Lemma}
\section{Decisive weighted majority games}\label{S.games}
\setcounter{equation}{0}
In this section, we describe the equivalence between $\mathbf{Ch}(({\mathbb R}_{\geq 0})^n)/\sym{n}$ and
the strategic equivalence classes of
decisive weighted majority games with $n$ players, as mentioned in \ref{OLEIS2}. Our references for game theory are
\cite[Chapter~10]{vNM} and \cite{Is}.
A \dfn{game} on a set of $n$ players (indexed by $\underline n$) is
a set of subsets of $\underline{n}$ called the \dfn{winning sets}, such that any set containing a winning
set is also winning. The game is \dfn{decisive} (or \dfn{simple}) if
for any $S\subset \underline{n}$, either $S$ or its complement is winning but not both.
Two games are \dfn{strategically equivalent} if there is a bijection
between their players identifying the families of winning sets.
A game with $n$ players may be extended to $m>n$ players by adding $m-n$ ``voteless'' players or \dfn{dummies}:
a subset $S$ of $\underline{m}$ is thus wining if and only if $S\cap\underline{n}$ is wining.
We see that a game ${\mathcal G}$ defines and is determined by a Boolean\ function $f_{{\mathcal G}}\:{\mathcal B}_n\to{\mathcal B}_1$,
given by $f(x)=1$ if and only if $x$ is the characteristic $n$-tuple of a set $S\subset\underline{n}$
which is a winning set.
As any superset of a winning wins, the function $f_{\mathcal G}$ is \dfn{monotone}, i.e. its value does not change from~$1$
to~$0$ when any of its variables changes from~$0$ to~$1$ \cite[p.~55]{Kn}.
That ${\mathcal G}$ is decisive translates into $f_{\mathcal G}$ being self-dual.
Two games ${\mathcal G}$ and ${\mathcal G}'$ are
strategically equivalent if and only if $f_{{\mathcal G}}$ and $f_{{\mathcal G}'}$ are in the same $\sym{n}$-orbit.
A game ${\mathcal G}$ is a \dfn{weighted majority game} if there exists $w\in{\mathbb R}^n$ such that $f_{{\mathcal G}}=f_{(w,\peri{w})}$.
We write ${\mathcal G}={\mathcal G}(w)$. If $w_i\leq 0$, then the player $i$ is a dummy.
Indeed, since every superset of a winning
set wins, $w_i\leq 0$ implies that $|w_i|$ is so small that it makes no difference.
Therefore, one can replace the negative weights by $0$ without changing the strategic equivalence class of the game.
We can thus suppose that no weight is negative.
By \lemref{Ld}, two decisive majority games ${\mathcal G}(w)$ and ${\mathcal G}(w')$ are then strategically equivalent if and only if
$w$ and $w'$ are in the same $\sym{n}$-orbit.
Note that if $a\in({\mathbb R}_{\geq0})^n$ is a length vector, the winning set of ${\mathcal G}_a$ are the $a$-long subsets of $\underline{n}$.
Also, $a$ is generic if and only if ${\mathcal G}_a$ is decisive.
The above considerations, together with Propositions~\ref{PstratCH} and~\ref{Pa}, gives the following result.
\begin{Proposition}\label{P.games}
The map $a\mapsto{\mathcal G}(a)$ induces a bijection from $\mathbf{Ch}(({\mathbb R}_{\geq 0}^n))/\sym{n}$ to the set of strategic equivalence classes
of decisive weighted majority games.
\end{Proposition}
\section{Self-dual regular Boolean\ functions}\label{S.sdf}
\setcounter{equation}{0}
A partial order on ${\mathcal B}_n$ is defined by saying that $x\preceq y$ if
$x_1+\cdots x_k \leq y_1+\cdots y_k$ for $1\leq k\leq n$ \cite[p.~92]{Kn}.
Note that $x\preceq y$ if and only if $\bar y\preceq\bar x$: indeed, $x_1+\cdots+x_k \leq y_1+\cdots+y_k$
if and only if $k-x_1-\cdots -x_k \geq k-y_1-\cdots -y_k$.
A Boolean\ $f\:{\mathcal B}_n\to{\mathcal B}_1$ is \dfn{regular} if $f(x)\preceq f(y)$ whenever $x\preceq y$ \cite[p.~93]{Kn}.
For example, a threshold function $f_{(w,t)}$ is regular if $w_1\geq w_2 \geq\cdots\geq w_n\geq 0$.
Let $\mathbf{SDR}(n)$ be the set of self dual regular Boolean\ functions on ${\mathcal B}_n$.
\begin{Proposition}\label{P2}
There is a bijection between $\mathbf{SDR}(n)$ and the set of virtual genetic codes (see \secref{S.polyg}).
\end{Proposition}
Before proving \proref{P2}, we note the following lemma, in which ${\mathcal B}_n^1=\{x\in{\mathcal B}_n\mid x_1=1\}$.
\begin{Lemma}\label{Lsdbun}
A self-dual Boolean\ $f\:{\mathcal B}_n\to{\mathcal B}_1$ is determined by its restriction to ${\mathcal B}_n^1$.
\end{Lemma}
\begin{proof}
We use that $x\in{\mathcal B}_n^1$ if and only if $\bar x\notin{\mathcal B}_n^1$. As $f$ is self-dual, one has
\begin{equation}\label{Lsdbun-eq}
f(x) =
\left\{
\begin{array}{lll}
f(x) & \hbox{if $x\in{\mathcal B}_n^1$} \\[1mm]
\overline{f(\bar x)} & \hbox{otherwise.}
\end{array}\right.
\end{equation}
\end{proof}
\begin{proof}[Proof of \proref{P2}]
To $f\in\mathbf{SDR}(n)$, we associate its \dfn{code} $\gamma(f)$ which is a subset of ${\mathcal B}_n^1$.
By definition, $\gamma(f)$ is the set of $\preceq$-maximal elements $x\in{\mathcal B}_n^1$ for which $f(x)=0$.
For instance $\gamma(f_{((2,1,1,1),5/2)})=\{1000\}$ and $\gamma(f_{((2,2,2,1),7/2)})=\{1001\}$.
If $\gamma(f)=\{b_1,\dots,b_k\}$, then
\begin{itemize}
\item[(i)] $b_i\not\preceq b_j$ for all $i\neq j$ and
\item[(ii)] $\bar b_i\not\preceq b_j$ for all $i,j$.
\end{itemize}
Let $\Gamma_n$ be the set of subsets of ${\mathcal B}_n^1$ satisfying (i) and (ii).
We first establish that $\gamma\:\mathbf{SDR}(n)\to\Gamma_n$ is a bijection.
Indeed, $\gamma(f)$ clearly determines the restriction of $f$ to ${\mathcal B}_n^1$, and then determines $f$ by \lemref{Lsdbun};
this proves that $\gamma$ is injective. For the surjectivity, let $R\in\Gamma_n$. The formula
$$
f(x) =
\left\{
\begin{array}{lll}
0 & \hbox{if $\exists\ r\in R$ with $x\preceq r$} \\[1mm]
1 & \hbox{otherwise}
\end{array}\right.
$$
defines a function on ${\mathcal B}_n^1$ which can be extended to $f\:{\mathcal B}_n\to{\mathcal B}_1$ by \eqref{Lsdbun-eq}.
Such a definition guarantees that $f$ is self-dual. For the regularity, let $x\preceq y$ be two elements in ${\mathcal B}_n$.
The condition $f(x)\preceq f(y)$ is automatic if $f(y)=1$. We can thus assume that $f(y)=0$, so we must prove that $f(x)=0$.
There are four cases.
{\it Case 1: $x_1=1=y_1$.} \ As $f(y)=0$, there exists $r\in R$ with $y\preceq r$. As $x\preceq y$, then $x\preceq r$ and thus $f(x)=0$.
{\it Case 2: $x_1<y_1$.} \ If $f(x)=1$, then $\bar x \preceq r$ for some $r\in R$.
As $f(y)=0$, there exists $s\in R$ with $y\preceq s$. Therefore, $\bar r \preceq x \preceq y \preceq s$ which contradicts (ii).
{\it Case 3: $x_1> y_1$}. This is impossible since $x\preceq y$.
{\it Case 4: $x_1=0=y_1$}. If $f(x)=1$, then $f(\bar x)=0$. As $f(\bar y)=1$ and $\bar y\preceq \bar x$, the pair $(\bar y,\bar x)$
would contradict Case~1, already established.
\sk{1}
It remains to establish a bijection from $\Gamma_n$ and the set of virtual genetic codes.
To $x\in{\mathcal B}_n^1$ one associates $x^\sharp\subset \underline{n}$ by the rule
$$
x^\sharp = \{i\in \underline{n} \mid x_{n+1-i} = 1 \} \, .
$$
For instance, $(1000)^\sharp = \{4\}$ while $(1010)^\sharp = \{2,4\}$. Obviously, $x\mapsto x^\sharp$
is a bijection between ${\mathcal B}_n^1$ and the subsets of $\underline{n}$ containing $n$. Conditions (i) and (ii) above
are intertwined with Conditions (a) and (b) of \cite[p.~37]{HR}. The latter define a virtual genetic code.
Hence, the correspondence $x\mapsto x^\sharp$ maps $\Gamma_n$ bijectively to the set of virtual genetic codes.
\end{proof}
\begin{Remark}\rm
For $n\geq 9$, not every self-dual regular Boolean\ function is equivalent to a threshold function. As an example,
the function with $\gamma(f)=\{100101010\}$, corresponding to the genetic code $\{9,6,4,2\}$ (see \cite[Lemma 4.5]{HR}).
\end{Remark}
\section{Non-generic strata}\label{S.ngen}
\setcounter{equation}{0}
In this section, we give an analogue of the bijection $\Xi$ of \proref{Pa}, extended to possibly non-generic strata,
taking advantage of 3-valued (3V) Boolean\ functions.
A \dfn{3V-Boolean\ function} is a map $f\:{\mathcal B}_n\to \{-1,0,1\}$. It is \dfn{self-dual} if $f(\bar x) = -f(x)$.
As in \secref{S.sptri}, a right ${\mathcal T}_n$-action on the set of 3V-Boolean\ functions using \eqref{E.aonfns}.
As in \lemref{La}, one proves that self-duality is preserved by this ${\mathcal T}_n$-action.
Let $w=(w_1,\dots,w_n)\in{\mathbb R}^n$ and $t\in {\mathbb R}$. The 3V-Boolean\ function
$$
f_{(w,t)}^{3V}(x) =
\left\{
\begin{array}{rll}
1 & \hbox{if } \llangle{x}{w} > t \\
0 & \hbox{if } \llangle{x}{w} = t \\
-1 & \hbox{if } \llangle{x}{w} < t
\end{array}\right.
$$
is called the \dfn{3V-threshold function} with \dfn{weights} $w_1,w_2,\dots,w_n$ and \dfn{threshold} $t$.
With essentially the same proofs, Lemma~\ref{La}--\ref{Ld} remain valid without change for 3V-threshold functions,
except for \lemref{Lsd} which requires the hypothesis $0\notin {Image}(f_{(w,t)}^{3V})$.
If this is not the case, one easily proves the following lemma.
\begin{Lemma}\label{lsdii}
Let $(w,t)\in{\mathbb R}^n\times{\mathbb R}$ such that $0\in {Image}(f_{(w,t)}^{3V})$. Then, $f_{(w,t)}^{3V}$ is self-dual
if and only if $t=\peri{w}$. \hfill \ensuremath{\Box}
\end{Lemma}
Let $\mathbf{SDT}^{3V}(n)$ be the set of self-dual 3V-threshold functions on ${\mathcal B}_n$.
As in \secref{S.sptri}, we define a map $\tilde\Xi^{3V}\:\mathbf{Str}({\mathbb R}^n)\to \mathbf{SDT}^{3V}(n)$
by associating associating to $S\in \mathbf{Str}({\mathbb R}^n)$ the 3V-threshold function
$f_{(a,\peri{a})}^{3V}$ for $a\in S$. We check that $\tilde\Xi^{3V}$ is well defined and ${\mathcal T}_n$-equivariant,
thus inducing a map $\tilde\Xi^{3V}\:\mathbf{Str}({\mathbb R}^n)/{\mathcal T}_n\to \mathbf{SDT}^{3V}(n)/{\mathcal T}_n$.
The same proof as for \proref{Pa} gives following proposition.
\begin{Proposition}\label{Pb}
The map $\Xi^{3V}:\mathbf{Str}({\mathbb R}^n)/{\mathcal T}_n \to \mathbf{SDT}^{3V}(n)/{\mathcal T}_n$ is a bijection.
\hfill \ensuremath{\Box}
\end{Proposition}
\begin{Remark}\rm The relationship between the maps $\Xi$ and $\Xi^{3V}$ of Propositions~\ref{Pa} and \ref{Pb}
is as follows.
There is an obvious injection $j\:\mathbf{SDT}(n)/{\mathcal T}_n\to \mathbf{SDT}^{3V}(n/{\mathcal T}_n)$
induced by $f\mapsto \epsilon\kern .7pt {\scriptstyle \circ} \kern 1pt f$ where $\epsilon(u)=(-1)^u$. Its image is the set of 3V-Boolean\ functions $f$ such that
$0\notin {\rm Image}(f)$. One has a commutative diagram
$$
\begin{array}{c}{\xymatrix@C-3pt@M+2pt@R-4pt{%
\mathbf{Ch}({\mathbb R}^n)/{\mathcal T}_n\kern 2pt \ar@{>->}[r] \ar[d]_{\Xi}^{\approx} &
\mathbf{Str}({\mathbb R}^n)//{\mathcal T}_n \ar[d]^{\Xi^{3V}}_{\approx} \\
\mathbf{SDT}(n)/{\mathcal T}_n\kern 2pt \ar@{>->}[r]^(.47){j} & \mathbf{SDT}^{3V}(n)/{\mathcal T}_n
}}\end{array} .
$$
\end{Remark}
\begin{ccote}\label{comptStr} Computing the number of strata. \rm \
Consider the following numbers
\begin{itemize}
\item $c(n) = \sharp\big(\mathbf{Ch}({\mathbb R}^n)/{\mathcal T}_n\big)$
\item $k(n)
=\sharp\big(\mathbf{Str}(({\mathbb R}_{\neq 0})^n)/{\mathcal T}_n\big)$
\item $tk(n) = \sharp\big(\mathbf{Str}({\mathbb R}^n)/{\mathcal T}_n \big)$.
\end{itemize}
For example, $c(1)=0$ and $k(1)=tk(1)=1$ (the stratum of $(0)$). For $n=2$, one has
$c(2)=1$ (the chamber of $(0,1)$), $k(2)=2$ (the previous chamber and the stratum of $(1,1)$), while
$tk(2)=3$ because the stratum of $(0)$ in $\mathbf{Str}({\mathbb R}^1)$ gives rise to that of $(0,0)$
in $\mathbf{Str}({\mathbb R}^2)$. In general, the injection
${\mathbb R}^{n-1}\approx\{0\}\times{\mathbb R}^{n-1}\hookrightarrow{\mathbb R}^n$ induces an injection
$[\mathbf{Str}({\mathbb R}^{n-1})-\mathbf{Ch}({\mathbb R}^{n-1})]/{\mathcal T}_{n-1}
\hookrightarrow \mathbf{Str}({\mathbb R}^{n})/{\mathcal T}_{n}$. This proves the recursion formula
\begin{equation}\label{E.rec}
tk(n) =
k(n) + tk(n-1)- c(n-1) \, .
\end{equation}
The number $k(n)$ was computed in \cite[\S\,5]{HR} for $n\leq 8$. Thanks to \proref{PstratCH},
the values of $c(n)$ may be taken from the table in the introduction. Using Formula~\eqref{E.rec},
we thus get the following table.
\begin{center}
\begin{minipage}{110mm}
\begin{tabular}{c|cccccccccccc} \small
$n$ & \footnotesize 1 & \footnotesize 2 &\footnotesize 3&\footnotesize 4&\footnotesize 5&\
\footnotesize 6 &\footnotesize 7&\footnotesize 8
\\[1mm]\hline \rule{0mm}{4mm}
\small $c(n)$ & \footnotesize 0 & \footnotesize 1 &\footnotesize 2 &\footnotesize 3&
\footnotesize 7&\footnotesize \kern 2.9pt 21&\footnotesize 135&
\footnotesize 2,470
\\[1mm]\hline \rule{0mm}{4mm}
\small $k(n)$ &\footnotesize 1 & \footnotesize 2
&\footnotesize 3 &\footnotesize 7&
\footnotesize 21&\footnotesize \kern 2.9pt 117&\footnotesize 1506&
\footnotesize 62254&
\\[1mm]\hline \rule{0mm}{4mm}
\small $tk(n)$ & \footnotesize 1 & \footnotesize 3
&\footnotesize 5 &\footnotesize 10 &
\footnotesize 28 &\footnotesize 138 &\footnotesize 1623 &
\footnotesize 63742
\end{tabular}
\end{minipage}
\end{center}
\end{ccote}
The sequences $k(n)$ and $tk(n)$ do not seem to occur in the {\em On-Line Encyclopedia of Integer Sequences}.
|
\section{Introduction}
As it is well known, geometry of space is associated with mathematical
group. The idea of invariance of geometry under transformation group may
imply that, on some spacetimes of maximum symmetry there should be a
principle of relativity which requires the invariance of physical laws
without gravity under transformations among inertial systems \cite{1}.
Besides, theory of curves and the curves of constant curvature in the
equiform differential geometry of the isotropic spaces $I_{3}^{1}$ , $%
I_{3}^{2}$ and the Galilean space $G_{3}$ are described in \cite{2} and \cite%
{3}, respectively. The pseudo-Galilean space is one of the real Cayley-Klein
spaces. It has projective signature $(0,0,+,-)$ according to \cite{2}. The
absolute of the pseudo-Galilean space is an ordered triple $\{w,f,I\}$ where
$w$ is the ideal plane, $f$ a line in $w$ and $I$ is the fixed hyperbolic
involution of the points of $f$. In \cite{4}, from the differential geometric point of
view, K. Arslan and A. West defined the notion of AW(k)-type submanifolds. Since
then, many works have been done related to AW(k)-type submanifolds (see, for
example, \cite{5,6,7,8,9,10}). In \cite{9}, \"{O}zg\"{u}r and Gezgin studied a Bertrand curve of AW$(k)$-type and furthermore, they showed that there is no such Bertrand curve of AW$(1)$ and AW$(3)$%
-types if and only if it is a right circular helix. In addition, they
studied weak AW$(2)$-type and AW$(3)$-type conical geodesic curves in
Euclidean 3-space $E^{3}$. Besides, In $3$-dimensional Galilean space and
Lorentz space, the curves of AW$(k)$-type were investigated in \cite{6,8}.
In \cite{7}, the authors gave curvature conditions and characterizations
related to AW$(k)$-type curves in $E^{n}$ and in \cite{10}, the authors
investigated curves of AW$(k)$-type in the $3$-dimensional null cone.
In this paper, to the best of author's knowledge, Bertrand curves of AW$(k)$%
-type have not been presented in the equiform geometry of the
pseudo-Galilean space $G_{3}^{1}$ in depth. Thus, the study is proposed to
serve such a need. Our paper is organized as follows. In Section $2$, the
basic notions and properties of a pseudo-Galilean geometry are reviewed. In
Section $3$, properties of the equiform geometry of the pseudo-Galilean
space $G_{3}^{1}$ are given. Section $4$ contains a study of AW$(k)$-type
equiform Frenet curves. Equiform Bertrand curves of AW$(k)$-type in $G_{3}^{1}$ included in
section $5$.
\section{Pseudo-Galilean geometric meanings}
In this section, let us first recall basic notions from pseudo-Galilean
geometry \cite{11,12}. In the inhomogeneous affine coordinates for points and vectors
(point pairs) the similarity group $H_{8}$ of $G_{3}^{1}$ has the following
form
\begin{align}
\bar{x}& =a+b.x, \notag \\
\bar{y}& =c+d.x+r.\cosh \theta .y+r.\sinh \theta .z, \notag \\
\bar{z}& =e+f.x+r.\sinh \theta .y+r.\cosh \theta .z,
\end{align}%
where $a,b,c,d,e,f,r$ and $\theta $ are real numbers.
Particularly, for $b=r=1,$ the group $(2.1)$ becomes the group $B_{6}\subset
H_{8}$ of isometries (proper motions) of the pseudo-Galilean space $G_{3}^{1}
$. The motion group leaves invariant the absolute figure and defines the
other invariants of this geometry. It has the following form
\begin{align}
\bar{x}& =a+x, \notag \\
\bar{y}& =c+d.x+\cosh \theta .y+\sinh \theta .z, \notag \\
\bar{z}& =e+f.x+\sinh \theta .y+\cosh \theta .z.
\end{align}%
According to the motion group in the pseudo-Galilean space, there are
non-isotropic vectors $A(A_{1},A_{2},A_{3})$ (for which holds $A_{1}\neq 0$)
and four types of isotropic vectors: spacelike ($A_{1}=0,$ $%
A_{2}^{2}-A_{3}^{2}>0$), timelike ($A_{1}=0,$ $A_{2}^{2}-A_{3}^{2}<0$) and
two types of lightlike vectors ($A_{1}=0,A_{2}=\pm A_{3}$). The scalar
product of two vectors $u=(u_{1},u_{2},u_{3})$ and $v=(v_{1},v_{2},v_{3})$
in $G_{3}^{1}$ is defined by
\begin{equation*}
\left\langle u,v\right\rangle =\left\{
\begin{array}{c}
u_{1}v_{1},\text{ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ if }%
u_{1}\neq 0\text{ or }v_{1}\neq 0, \\
u_{2}v_{2}-u_{3}v_{3}\text{ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ if \ }%
u_{1}=0\text{ and }v_{1}=0.%
\end{array}%
\right\}
\end{equation*}%
We introduce a pseudo-Galilean cross product in the following way
\begin{equation*}
u\times _{G_{3}^{1}}v=\left\vert
\begin{array}{ccc}
0 & -j & k \\
u_{1} & u_{2} & u_{3} \\
v_{1} & v_{2} & v_{3}%
\end{array}%
\right\vert ,
\end{equation*}%
where $j=(0,1,0)$ and $k=(0,0,1)$ are unit spacelike and timelike vectors,
respectively. Let us recall basic facts about curves in $G_{3}^{1}$, that
were introduced in \cite{15}.
A curve $\gamma (s)=(x(s),y(s),z(s))$ is called an admissible curve if it has no
inflection points $(\dot{\gamma}\times \ddot{\gamma}\neq 0)$ and no
isotropic tangents $(\dot{x}\neq 0)$ or normals whose projections on the
absolute plane would be lightlike vectors $(\dot{y}\neq \pm \dot{z})$. An
admissible curve in $G_{3}^{1}$ is an analogue of a regular curve in
Euclidean space \cite{12}.
For an admissible curve $\gamma :I\subseteq \mathbb{R}\rightarrow G_{3}^{1},$
the curvature $\kappa (s)$ and torsion $\tau (s)$ are defined by%
\begin{equation}
\kappa (s)=\frac{\sqrt{\left\vert \ddot{y}(s)^{2}-\ddot{z}(s)^{2}\right\vert
}}{(\dot{x}(s))^{2}},\text{ }\tau (s)=\frac{\ddot{y}(s)\dddot{z}(s)-\dddot{y}%
(s)\ddot{z}(s)}{\left\vert \dot{x}(s)\right\vert ^{5}\cdot \kappa ^{2}(s)},%
\text{\ }
\end{equation}%
expressed in components. Hence, for an admissible curve $\gamma :I\subseteq
\mathbb{R}\rightarrow G_{3}^{1}$ parameterized by the arc length $s$ with
differential form $ds=dx$, given by
\begin{equation}
\gamma (x)=(x,y(x),z(x)),
\end{equation}%
the formulas $(2.3)$ have the following form
\begin{equation}
\kappa (x)=\sqrt{\left\vert y^{^{\prime \prime }}(x)^{2}-z^{^{\prime \prime
}}(x)^{2}\right\vert },\text{ }\tau (x)=\frac{y^{^{\prime \prime
}}(x)z^{^{\prime \prime \prime }}(x)-y^{^{\prime \prime \prime
}}(x)z^{^{\prime \prime }}(x)}{\kappa ^{2}(x)}.
\end{equation}%
The associated trihedron is given by
\begin{align}
\mathbf{e}_{1}& =\gamma ^{\prime }(x)=(1,y^{^{\prime }}(x),z^{^{\prime
}}(x)), \notag \\
\mathbf{e}_{2}& =\frac{1}{\kappa (x)}\gamma ^{^{\prime \prime }}(x)=\frac{1}{%
\kappa (x)}(0,y^{^{\prime \prime }}(x),z^{^{\prime \prime }}(x)), \notag \\
\mathbf{e}_{3}& =\frac{1}{\kappa (x)}(0,\epsilon z^{^{\prime \prime
}}(x),\epsilon y^{^{\prime \prime }}(x)),
\end{align}%
where $\epsilon =+1$ or $\epsilon =-1$, chosen by criterion det$%
(e_{1},e_{2},e_{3})=1$, that means
\begin{equation*}
\left\vert y^{^{\prime \prime }}(x)^{2}-z^{^{\prime \prime
}}(x)^{2}\right\vert =\epsilon (y^{^{\prime \prime }}(x)^{2}-z^{^{\prime
\prime }}(x)^{2})\text{.}
\end{equation*}%
The curve $\gamma $ given by $(2.4)$ is timelike (resp. spacelike) if $%
\mathbf{e}_{2}(s)$ is a spacelike (resp. timelike) vector. The principal
normal vector or simply normal is spacelike if $\epsilon =+1$ and timelike
if $\epsilon =-1$. For derivatives of the tangent $\mathbf{e}_{1}$, normal $%
\mathbf{e}_{2}$ and binormal $\mathbf{e}_{3}$ vector fields, the following
Frenet formulas in $G_{3}^{1}$ hold:
\begin{align}
\mathbf{e}_{1}^{\prime }(x)& =\kappa (x)\mathbf{e}_{2}(x), \notag \\
\mathbf{e}_{2}^{\prime }(x)& =\tau (x)\mathbf{e}_{3}(x), \notag \\
\mathbf{e}_{3}^{\prime }(x)& =\tau (x)\mathbf{e}_{2}(x).
\end{align}
\section{Frenet formulas according to the equiform geometry of $G_{3}^{1}$}
This section contains some important facts about equiform geometry. The
equiform differential geometry of curves in the pseudo-Galilean space $%
G_{3}^{1}$ has been described in \cite{11}. In the equiform geometry a few
specific terms will be introduced. So, let $\gamma (s):I\rightarrow
G_{3}^{1} $ be an admissible curve in the pseudo-Galilean space $G_{3}^{1}$,
the equiform parameter of $\gamma $ is defined by
\begin{equation*}
\sigma :=\int \frac{1}{\rho }ds=\int \kappa ds,
\end{equation*}%
where $\rho =\frac{1}{\kappa }$ is the radius of curvature of the curve $%
\gamma $. Then, we have
\begin{equation}
\frac{ds}{d\sigma }=\rho .
\end{equation}%
Let $h$ be a homothety with the center in the origin and the coefficient $
\mu $. If we put $\bar{\gamma}=h(\gamma )$, then it follows
\begin{equation*}
\bar{s}=\mu s\text{ \ and \ }\bar{\rho}=\mu \rho ,\text{ }
\end{equation*}%
where $\bar{s}$ is the arc-length parameter of $\bar{\gamma}$ and $\bar{%
\rho}$ the radius of curvature of this curve. Therefore, $\sigma $ is an
equiform invariant parameter of $\gamma $ \cite{11}.
\begin{notation}
The functions $\kappa $ and $\tau $ are not invariants of the homothety
group, then from $(2.3)$ it follows that $\bar{\kappa}=\frac{1}{\mu }\kappa $
and $\bar{\tau}=\frac{1}{\mu }\tau $.
\end{notation}
From now on, we define the Frenet formulas of the curve $\gamma $ with
respect to its equiform invariant parameter $\sigma $ in $G_{3}^{1}.$ The
vector
\begin{equation*}
\mathbf{T}=\frac{d\gamma }{d\sigma },
\end{equation*}%
is called a tangent vector of the curve $\gamma .$ From $(2.6)$ and $(3.1)$
we get
\begin{equation}
\mathbf{T}=\frac{d\gamma }{ds}\frac{ds}{d\sigma }=\rho \cdot \frac{d\gamma }{%
ds}=\rho \cdot \mathbf{e}_{1}.
\end{equation}%
Further, we define the principal normal vector and the binormal vector by
\begin{equation}
\mathbf{N}=\rho \cdot \mathbf{e}_{2},\text{ \ }\mathbf{B}=\rho \cdot \mathbf{%
e}_{3}.
\end{equation}%
It is easy to show that $\left\{ \mathbf{T},\mathbf{N},\mathbf{B}\right\} $
is an equiform invariant frame of $\gamma .$ On the other hand, the
derivatives of these vectors with respect to $\sigma $ are given by
\begin{equation}
\left[
\begin{array}{c}
\mathbf{T} \\
\mathbf{N} \\
\mathbf{B}%
\end{array}%
\right] ^{\prime }=\left[
\begin{array}{ccc}
\dot{\rho} & 1 & 0 \\
0 & \dot{\rho} & \rho \tau \\
0 & \rho \tau & \dot{\rho}%
\end{array}%
\right] \left[
\begin{array}{c}
\mathbf{T} \\
\mathbf{N} \\
\mathbf{B}%
\end{array}%
\right] .
\end{equation}
The functions $\mathcal{K}:I\rightarrow \mathbb{R}$ defined by $\mathcal{K}=%
\dot{\rho}$ is called the equiform curvature of the curve $\gamma $ and $%
\mathcal{T}:I\rightarrow \mathbb{R}$ defined by $\mathcal{T}=\rho \tau =%
\frac{\tau }{\kappa }$ is called the equiform torsion of this curve. In the light of this, the formulas $(3.4)$ analogous to the Frenet
formulas in the equiform geometry of the pseudo-Galilean space $G_{3}^{1}$
can be written as
\begin{equation}
\left[
\begin{array}{c}
\mathbf{T} \\
\mathbf{N} \\
\mathbf{B}%
\end{array}%
\right] ^{\prime }=\left[
\begin{array}{ccc}
\mathcal{K} & 1 & 0 \\
0 & \mathcal{K} & \mathcal{T} \\
0 & \mathcal{T} & \mathcal{K}%
\end{array}%
\right] \left[
\begin{array}{c}
\mathbf{T} \\
\mathbf{N} \\
\mathbf{B}%
\end{array}%
\right] .
\end{equation}%
The equiform parameter $\sigma =\int \kappa (s)ds$ for closed curves is
called the total curvature, and it plays an important role in global
differential geometry of Euclidean space. Also, the function $\frac{\tau }{%
\kappa }$ has been already known as a conical curvature and it also has
interesting geometric interpretation.
\begin{notation}
Let $\gamma :I\rightarrow G_{3}^{1}$ be a Frenet curve in the equiform
geometry of the $G_{3}^{1}$, the following statements are true $($ see for
details \cite{11,13} $)$:
\end{notation}
\begin{enumerate}
\item If $\gamma (s)$ is an isotropic logarithmic spiral in $G_{3}^{1}$.
Then, $\mathcal{K=}$const$.\neq 0$ and $\mathcal{T}=0,$
\item If $\gamma (s)$ is a circular helix in $G_{3}^{1}$. Then, $\mathcal{K=}%
0$ and $\mathcal{T}=$const$.\neq 0,$
\item If $\gamma (s)$ is an isotropic circle in $G_{3}^{1}$. Then, $\mathcal{%
K=}0$ and $\mathcal{T}=0.$
\end{enumerate}
\section{AW($k$)-type curves in the equiform geometry of $G_{3}^{1}$}
Let $\gamma :I\rightarrow G_{3}^{1}$ be a curve in the equiform geometry of
the pseudo-Galilean space $G_{3}^{1}$. The curve $\gamma $ is called a
Frenet curve of osculating order $l$ if its derivatives $\gamma ^{\prime
}(s),\gamma ^{\prime \prime }(s),\gamma ^{\prime \prime \prime
}(s),...,\gamma ^{(l)}(s)$ are linearly dependent and $\gamma ^{\prime
}(s),\gamma ^{\prime \prime }(s),\gamma ^{\prime \prime \prime
}(s),...,\gamma ^{(l+1)}(s)$ are no longer linearly independent for all $%
s\in I$ . To each Frenet curve of order $3$ one can associate an orthonormal
$3$-frame $\left\{ \mathbf{T},\mathbf{N},\mathbf{B}\right\} $ along $\gamma $,
such that $\gamma ^{\prime }(s)=\frac{1}{\rho }\mathbf{T}$, called the
equiform Frenet frame (Eqs. $(3.5)$).
Now, we consider equiform Frenet curevs of osculating order $3$ in $G_{3}^{1}
$ and start with some important results.
Let $\gamma :I\rightarrow G_{3}^{1}$ be a Frenet curve in the equiform
geometry of the pseudo-Galilean space. By the use of Frenet formulas $(3.5)$%
, we obtain the higher order derivatives of $\gamma $ as follows
\begin{align*}
\gamma ^{\prime }(s)& =\frac{d\gamma }{d\sigma }\frac{d\sigma }{ds}=\frac{1}{%
\rho }\mathbf{T}, \\
\gamma ^{\prime \prime }(s)& =\frac{1}{\rho ^{2}}\mathbf{N}, \\
\gamma ^{\prime \prime \prime }(s)& =\frac{1}{\rho ^{3}}\left( -\mathcal{K}%
\mathbf{N}\mathcal{+T}\mathbf{B}\right) , \\
\gamma ^{\prime \prime \prime \prime }(s)& =\frac{1}{\rho ^{4}}[(2\mathcal{K}%
^{2}\mathcal{+\mathcal{T}}^{2}-\mathcal{K}^{\prime })\mathbf{N}+(\mathcal{T}%
^{\prime }-3\mathcal{KT})\mathbf{B}].
\end{align*}
\begin{notation}
Let us write
\begin{align}
Q_{1}& =\frac{1}{\rho ^{2}}\mathbf{N}, \\
Q_{2}& =\frac{1}{\rho ^{3}}\left( -\mathcal{K}\mathbf{N}\mathcal{+T}\mathbf{B%
}\right) , \\
Q_{3}& =\frac{1}{\rho ^{4}}[(2\mathcal{K}^{2}\mathcal{+\mathcal{T}}^{2}-%
\mathcal{K}^{\prime })\mathbf{N}+(\mathcal{T}^{\prime }-3\mathcal{KT})%
\mathbf{B}].
\end{align}
\end{notation}
\begin{notation}
$\gamma ^{\prime }(s),\gamma ^{\prime \prime }(s),\gamma ^{\prime \prime
\prime }(s)$ and $\gamma ^{\prime \prime \prime \prime }(s)$ are linearly
dependent if and only if $Q_{1},Q_{2}$ and $Q_{3}$ are linearly dependent.
\end{notation}
\begin{definition}
Frenet curves (of osculating order $3$) in the equiform geometry of the
pseudo-Galilean space $G_{3}^{1}$ are called \cite{5}:
\begin{enumerate}
\item of type equiform AW$(1)$ if they satisfy \ $Q_{3}=0,$
\item of type equiform AW$(2)$ if they satisfy $\left\Vert Q_{2}\right\Vert
^{2}$ $Q_{3}=\langle Q_{3}\left( s\right) ,Q_{2} \rangle
Q_{2} ,$
\item of type equiform AW$(3)$ if they satisfy $\left\Vert Q_{1}\right\Vert
^{2}$ $Q_{3}=\langle Q_{3} ,Q_{1}\left( s\right) \rangle
Q_{1} ,$
\item of type weak equiform AW$(2)$ if they satisfy%
\begin{equation}
Q_{3}=\left\langle Q_{3},Q_{2}^{\ast }\right\rangle Q_{2}^{\ast },
\end{equation}
\item of type weak equiform AW$\left( 3\right) $ if they satisfy%
\begin{equation}
Q_{3}=\left\langle Q_{3},Q_{1}^{\ast }\right\rangle Q_{1}^{\ast },
\end{equation}%
where%
\begin{eqnarray}
Q_{1}^{\ast } &=&\frac{Q_{1}}{\left\Vert Q_{1}\right\Vert }, \notag \\
Q_{2}^{\ast } &=&\frac{Q_{2}-\left\langle Q_{2},Q_{1}^{\ast }\right\rangle
Q_{1}^{\ast }}{\left\Vert Q_{2}-\left\langle Q_{2},Q_{1}^{\ast
}\right\rangle Q_{1}^{\ast }\right\Vert }.
\end{eqnarray}
\end{enumerate}
\end{definition}
\begin{proposition}
Let $\gamma :I\rightarrow G_{3}^{1}$ be a Frenet curve $($of osculating
order $3)$ in the equiform geometry of the pseudo-Galilean space $G_{3}^{1}$,
(i) $\gamma $ is of type weak equiform AW$(2)$ if and only if
\begin{equation}
2\mathcal{K}^{2}+\mathcal{T}^{2}-\mathcal{K}^{\prime }=0,
\end{equation}
(ii) $\gamma $ is of type weak equiform AW$(2)$ if and only if
\begin{equation}
\mathcal{T}^{\prime }-3\mathcal{KT(}s\mathcal{)}=0.
\end{equation}
\end{proposition}
\begin{proof}
According to Definition 4.1 and Notation 4.1, the proof is obvious.
\end{proof}
\begin{theorem}
Let $\gamma :I\rightarrow G_{3}^{1}$ be a Frenet curve $($of osculating
order $3)$ in the equiform geometry of the pseudo-Galilean space $G_{3}^{1}$%
. Then $\gamma $ is of type equiform AW$(2)$ if and only if%
\begin{equation*}
-\mathcal{K}^{\prime }+2\mathcal{K}^{2}+\mathcal{T}^{2}=0,
\end{equation*}%
\begin{equation}
\text{\ }3\mathcal{KT-T}^{\prime }=0.
\end{equation}
\end{theorem}
\begin{proof}
Since $\gamma $ is of type equiform AW$(2)$, then from $(4.3)$,
we obtain%
\begin{equation*}
\frac{1}{\rho ^{4}}[(2\mathcal{K}^{2}\mathcal{+\mathcal{T}}^{2}(s)-\mathcal{K%
}^{\prime })\mathbf{N}+(\mathcal{T}^{\prime }-3\mathcal{KT})\mathbf{B}]=0.
\end{equation*}%
As we know, the vectors $\mathbf{N}$ and $\mathbf{B}$ are linearly
independent, so we can write%
\begin{equation*}
2\mathcal{K}^{2}\mathcal{+\mathcal{T}}^{2}-\mathcal{K}^{\prime }=0\text{ and
\ }\mathcal{T}^{\prime }-3\mathcal{KT}=0.
\end{equation*}%
The converse statement is straightforward and therefore the proof is
completed.
\end{proof}
\begin{theorem}
Let $\gamma :I\rightarrow G_{3}^{1}$ be a Frenet curve $($of osculating
order $3)$ in the equiform geometry of the pseudo-Galilean space $G_{3}^{1}$%
. Then $\gamma $ is of type equiform AW$(2)$ if and only if%
\begin{equation}
\mathcal{K}^{2}\mathcal{T}-\mathcal{KT}^{\prime }+\mathcal{TK}^{\prime }-%
\mathcal{T}^{3}=0.
\end{equation}
\end{theorem}
\begin{proof}
Assuming that $\gamma $ is a Frenet curve in the equiform geometry of $%
G_{3}^{1}$ , then from $(4.2)$ and $(4.3)$, one can write%
\begin{align*}
Q_{2}& =a_{11}\mathbf{N}+a_{12}\mathbf{B}, \\
Q_{3}& =a_{21}\mathbf{N}+a_{22}\mathbf{B},
\end{align*}%
where $a_{11}$,$a_{12}$, $a_{21}$ and $a_{22}$ are differentiable functions.
Since $Q_{2}$ and $Q_{3}$ are linearly dependent, coefficients
determinant equals zero and hence
\begin{equation}
\left\vert
\begin{array}{cc}
a_{11} & a_{12} \\
a_{21} & a_{22}%
\end{array}%
\right\vert =0,
\end{equation}%
where
\begin{eqnarray}
a_{11} &=&\frac{-1}{\rho ^{3}}\mathcal{K},\text{ }a_{12}=\frac{1}{\rho ^{3}}%
\mathcal{T}, \notag \\
a_{21} &=&\frac{1}{\rho ^{4}}[-\mathcal{K}^{\prime }+2\mathcal{K}^{2}+%
\mathcal{T}^{2}],\text{ } \notag \\
a_{22} &=&\frac{1}{\rho ^{4}}[-3\mathcal{KT+T}^{\prime }].
\end{eqnarray}%
From $(4.11)$ and $(4.12)$, we obtain $(4.10)$. It can be easily shown that
the converse assertion is also true.
\end{proof}
\begin{corollary}
Let $\gamma :I\rightarrow G_{3}^{1}$ be a Frenet curve (of osculating
order3) in the equiform geometry of the pseudo-Galilean space $G_{3}^{1}$,
(i) If $\gamma $ is an isotropic logarithmic spiral in $G_{3}^{1}$, then $%
\gamma $ is of equiform AW$(2)$-type curve.
(ii) If $\gamma $ is an equiform space or timelike general (circular) helix
in $G_{3}^{1}$, then it is not of equiform AW$(k)$, weak AW$(2)$ and weak
AW$(3)$-types.
\end{corollary}
\begin{theorem}
Let $\gamma :I\rightarrow G_{3}^{1}$ be a Frenet curve (of osculating order $%
3$) in the equiform geometry of $G_{3}^{1}$. Then $\gamma $ is of equiform AW%
$\left( 3\right) $-type if and only if
\begin{equation}
\mathcal{T}^{\prime }-3\mathcal{KT}=0.
\end{equation}
\end{theorem}
\begin{proof}
Using Definition $4.1$ and Eqs. $(4.1)$ and $(4.3)$, we obtain $(4.13)$. The
converse direction is obvious, hence our Theorem is proved.
\end{proof}
\section{Bertrand curves of AW$(k)$-type}
\begin{definition}
A curve $\gamma :I\rightarrow G_{3}^{1}$ with equiform curvature $\mathcal{K}%
=0$ is called an equiform Bertrand curve if there exist a curve $\bar{\gamma}%
:I\rightarrow G_{3}^{1}$ with equiform curvature $\mathcal{\bar{K}}=0$ such
that the principal normal lines of $\gamma $ and $\bar{\gamma}$ are parallel
at the corresponding points. In this case $\bar{\gamma}$ is called an
equifrm Bertrand mate of $\gamma $ and vise versa.
\end{definition}
By Definition $5.1$, we can say that for given an equiform Bertrand pair $%
\left( \gamma ,\bar{\gamma}\right) $, there exist a functional relation $%
\bar{s}=\bar{s}(s)$ such that $\lambda (\bar{s}(s)=\lambda (s)$, then the
equiform Bertrand mate of $\gamma $ is given by
\begin{equation}
\bar{\gamma}(s)=\gamma (s)+\lambda \mathbf{N}.
\end{equation}
\begin{theorem}
If $\left( \gamma ,\bar{\gamma}\right) $ is an equiform Bertrand pair in the
equiform geometry of the pseudo-Galilean space $G_{3}^{1}$, then
\begin{description}
\item[(i)] The function $\lambda $ is constant.
\item[(ii)] $\gamma $ with non-zero constant equiform torsion is a circular
helix in $G_{3}^{1}$.
\item[(iii)] $\gamma $ with zero equiform torsion is an isotropic circle of $%
G_{3}^{1}$
\end{description}
\end{theorem}
\begin{proof}
Along $\gamma $ and $\bar{\gamma}$, let $\{\mathbf{T},\mathbf{N},\mathbf{B}\}
$ and $\{\mathbf{\bar{T}},\mathbf{\bar{N}},\mathbf{\bar{B}}\}$ be the Frenet
frames according to the equiform geometry of the pseudo-Galilean space $%
G_{3}^{1}$, respectively. Differentiate $(5.1)$ with respect to $s$, we
obtain
\begin{equation}
\mathbf{\bar{T}}=\mathbf{T}+\lambda \mathbf{N}^{\prime }+\lambda ^{\prime }%
\mathbf{N}.
\end{equation}%
By using $(3.5)$, we have
\begin{equation*}
\mathbf{\bar{T}}=\mathbf{T}+\left( \lambda \mathcal{K+\lambda }^{\prime
}\right) \mathbf{N}+\lambda \mathcal{T}\mathbf{B}\text{.}
\end{equation*}%
Since $\mathbf{\bar{N}\ }$is parallel to $\mathbf{N}$, we get
\begin{equation*}
\lambda \mathcal{K+\lambda }^{\prime }\mathcal{=}0,
\end{equation*}%
it follows that
\begin{equation*}
\lambda =const.
\end{equation*}
If $\gamma $ has a non-zero constant equiform torsion, then $\gamma $ is
characterized by
\begin{equation*}
\kappa =const.\neq 0,~\tau =const.\neq 0,
\end{equation*}%
and therefore $\tau /\kappa =const.$holds.
On the other hand, whenever $\mathcal{T}=0$, the natural equations of $%
\gamma $ is given by
\begin{equation*}
\kappa =const.\neq 0,~\tau =0,
\end{equation*}%
and so, the curve $\gamma $ is an isotropic circle in $G_{3}^{1}$ \cite{14}.
Thus the proof is completed.
\end{proof}
\begin{theorem}
If $\left( \gamma ,\bar{\gamma}\right) $ is a Bertrand pair in the equiform
geometry of the pseudo-Galilean space $G_{3}^{1}$, then the angle between
tangent vectors at corresponding points is constant.
\end{theorem}
\begin{proof}
To prove that the angle is constant, we need to show that $\left\langle
\mathbf{\bar{T}},\mathbf{T}\right\rangle ^{\prime }=0$. For this purpose
using $(3.5)$ to obtain
\begin{align}
\left\langle \mathbf{\bar{T}},\mathbf{T}\right\rangle ^{\prime }&
=\left\langle \mathbf{\bar{T}}^{\prime },\mathbf{T}\right\rangle
+\left\langle \mathbf{\bar{T}},\mathbf{T}^{\prime }\right\rangle \notag \\
& =\left\langle \mathcal{\bar{K}}\mathbf{\bar{T}}+\mathbf{\bar{N}},\mathbf{T}%
\right\rangle +\left\langle \mathbf{\bar{T}},\mathcal{K}\mathbf{T}+\mathbf{N}%
\right\rangle \notag \\
& =\mathcal{\tilde{K}}\left\langle \mathbf{\bar{T}},\mathbf{T}\right\rangle
+\left\langle \mathbf{\bar{N}},\mathbf{T}\right\rangle +\mathcal{K}%
\left\langle \mathbf{\bar{T}},\mathbf{T}\right\rangle \notag \\
& \ \ \ +\left\langle \mathbf{\bar{T}},\mathbf{N}\right\rangle .\text{ \ \ \
\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ }
\end{align}%
Because of $\mathbf{\bar{N}}$ is parallel to $\mathbf{N},$ then%
\begin{equation}
\left\langle \mathbf{\bar{N}},\mathbf{T}\right\rangle =0,\left\langle
\mathbf{\bar{T}},\mathbf{N}\right\rangle =0.
\end{equation}%
Since $\left( \gamma ,\bar{\gamma}\right) $ is a Berrand pair in the
equiform geometry of $G_{3}^{1}$, then from Theorem $5.1$, we have%
\begin{equation}
\mathcal{K=}0\text{ and }\mathcal{\bar{K}=}0.
\end{equation}%
After substituting $(5.4)$ and $(5.5)$ into $(5.3)$, we get
\begin{equation}
\left\langle \mathbf{\bar{T}},\mathbf{T}\right\rangle ^{\prime }=0.
\end{equation}%
In the light of $(5.6)$ the angle between $\mathbf{\bar{T}},\mathbf{T}$ is
constant. Thus this completes the proof.
\end{proof}
\begin{corollary}
Let $\gamma (s):I\rightarrow G_{3}^{1}$ be a Bertrand curve in the equiform
geometry of $G_{3}^{1}.$ Then
(i) $\gamma $ is a weak equiform AW$(3)$-type but not a weak equiform AW$(2)$%
-type.
(ii) $\gamma $ is equiform AW$(3)$-type but not equiform AW$(1)$ and AW$(2)$%
-types.
\end{corollary}
\section{Examples}
We consider some examples (timelike and spacelike curves \cite{11,12}) which
characterize equiform general (circular) helices with respect to the Frenet
frame $\left\{ \mathbf{T},\mathbf{N},\mathbf{B}\right\} $ in the equiform
geometry of $G_{3}^{1}$ which satisfy some conditions of equiform curvatures
($\mathcal{K=K}(s),\mathcal{T=T}(s);~\mathcal{K=}const.\neq 0,\mathcal{T=}%
const.\neq 0;~\mathcal{K=}const.\neq 0,\mathcal{T=}0$).
\begin{example}
Consider the equiform \textbf{timelike} general helix $\mathbf{r}%
:I\longrightarrow G_{3}^{1},I\subseteq \mathbb{R}$ parameterized by the arc
length $s$ with differential form $ds=dx,$ given by
\begin{equation*}
\mathbf{r}(x)=(x,y(x),z(x)),
\end{equation*}%
where
\begin{eqnarray*}
x(s) &=&s, \\
y(s) &=&\frac{e^{-as}}{\left( a^{2}-b^{2}\right) ^{2}}\left( \left(
a^{2}+b^{2}\right) \cosh \left( bs\right) +2ab\sinh \left( bs\right) \right)
, \\
z(s) &=&\frac{e^{-as}}{\left( a^{2}-b^{2}\right) ^{2}}\left( 2ab\cosh \left(
bs\right) +\left( a^{2}+b^{2}\right) \sinh \left( bs\right) \right) ; \\
a,b&\in&\mathbb{R}-\left\{ 0\right\}.
\end{eqnarray*}%
The corresponding derivatives of $\mathbf{r}$\textbf{\ }are as follows
\begin{eqnarray*}
\mathbf{r}^{\prime } &=&\left( 1,\frac{-e^{-as}}{\left( a^{2}-b^{2}\right) }%
\left( a\cosh \left( bs\right) +b\sinh \left( bs\right) \right) ,\frac{%
e^{-as}}{\left( b^{2}-a^{2}\right) }\left( b\cosh \left( bs\right) +a\sinh
\left( bs\right) \right) \right) , \\
\mathbf{r}^{\prime \prime } &=&\left( 0,e^{-as}\cosh \left( bs\right)
,e^{-as}\sinh \left( bs\right) \right) , \\
\mathbf{r}^{\prime \prime \prime } &=&\left( 0,e^{-as}\left( -a\cosh \left(
bs\right) +b\sinh \left( bs\right) \right) ,e^{-as}\left( b\cosh \left(
bs\right) -a\sinh \left( bs\right) \right) \right) .
\end{eqnarray*}
First of all, we find that the tangent vector of $\mathbf{r}$ has the form
\begin{eqnarray*}
\mathbf{e}_{1} &=&\left( x^{\prime },y^{\prime },z^{\prime }\right) \\
&=&\left( 1,\frac{-e^{-as}}{\left( a^{2}-b^{2}\right) }\left( a\cosh \left(
bs\right) +b\sinh \left( bs\right) \right) ,\frac{e^{-as}}{\left(
b^{2}-a^{2}\right) }\left( b\cosh \left( bs\right) +a\sinh \left( bs\right)
\right) \right) .
\end{eqnarray*}
Then the two normals (normal and binormal) of the curve are, respectively%
\begin{eqnarray*}
\mathbf{e}_{2} &=&\left( 0,\cosh \left( bs\right) ,\sinh \left( bs\right)
\right) , \\
\mathbf{e}_{3} &=&\left( 0,\sinh \left( bs\right) ,\cosh \left( bs\right)
\right) ;~~\det [\mathbf{e}_{1},\mathbf{e}_{2},\mathbf{e}_{3}]=1.
\end{eqnarray*}
Thus the computations of the coordinate functions of $\mathbf{r}$ lead to%
\begin{equation*}
\kappa =e^{-as},~\tau =b~.
\end{equation*}
From the equiform Frenet formulas $(3.5)$ we can express vector fields $%
\mathbf{T},\mathbf{N},\mathbf{B}$ as follows
\begin{eqnarray*}
\mathbf{T} &=&\left( e^{as},\frac{-1}{\left( a^{2}-b^{2}\right) }\left(
a\cosh \left( bs\right) +b\sinh \left( bs\right) \right) ,\frac{1}{\left(
b^{2}-a^{2}\right) }\left( b\cosh \left( bs\right) +a\sinh \left( bs\right)
\right) \right) , \\
\mathbf{N} &=&\left( 0,e^{as}\cosh \left( bs\right) ,e^{as}\sinh \left(
bs\right) \right) , \\
\mathbf{B} &=&\left( 0,e^{as}\sinh \left( bs\right) ,e^{as}\cosh \left(
bs\right) \right) ,
\end{eqnarray*}%
respectively.
In the light of this, the equiform curvatures are given by%
\begin{equation*}
~\mathcal{K}=ae^{as},\mathcal{T}=-be^{as}.
\end{equation*}
\end{example}
\begin{center}
\begin{figure}[h]
\centering
\includegraphics[width=5cm]{timelikeg2.png} \label{fig:timelikeg2}
\caption{Equiform timelike general helix with $\mathcal{K}(s)=e^{s},\mathcal{%
T}(s)=2e^{s}$.}
\end{figure}
\end{center}
\begin{example}
Let $\mathbf{r}:I\longrightarrow G_{3}^{1},I\subseteq \mathbb{R}$ be the equiform
\textbf{spacelike} general helix, given by
\begin{equation*}
\mathbf{r}(x)=(x,y(x),z(x)),
\end{equation*}%
where
\begin{eqnarray*}
x(s) &=&s, \\
y(s) &=&\frac{e^{-as}}{\left( a^{2}-b^{2}\right) ^{2}}\left( 2ab\cosh \left(
bs\right) +\left( a^{2}+b^{2}\right) \sinh \left( bs\right) \right) , \\
z(s) &=&\frac{e^{-as}}{\left( a^{2}-b^{2}\right) ^{2}}\left( \left(
a^{2}+b^{2}\right) \cosh \left( bs\right) +2ab\sinh \left( bs\right) \right)
; \\
a,b &\in &\mathbb{R}-\left\{ 0\right\} .
\end{eqnarray*}%
For the coordinate functions of $\mathbf{r}$, we have
\begin{eqnarray*}
\mathbf{r}^{\prime } &=&\left( 1,\frac{e^{-as}}{\left( b^{2}-a^{2}\right) }%
\left( b\cosh \left( bs\right) +a\sinh \left( bs\right) \right) ,\frac{%
-e^{-as}}{\left( a^{2}-b^{2}\right) }\left( a\cosh \left( bs\right) +b\sinh
\left( bs\right) \right) \right) , \\
\mathbf{r}^{\prime \prime } &=&\left( 0,e^{-as}\sinh \left( bs\right)
,e^{-as}\cosh \left( bs\right) \right) , \\
\mathbf{r}^{\prime \prime \prime } &=&\left( 0,e^{-as}\left( b\cosh \left(
bs\right) -a\sinh \left( bs\right) \right) ,e^{-as}\left( b\sinh \left(
bs\right) -a\cosh \left( bs\right) \right) \right) .
\end{eqnarray*}%
Also, the associated trihedron is given by
\begin{eqnarray*}
\mathbf{e}_{1} &=&\left( 1,\frac{e^{-as}}{\left( b^{2}-a^{2}\right) }\left(
b\cosh \left( bs\right) +a\sinh \left( bs\right) \right) ,\frac{-e^{-as}}{%
\left( a^{2}-b^{2}\right) }\left( a\cosh \left( bs\right) +b\sinh \left(
bs\right) \right) \right) , \\
\mathbf{e}_{2} &=&\left( 0,\sinh \left( bs\right) ,\cosh \left( bs\right)
\right) , \\
\mathbf{e}_{3} &=&\left( 0,-\cosh \left( bs\right) ,-\sinh \left( bs\right)
\right) .~
\end{eqnarray*}
The curvature and torsion of this curve are
\begin{equation*}
\kappa =e^{-as},~\tau =-b~.~
\end{equation*}
Furthermore, the tangent, normal and binormal vector fields in the equiform
geometry of $G_{3}^{1}$ are obtained as follows
\begin{eqnarray*}
\mathbf{T} &=&\left( e^{as},\frac{1}{\left( b^{2}-a^{2}\right) }\left(
b\cosh \left( bs\right) +a\sinh \left( bs\right) \right) ,\frac{-1}{\left(
a^{2}-b^{2}\right) }\left( a\cosh \left( bs\right) +b\sinh \left( bs\right)
\right) \right) , \\
\mathbf{N} &=&\left( 0,e^{as}\sinh \left( bs\right) ,e^{as}\cosh \left(
bs\right) \right) , \\
\mathbf{B} &=&\left( 0,-e^{as}\cosh \left( bs\right) ,-e^{as}\sinh \left(
bs\right) \right) ,
\end{eqnarray*}%
respectively.
The equiform curvatures of $\mathbf{r}$ are
\begin{equation*}
~\mathcal{K}=ae^{as},\mathcal{T}=-be^{as}.
\end{equation*}
\end{example}
\begin{center}
\begin{figure}[h]
\centering
\includegraphics[width=6.5cm]{spacelikeg2.png} \label{fig:spacelikeg2}
\caption{Equiform spacelike general helix with $\mathcal{K}(s)=e^{s},%
\mathcal{T}(s)=-2e^{s}$.}
\end{figure}
\end{center}
\begin{example}
In this example, let us consider the equiform timelike\textbf{\ circular}
helix $\mathbf{r}:I\longrightarrow G_{3}^{1}$ given by
\begin{equation*}
\mathbf{r}(x)=(x,y(x),z(x)),
\end{equation*}%
where
\begin{eqnarray*}
x(s) &=&s, \\
y(s) &=&\frac{a^{3}s}{b\left( b^{2}-a^{2}\right) }\left( b\sinh \left( \frac{%
b}{a}\ln (as)\right) -a\cosh \left( \frac{b}{a}\ln (as)\right) \right) , \\
z(s) &=&\frac{a^{3}s}{b\left( b^{2}-a^{2}\right) }\left( b\cosh \left( \frac{%
b}{a}\ln (as)\right) -a\sinh \left( \frac{b}{a}\ln (as)\right) \right) ; \\
a,b &\in &\mathbb{R}-\left\{ 0\right\} .
\end{eqnarray*}%
For this curve, the equiform vector fields are obtained as follows
\begin{eqnarray*}
\mathbf{T} &=&\left( \frac{s}{a},\frac{as}{b}\cosh \left( \frac{b}{a}\ln
(as)\right) ,\frac{as}{b}\sinh \left( \frac{b}{a}\ln (as)\right) \right) , \\
\mathbf{N} &=&\left( 0,\frac{s}{a}\sinh \left( \frac{b}{a}\ln (as)\right) ,%
\frac{s}{a}\cosh \left( \frac{b}{a}\ln (as)\right) \right) , \\
\mathbf{B} &=&\left( 0,\frac{s}{a}\cosh \left( \frac{b}{a}\ln (as)\right) ,%
\frac{s}{a}\sinh \left( \frac{b}{a}\ln (as)\right) \right) ,
\end{eqnarray*}%
respectively.
It follows that%
\begin{equation*}
\mathcal{K}=\frac{1}{a},\mathcal{T}=\frac{-b}{a^{2}}.
\end{equation*}
\end{example}
\begin{center}
\begin{figure}[h]
\centering
\includegraphics[width=6cm]{circulart.png} \label{fig:circulart}
\caption{Equiform timelike circular helix with $\mathcal{K}=\frac{1}{a},%
\mathcal{T}=\frac{-b}{a^{2}}$.}
\end{figure}
\end{center}
\begin{example}
Let the equiform \textbf{spacelike }circular helix $\mathbf{r}%
:I\longrightarrow G_{3}^{1},I\subseteq \mathbb{R}$ in the form
\begin{equation*}
\mathbf{r}(x)=(x,y(x),z(x)),
\end{equation*}%
where
\begin{eqnarray*}
x(s) &=&s, \\
y(s) &=&\frac{a^{3}s}{b\left( b^{2}-a^{2}\right) }\left( b\cosh \left( \frac{%
b}{a}\ln (as)\right) -a\sinh \left( \frac{b}{a}\ln (as)\right) \right) , \\
z(s) &=&\frac{a^{3}s}{b\left( b^{2}-a^{2}\right) }\left( b\sinh \left( \frac{%
b}{a}\ln (as)\right) -a\cosh \left( \frac{b}{a}\ln (as)\right) \right) ; \\
a,b &\in &\mathbb{R}-\left\{ 0\right\} .~
\end{eqnarray*}%
Here, the equiform differntial vectors are respectively, as follows
\begin{eqnarray*}
\mathbf{T} &=&\left( \frac{s}{a},\frac{as}{b}\sinh \left( \frac{b}{a}\ln
(as)\right) ,\frac{as}{b}\cosh \left( \frac{b}{a}\ln (as)\right) \right) , \\
\mathbf{N} &=&\left( 0,\frac{s}{a}\cosh \left( \frac{b}{a}\ln (as)\right) ,%
\frac{s}{a}\sinh \left( \frac{b}{a}\ln (as)\right) \right) , \\
\mathbf{B} &=&\left( 0,-\frac{s}{a}\sinh \left( \frac{b}{a}\ln (as)\right) ,-%
\frac{s}{a}\cosh \left( \frac{b}{a}\ln (as)\right) \right) .
\end{eqnarray*}%
Equiform curvature and equiform torsion are calculated as follows%
\begin{equation*}
\mathcal{K}=\frac{1}{a},\mathcal{T}=\frac{b}{a^{2}}.
\end{equation*}
\end{example}
\begin{center}
\begin{figure}[h]
\centering
\includegraphics[width=5cm]{circulars.png} \label{fig:circulars}
\caption{Equiform spacelike circular helix with $\mathcal{K}=\frac{1}{a},%
\mathcal{T}=\frac{b}{a^{2}}$.}
\end{figure}
\end{center}
\begin{example}
If we consider the equiform \textbf{timelike} isotropic logarithmic spiral $%
\mathbf{r}:I\longrightarrow G_{3}^{1},I\subseteq \mathbb{R}$ parameterized by the arc
length $s$ with differential form $ds=dx,$ given by%
\begin{equation*}
\mathbf{r}(x)=(x,y(x),0),
\end{equation*}%
where
\begin{eqnarray*}
x(s) &=&s, \\
y(s) &=&\frac{as+b}{a^{2}}\left( \ln (as+b)-1\right) , \\
z(s) &=&0; \\
a,b &\in &\mathbb{R}-\left\{ 0\right\} .
\end{eqnarray*}%
For this curve, we get%
\begin{eqnarray*}
\mathbf{r}^{\prime } &=&\left( 1,\frac{\ln (as+b)}{a},0\right) , \\
\mathbf{r}^{\prime \prime } &=&\left( 0,\frac{1}{as+b},0\right) , \\
\mathbf{r}^{\prime \prime \prime } &=&\left( 0,\frac{-a}{\left( as+b\right)
^{2}},0\right) ,
\end{eqnarray*}%
and
\begin{eqnarray*}
\mathbf{e}_{1} &=&\left( 1,\frac{\ln (as+b)}{a},0\right) , \\
\mathbf{e}_{2} &=&\left( 0,1,0\right) , \\
\mathbf{e}_{3} &=&\left( 0,0,1\right) ;~\kappa =\frac{1}{as+b},~\tau =0.
\end{eqnarray*}%
In this case, equiform Frenet vectors and equiform curvatures are as follows
\begin{eqnarray*}
\mathbf{T} &=&\left( as+b,\frac{\left( as+b\right) \ln (as+b)}{a},0\right) ,
\\
\mathbf{N} &=&\left( 0,as+b,0\right) , \\
\mathbf{B} &=&\left( 0,0,as+b\right) ,~\mathcal{K}=a,\mathcal{T}=0.
\end{eqnarray*}%
respectively.
\end{example}
\begin{center}
\begin{figure}[h]
\centering
\includegraphics[width=5cm]{spiral11.png} \label{fig:spiral11}
\caption{Equiform timelike isotropic logarithmic spiral with $\mathcal{K}%
(s)=1,\mathcal{T}(s)=0$.}
\end{figure}
\end{center}
From aforementioned calculations, according to (\textbf{Proposition }$%
\mathbf{4.2}$\textbf{\ and Theorems }$\mathbf{4.1-4.3}$), examples $1-4$ are
not characterize curves of equiform AW$(k)$, weak equiform AW$(2)$ and weak
equiform AW$(3)$-types. On the other hand, the last example shows that the
curve is of equiform AW$(2)$ and AW$(3)$-types and it is not of equiform AW$%
(1)$-type. Also, it is of weak equiform AW$(2)$ and not of weak equiform AW$%
(3)$-types.
\section{Conclusion}
In this paper, we have considered some special curves of equiform AW$(k)$-type of the pseudo-Galilean $3$-space. Also, using the equiform curvature
conditions of these curves, the necessary and sufficient conditions for them
to be equiform AW$(k)$ and weak equiform AW$(k)$-types are given.
Furthermore, several examples to confirm our main results have been given
and illustrated.
|
\section{Introduction}
\label{sec:intro}
Inflationary cosmology which was proposed in the early 1980's, extends the standard Big-Bang model by postulating an early epoch of nearly exponential expansion in order to resolve a number of puzzles of the Big-Bang cosmology such as flatness, horizon and monopole problems \cite{ref1,ref2,ref3}. Inflation also explains the origin of the CMB anisotropies and large scale structure of the cosmos, indeed quantum vacuum fluctuations of the inflation field(s) magnified to the cosmic sized classical perturbations after the horizon exit time and became the seeds for the growth of the structure and CMB anisotropies in the universe\cite{ref4,ref5,ref6,ref7}. Before the advent of inflation the initial perturbations were \textit{postulated} and their spectrum was supposed to be scalar-invariant in order to fit observational data\cite{ref8,ref9,ref10}. On the other hand, inflationary theory not only truly explains the origin of the primordial inhomogeneities but also predicts their spectrum. The spectrum of these inhomogeneities, as well as the spectrum of cosmological gravitational waves produced during the inflation, are about the only observational test of the inflationary theories. Cosmological observations are consistent with the simplest model of inflation within the slow-roll paradigm \cite{ref11,ref12}. According to this scenario, the curvature power spectrum is nearly flat\cite{ref13,ref14,ref15,ref16,ref17,ref18,ref19}i.e.
\begin{equation}\label{1}
\mathcal{R}^o_q\propto q^{-\frac{3}{2}-2\epsilon-\delta},
\end{equation}
where $ \mathcal{R}_q $ is the Fourier component of comoving curvature perturbation with comoving wave number $ q $ (the superscript "o" is standing for "outside the Hubble horizon"). Furthermore, $ \epsilon $ and $ \delta $ are respectively first and second slow-roll parameters. According to the observational data $ \epsilon \leq 0.008 $ and $ \delta\leq 0.018 $ \cite{ref11}. $\mathcal{R}$ characterizes the adiabatic scalar perturbations which for super-Hubble scales has a constant value \cite{ref20,ref21}. On the other hand, all inflationary models predict the existence of cosmological gravitational waves which produce a B-mode polarization pattern in the CMB anisotropies. Recently this mode has been detected by the BICEP2 collaboration \cite{ref22}. In the slow-roll approximation we have \cite{ref13}
\begin{equation}\label{2}
\mathcal{D}^o_q\propto q^{-\frac{3}{2}-\epsilon},
\end{equation}
where $ \mathcal{D}_q $ is the amplitude of inflationary gravitational waves. The relative amplitude which is characterized by the tensor-scalar ratio $ r=4\vert\frac{\mathcal{D}^o_q}{\mathcal{R}^o_q}\vert^2 $ is a probe of the energy scale in the inflationary epoch. It can be shown that in slow-roll approximation with a single scalar field $ r=16\epsilon $ \cite{ref13}. The BICEP2 collaboration has reported $ r\simeq 0.2 $ which is greater than the upper limit $ r<0.11 $ obtained by the Planck collaboration \cite{ref11}. In addition to this inconsistency, there are another discrepancy which refers to the running parameter of the curvature spectral index. In the slow-roll single field inflation the running parameter is of the second order in terms of slow-roll parameters, but the Planck data prefer $ \mathfrak{N}_r=\frac{\partial \mathfrak{N}_s}{\partial ln q}\simeq -0.015 $ \cite{ref11} which is of the first order slow-roll parameters and so has no any justification in slow-rolling inflationary model. In other words, the running parameter has magnitude significantly greater then the slow-roll paradigm prediction in spatially flat inflationary universe. On the other hand, there are some anomalies in the CMB power spectrum such as suppression of the lowest CMB multiples \cite{ref23} and lack of temperature correlations on scales beyond $ 70^\circ $ \cite{ref24} which may be evidences for discrete spectrum and non-trivial spherical topologies \cite{ref25,ref23}. In other words, some positive curvature models with non-trivial topology can solve the problem of the CMB quadrupole and octopole suppression and also mystery of missing fluctuations which apear in the concordance model \cite{ref26,ref27}. Furthermore, as we know the inflationary universe background is described by quasi-de Sitter space. However, maximally extended de Sitter space which is known as the \textit{Lorentzian de Sitter space} is included in the FLRW models with $ K=+1 $ and so has the positive spatial curvature \cite{ref28}. Lorentzian de Sitter is geodesically complete too. Besides, the last observational data does not roll out $ \Omega_K<0 $ case as well \cite{ref29}. It is noticeable that if the spatial curvature of the universe is positive then, the curvature dominates at early times in inflationary era\cite{ref25,ref30} , so the curvature may be significant in primordial spectra of perturbations and cannot be ignored. Dynamics of the inflationary universe with positive spatial curvature has been studied by Ellis and his collaborators\cite{ref30,ref31}. They showed that whatever the number of e-foldings increases, the curvature parameter decreases and the universe would be closer to flat today\cite{ref30}. On the other hand, Vilenkin discussed a cosmological model in which the inflationary universe is created by a quantum tunneling from \textit{nothing}\cite{ref32,ref33,ref34}. This model doesn’t have Big-Bang singularity and predicts that the inflationary universe is positively curved. Although Linde has claimed that it is very difficult to obtain a realistic model of a closed inflationary universe\cite{ref35}, Ellis and Maartens has constructed a single field inflationary model in the closed universe which is known as the \textit{eternal emergent universe scenario}\cite{ref36,ref37}. This model is a nonsingular closed inflationary cosmology that begins from a meta-stable Einstein static state. Another closed inflationary model with positive curvature index has been introduced by Lasenby and Doran\cite{ref38}.\\
In this article we investigate the slow-rolling inflationary scenario in a spatially closed background with trivial topology namely positively curved FLRW universe. The layout of the article is as follows: In section 2, we derive the generalized Mukhanov-Sasaki equation associated to the positively curved universe. Then, the slow-roll parameters and also Sasaki-Mukhanov variable is generalized to the inflationary universe with positive curvature index. Section 2 is concluded with the calculation of the comoving power spectrum. Section 3 is allotted the investigation of gravitational wave spectrum in the positively curved universe and section 4 includes the calculation of the tensor-scalar ratio in the FLRW universe with positive curvature index. Conclusion is given in the final section.
\section{Curvature power spectrum in the positively curved FLRW universe }
\label{sec:2}
\subsection{the Mukhanov-Sasaki equation associated with the positively curved inflationary universe}
\label{sec:2.1}
In order to find the curvature power spectrum in a spatially closed universe, we should generalize the ordinary Mukhanov-Sasaki equation \cite{ref7,ref13,ref39,ref40} to the case $ K=+1 $ ($ K $ is the curvature index in the FLRW metric). This equation describes the evolution of the comoving curvature perturbation in the inflationary epoch. For this purpose, we suppose the homogeneous inflation field $ \bar{\Phi}\left( t\right) $ has been perturbed by a small fluctuations $ \delta \Phi\left(t,\mathbf{x} \right) $ during the inflation era (hereafter bar over any quantity stands for its unperturbed value). This fluctuations are accompanied by the (scalar) perturbation in the FLRW metric (with $ K=+1 $ ). According to this perturbation, the line element of the universe may be written as\cite{ref13}
\begin{align}
&ds^2 = - \left( {1 + E} \right)dt^2 + 2a\left( \partial _i F \right)dt d{x^i} +a^2 \left(1 + 2\mathcal{R} \right)\tilde{ g}_{ij}dx^idx^j ,\label{3}\\
&\tilde{g}_{ij}=\delta_{ij}+\frac{x^ix^j}{1-\mathbf{x}^2},\nonumber
\end{align}
which is the FLRW metric with $ K=+1 $ in the comoving quasi-Cartesian coordinates $ x^{i} $ plus the scalar linear perturbation in the \textit{comoving gauge}. Here $ E $ and $ \mathcal{R} $ are respectively the \textit{lapse function} and \textit{comoving curvature perturbation}. It can be shown that in the comoving gauge\cite{ref13}
\begin{eqnarray}
&&\delta{\rho} = \delta p =- \frac{1}{2}E\dot{\bar{\Phi}}^2,\label{4}\\
&&\delta{\Phi} = 0,\label{5}
\end{eqnarray}
where $ \rho $ and $ p $ are energy density and pressure of the perfect fluid associated to the inflaton (dot stands for the derivation respect to the cosmic time $ t $ ). On the other hand, according to the perturbative field equations as well as the energy conservation law $ E $ , $ F $ and $ \mathcal{R} $ don’t evolve independently and we have\cite{ref41}
\begin{align}
& \frac{4\mathcal{R}}{a^2}+Ha\nabla ^2 F-6H \dot{\mathcal{R}}-\ddot{\mathcal{R}}+\left(3H^2+\dot{H} \right)E+\frac{1}{2}H\dot{E}+\nabla ^2 \mathcal{R}=0,\label{6} \\
&2\dot{\mathcal{R}}-HE+\frac{2}{a}F=0,\label{7} \\
& 4HaF+2a\dot{F}+E+2\mathcal{R}=0,\label{8}\\
&\dot{\delta\rho}-\dot{\bar{\Phi}}^2\left(a\nabla ^2 F-3\dot{\mathcal{R}}+3H E \right)=0,\label{9}
\end{align}
where equation (\ref{9}) is the energy conservation law. Notice that $ \nabla ^2 =\frac{1}{a^2}\tilde{g}^{ij}\nabla _i\nabla_j $ is the \textit{Laplace-Beltrami operator} respect to the $ a^2\tilde{g}_{ij} $. After some tedious and lengthy calculation, we may combine these equations and extract an explicit equation in terms of $ \mathcal{R} $
\begin{align}
&\Bigg[Ha^2\left(n^2-4 \right)+\frac{1}{H}-\frac{\dot{H}a^2}{H} \Bigg] \ddot{\mathcal{R}}_\mathbf{n}+\Bigg[Ha^2\frac{\dot{\boldsymbol{\chi}}}{\boldsymbol{\chi}}\left(n^2-4 \right)-\dot{H}a^2 \left(2n^2-5 \right)\nonumber\\
& \qquad +3H^2a^2\left(n^2-4 \right)+3\Bigg] \dot{\mathcal{R}}_\mathbf{n}+\Bigg[H\left(n^2-4 \right)\left(n^2-5 \right)+\frac{\dot{H}}{H}\left(n^2-3\right)\nonumber\\
&\qquad +\frac{1}{Ha^2}\left(n^2-5\right)-\frac{\dot{\boldsymbol{\chi}}}{\boldsymbol{\chi}}\left(n^2-4 \right) \Bigg] \mathcal{R}_\mathbf{n}=0, \label{10}
\end{align}
where $ \boldsymbol{\chi}=\dot{H}-\frac{1}{a^2} $ and $ \mathcal{R}_\mathbf{n} $ is the Fourier component of $ \mathcal{R} $ with comoving canonical wave number $ n $. Notice that $ \mathbf{n}=\left(n,l,m \right) $ where $ n=3,4,..., 0\leq l\leq n-1 $ and $ \vert m\vert\leq l $ \cite{ref41}. Here due to the compactness of spatial section of spacetime, the comoving wave number is discrete. Furthermore, wave numbers $ n=1,2 $ correspond to the pure gauge \cite{ref42,ref43}, so we ignore them. One can rewrite equation (\ref{10}) in terms of conformal time $ \tau $
\begin{align}
&\Bigg[\left(n^2 - 3 \right)\mathcal{H} + \frac{1}{\mathcal{H}} - \frac{\mathcal{H}'}{\mathcal{H}} \Bigg]\mathcal{R}''_\mathbf{n} + \Bigg[2\left(n^2 - 3 \right)\mathcal{H}^2- 2\left(n^2 - 3 \right)\mathcal{H}' + \left(n^2 - 4 \right)\mathcal{H}\frac{\chi '}{\chi} + 2 \Bigg]\mathcal{R}'_\mathbf{n} \nonumber\\
&+ \Bigg[ \left(n^2-3\right)\left( n^2-5\right) \mathcal{H} + \left( n^2 - 3 \right)\frac{\mathcal{H}'}{\mathcal{H}} + \left( n^2 - 5 \right)\frac{1}{\mathcal{H}} - \left( n^2 - 4 \right)\frac{\chi '}{\chi} \Bigg]\mathcal{R}_\mathbf{n} = 0,\label{11}
\end{align}
where the prime symbol indicates derivation respect to the conformal time. Moreover, $ \mathcal{H}=Ha $ is the comoving Hubble parameter and $ \chi=\mathcal{H}^2-\mathcal{H}'+1=4\pi G \bar{\phi}'^2 $ (indeed $ \boldsymbol{\chi}=-\frac{\chi}{a^2} $). Equation (\ref{11}) is the \textit{generalized Mukhanov-Sasaki equation} for the inflationary universe with positive curvature index.
\subsection{Re-definition of the slow-roll parameters; generalized Sasaki-Mukhanov variable}
\label{sec:2.2}
Now let’s consider the slow-roll inflation which guarantees slowly variation of inflaton by considering a Coleman–Weinberg type potential. In general, slow-roll inflation may be described by two flatness conditions\cite{ref7,ref13}
\begin{align}
&\dot{\bar{\phi}}^2\ll V\left( \bar{\phi} \right),\label{12}\\
&\vert \ddot{\bar{\phi}}\vert \ll H\vert \dot{\bar{\phi}}\vert . \label{13}
\end{align}
In spatially flat case relations (\ref{12}) and (\ref{13}) reduce to
\begin{align}
&\epsilon:=-\frac{\dot{H}}{H^2}\ll 1,\label{14}\\
&\delta:=\frac{\ddot{H}}{2H\dot{H}}\ll 1\label{15}.
\end{align}
Here $ \epsilon $ and $ \delta $ are respectively first and second slow-roll parameters which are considered roughly constant. On the other hand, for the positively curved inflationary universe, the flatness conditions may be written as the same as equations (\ref{14}) and (\ref{15}) by re-definition of the slow-roll parameters
\begin{align}
&\epsilon:=-\frac{\boldsymbol{\chi}}{H^2+\frac{1}{a^2}}= \frac{\mathcal{H}^2-\mathcal{H}'+1}{\mathcal{H}^2+1}\ll 1,\label{16}\\
&\delta:= \frac{1}{2H}\frac{\dot{\boldsymbol{\chi}}}{\boldsymbol{\chi}}= \frac{1}{2\mathcal{H}}\frac{\chi'}{\chi}-1\ll 1.\label{17}
\end{align}
One can re-write equation (\ref{16}) as
\begin{equation}
\left(\frac{1}{\mathcal{H}} \right) '=-\left(1- \epsilon\right) \left(1+\frac{1}{\mathcal{H}^2} \right),
\end{equation}
which results in
\begin{equation}\label{18}
\mathcal{H} = - \cot \Big[\left( 1 - \epsilon \right)\tau - \cot ^{ - 1}n\Big].
\end{equation}
Here we assumed $ \tau=\tau_n=-\int_t^{t_n}\frac{d\eta}{a\left(\eta\right)} $ where $ t_n $ is the horizon exit time for the inhomogeneity mode $ n $ ($ n=\mathcal{H}\left( t_n\right) $). Furthermore, combination of equations (\ref{17}) and (\ref{18}) results in
\begin{equation}\label{19}
\frac{\chi '}{\chi}=-2\left( 1+\delta \right) \cot \Theta ,
\end{equation}
where $ \Theta=\left(1 - \epsilon \right)\tau - \cot ^{ - 1}n $.
Now by substituting of equations (\ref{18}) and (\ref{19}) in equation (\ref{11}) we can deduce
\begin{align}
&\Bigg[\left( n^2 - 4 \right) + \frac{\epsilon }{\cos ^2\Theta} \Bigg]\mathcal{R}''_\mathbf{n} - \Bigg[ 4\left( n^2 - 4\right)\cot 2 \Theta + 4\epsilon \frac{n^2 - 3}{\sin 2 \Theta } + 2\delta\left(n^2 - 4\right) \cot \Theta \Bigg]\mathcal{R}'_\mathbf{n}+\nonumber\\
&\quad \Bigg[\left(n^2 - 4\right)\left(n^2 - 5\right) + 2\left(n^2 - 4 \right)\tan ^2 \Theta - \epsilon \frac{n^2 - 3}{\cos ^2 \Theta} - 2\delta \left( n^2 - 4\right) \Bigg]\mathcal{R}_\mathbf{n} = 0 .\label{20}
\end{align}
Hereafter, we just investigate linear perturbations i.e. the terms such as $ \epsilon^2,\delta^2,\epsilon\delta , $ etc. shall be ignored. Now let’s define the new variable $ \mathcal{V}_\mathbf{n} $ as
\begin{equation}\label{21}
\mathcal{V}_\mathbf{n} =\mathscr{T}\mathcal{R}_\mathbf{n}, \quad \mathscr{T}=\mathscr{C}\frac{\exp\left[ - \frac{\epsilon }{2\left(n^2 - 4 \right)\cos ^2 \Theta }\right]} {\left| \sin\Theta\right|^{1 + 2\epsilon + \delta }\left|\cos \Theta \right|}.
\end{equation}
(Here $ \mathscr{C} $ is a constant which is obtained soon) Thus, equation (\ref{20}) may be written in terms of $ \mathcal{V}_\mathbf{n} $
\begin{multline}\label{22}
\mathcal{V}''_\mathbf{n} + \Bigg[ \left( n^2 - 5 \right) - 2\cot ^2\Theta + \epsilon \left(1 + \cot ^2 \Theta\right)\left( 2\frac{1 - \cot ^2 \Theta}{\cot ^2 \Theta} + \frac{1}{n^2 - 4}\frac{3 - \cot ^2 \Theta}{\cot ^4\Theta} \right)\\ - \delta \left(1 + 3\cot ^2 \Theta \right) \Bigg]\mathcal{V}_\mathbf{n} = 0 .
\end{multline}
Before solving equation (\ref{22}), let’s find constant $ \mathscr{C} $. For this purpose, we may invoke the relation
\begin{equation}\label{23}
\frac{\epsilon '}{\epsilon}=2\mathcal{H}\left(\epsilon +\delta \right),
\end{equation}
which can be derived from the logarithmic derivation of equation (\ref{16}). Provided that $ \epsilon $ and $ \delta $ are supposed to be constant, equation (\ref{23}) yields
\begin{equation}\label{24}
\left(\frac{a}{\mathfrak{R}} \right) ^{\epsilon +\delta}=\sqrt{\epsilon} .
\end{equation}
Here $ \mathfrak{R} $ is a characteristic scale which is appeared as the integral constant in equation (\ref{24}).\\
On the other hand, equation (\ref{18}) results in
\begin{equation}\label{25}
a=\frac{1}{\mathfrak{H}}\left|\sin\Theta\right| ^{-\left(\epsilon +\delta\right) }, \quad \mathfrak{H}=\sqrt{\frac{8\pi G}{3}\bar{\rho}} ,
\end{equation}
Consequently
\begin{equation}\label{26}
\left|\sin\Theta\right| ^{\epsilon +\delta}=\frac{\left(\mathfrak{RH}\right) ^{-\left(\epsilon +\delta\right) }}{\sqrt{\epsilon}} .
\end{equation}
Meanwhile, using equations (\ref{18}), (\ref{21}) as well as (\ref{16}) and (\ref{17}) one can show
\begin{equation}\label{27}
\frac{\mathscr{T}'}{\mathscr{T}}=\frac{a'}{a}-\frac{\mathcal{H}'}{\mathcal{H}}+\frac{\bar{\phi}''}{\bar{\phi}'}+\frac{1}{n^2-4}\left(\frac{1}{\mathcal{H}}-\frac{\mathcal{H}'}{\mathcal{H}^3}+\frac{1}{\mathcal{H}^3} \right) ,
\end{equation}
which results in
\begin{equation}\label{28}
\mathscr{T}=\frac{a\bar{\phi}'}{\mathcal{H}}\exp\left[ \frac{1}{n^2-4}\left(\frac{1}{2\mathcal{H}^2}+\int\frac{\mathcal{H}^2+1} {\mathcal{H}^3} d\tau\right)\right] .
\end{equation}
Now let’s suppose $ n\longrightarrow+\infty $, thus equation (\ref{28}) takes the form
\begin{equation}\label{29}
\mathop {\lim }\limits_{n \to +\infty }\mathscr{T}=\frac{\mathscr{C}} {\left| \sin\Theta\right|^{1 + 2\epsilon + \delta }\left|\cos \Theta \right|}=\frac{a\bar{\phi}'}{\mathcal{H}}=\mathscr{Z} .
\end{equation}
For the severe sub-Hubble modes, the curvature has a negligible imprint and may be disregarded, so it coincides with $ K=0 $ case. Besides, it can be shown that
\begin{equation}\label{30}
\frac{a\bar{\phi}'}{\mathcal{H}}=\frac{a}{\mathcal{H}}\sqrt{\frac{\mathcal{H}^2-\mathcal{H}'+1}{4\pi G}}=a\sqrt{\left( 1+\frac{1}{\mathcal{H}^2}\right)\frac{\epsilon}{4\pi G} }=\frac{1}{\mathfrak{H}}\left|\sin\Theta\right|^{-1-\epsilon}\left|\cos\Theta\right|^{-1}\sqrt{\frac{\epsilon}{4\pi G}},
\end{equation}
Therefore, combination of equations (\ref{30}), (\ref{29}) and (\ref{26}) yields
\begin{equation}
\mathscr{C}=\frac{1}{\sqrt{4\pi G}}\frac{1}{\mathfrak{H}\left( \mathfrak{RH}\right) ^{\epsilon +\delta}} ,
\end{equation}
So
\begin{equation}\label{31}
\mathcal{R}_\mathbf{n} =\sqrt{4\pi G}\mathfrak{H}\left( \mathfrak{RH}\right) ^{\epsilon +\delta}\left| \sin\Theta\right|^{1 + 2\epsilon + \delta }\left|\cos \Theta \right| \exp\left[\frac{\epsilon }{2\left(n^2 - 4 \right)\cos ^2 \Theta }\right]\mathcal{V}_\mathbf{n} .
\end{equation}
Notice that $ \mathcal{V}_\mathbf{n} $ is the \textit{generalized Sasaki-Mukhanov variable} for the inflationary universe with positive curvature index.
\subsection{Curvature power spectrum }
\label{sec:2.3}
Now let’s find the solutions of the equation (\ref{22}). For this purpose, we assume $ x=\cos\Theta $, thus equation (\ref{22}) reduces to
\begin{multline}\label{32}
\left(1-x^2 \right)\frac{d^2\mathcal{V}_\mathbf{n}}{dx^2}-x\frac{d\mathcal{V}_\mathbf{n}}{dx}+\\
\left[ \left(n^2-3 \right)\left(1+2\epsilon \right)+2\delta-\frac{2+6\epsilon +3\delta}{1-x^2}+2\epsilon\left(1-\frac{2}{n^2-4}\right) \frac{1}{x^2}+\frac{3\epsilon}{n^2-4}\frac{1}{x^4} \right] \mathcal{V}_\mathbf{n}=0 .
\end{multline}
It may be proposed the solution as
\begin{equation}\label{33}
\mathcal{V}_\mathbf{n}=\EuScript{V}_\mathbf{n}+\epsilon \mathfrak{V}_\mathbf{n} ,
\end{equation}
where
\begin{equation}\label{34}
\EuScript{V}_\mathbf{n}=\mathscr{A}\sqrt[4]{1-x^2} P^\mu_\nu\left(x \right) +\mathscr{B}\sqrt[4]{1-x^2} Q^\mu_\nu\left(x \right),
\end{equation}
\begin{equation}\label{35}
\left\{
\begin{aligned}
&\mu:=\frac{3}{2}+2\epsilon+\delta ,\\
&\nu:=\left(1+\epsilon \right)\sqrt{n^2-3}+\frac{\delta}{\sqrt{n^2-3}} -\frac{1}{2}.
\end{aligned}
\right.
\end{equation}
Notice that $ P^\mu_\nu $ and $ Q^\mu_\nu $ are associated Legendre functions. By inserting ansatz (\ref{33}) in equation (\ref{32}) and neglecting higher order infinitesimal terms, one obtains a second order nonhomogeneous equation in terms of $ \mathfrak{V}_\mathbf{n} $
\begin{multline}\label{36}
\left(1-x^2 \right)\frac{d^2\mathfrak{V}_\mathbf{n}}{dx^2}-x\frac{d\mathfrak{V}_\mathbf{n}}{dx}+\left[ \left(n^2-3 \right)\left(1+2\epsilon \right)+2\delta-\frac{2+6\epsilon+3\delta}{1-x^2}\right] \mathfrak{V}_\mathbf{n}=\\
-\frac{1}{n^2-4}\left(2\frac{n^2-6}{x^2}+\frac{3}{x^4} \right) \left[ \mathscr{A}\left(1-x^2 \right)^{\frac{1}{4}} P^\mu_\nu\left(x \right) +\mathscr{B}\left(1-x^2 \right)^{\frac{1}{4}} Q^\mu_\nu\left(x \right)\right] .
\end{multline}
which has the special solution as
\begin{multline}\label{37}
\mathfrak{V}_\mathbf{n}=\frac{1}{n^2-4}\frac{\Gamma\left(\nu-\mu+1 \right) }{\Gamma\left(\nu+\mu+1 \right)}\sqrt[4]{1-x^2} \Bigg\{ \Big[ \mathscr{A}P^\mu_\nu\left(x \right) -\mathscr{B}Q^\mu_\nu\left(x \right) \Big] \int_{x_0}^x \left(1-y^2 \right)\left(2\frac{n^2-6}{y^2}+\frac{3}{y^4} \right)\times \\
P^\mu_\nu\left(y \right)Q^\mu_\nu\left(y \right)dy - \mathscr{A}Q^\mu_\nu\left(x \right)\int_{x_0}^x \left(1-y^2 \right)\left(2\frac{n^2-6}{y^2}+\frac{3}{y^4} \right)\Big[ P^\mu_\nu\left(y \right)\Big]^2dy \\ +\mathscr{B}P^\mu_\nu\left(x \right)\int_{x_0}^x \left(1-y^2 \right)\left(2\frac{n^2-6}{y^2}+\frac{3}{y^4} \right)\Big[ Q^\mu_\nu\left(y \right)\Big] ^2 dy \Bigg\} .
\end{multline}
Here, $ x_0 $ is an arbitrary constant for which $ \left| x_0\right|\leqslant 1 $. Consequently, the general solution of equation (\ref{22}) reduces to
\begin{align}\label{38}
&\mathcal{V}_\mathbf{n}\left(\tau \right) =\sqrt{\left|\sin\Theta\right|}\Big[\mathscr{A} P^\mu_\nu\left(\cos\Theta \right) +\mathscr{B} Q^\mu_\nu\left(\cos\Theta \right)\Big] + \frac{\epsilon}{n^2-4}\frac{\Gamma\left(\nu-\mu+1 \right) }{\Gamma\left(\nu+\mu+1 \right)}\sqrt{\left|\sin\Theta\right|}\times\nonumber\\
&\Bigg\{\Big[-\mathscr{A}P^\mu_\nu\left(\cos\Theta \right) +\mathscr{B}Q^\mu_\nu\left(\cos\Theta \right) \Big]\int_{\Theta_0}^\Theta\sin ^3\Upsilon\left(2\frac{n^2-6}{\cos ^2 \Upsilon}+\frac{3}{\cos^4\Upsilon} \right)P^\mu_\nu\left(\cos\Upsilon \right)\times\nonumber\\
&Q^\mu_\nu\left(\cos\Upsilon \right)d\Upsilon+\mathscr{A}Q^\mu_\nu\left(\cos\Theta \right)\int_{\Theta_0}^\Theta \sin^3 \Upsilon\left(2\frac{n^2-6}{\cos^2\Upsilon}+\frac{3}{\cos^4\Upsilon} \right)\Big[ P^\mu_\nu\left(\cos\Upsilon \right)\Big]^2d\Upsilon \nonumber\\
&- \mathscr{B}P^\mu_\nu\left(\cos\Theta \right) \int_{\Theta_0}^\Theta\sin^3 \Upsilon\left(2\frac{n^2-6}{\cos^2\Upsilon}+\frac{3}{\cos^4\Upsilon} \right)\Big[ Q^\mu_\nu\left(\cos\Upsilon \right)\Big] ^2 d\Upsilon \Bigg\} .
\end{align}
Hereafter, we put $ \Theta_0=-\cot^{-1}n $ ( it is completely compatible with the conformal initial condition which is introduced below).\\
In order to determine the constants $ \mathscr{A} $ and $ \mathscr{B} $ we may use the \textit{conformal (Bunch-Davies) initial condition} which states\cite{ref44,ref45}
\begin{equation}\label{39}
\mathop {\lim }\limits_{n \to +\infty }\mathcal{V}_\mathbf{n}=\frac{1}{\sqrt{2n}}\exp\left( -in\tau\right) .
\end{equation}
Thus, according to the asymptotic formulas of $ P^\mu_\nu $ and $ Q^\mu_\nu $ for large value of $ \nu $ i.e. \cite{ref46}
\begin{align}
&P^\mu_\nu\left(\cos\theta \right) \sim\frac{\Gamma\left(\mu+\nu+1 \right) }{\Gamma\left(\nu+\frac{3}{2} \right) }\sqrt{\frac{2}{\pi\sin\theta}}\sin\left[ \left(\nu+\frac{1}{2} \right)\theta+\frac{\pi}{4} +\frac{\mu\pi}{2} \right]+\mathcal{O}\left(\nu^{-1} \right) ,\label{40} \\
&Q^\mu_\nu\left(\cos\theta \right) \sim\frac{\Gamma\left(\mu+\nu+1 \right) }{\Gamma\left(\nu+\frac{3}{2} \right) }\sqrt{\frac{\pi}{2\sin\theta}}\cos\left[ \left(\nu+\frac{1}{2} \right)\theta+\frac{\pi}{4} +\frac{\mu\pi}{2} \right]+\mathcal{O}\left(\nu^{-1} \right) ,\label{41}
\end{align}
and noting that $ \mathop {\lim}\limits_{n \to +\infty }\frac{\left(n+\alpha \right)!}{n!}\sim n^\alpha $, after alot of lengthy but straightforward calculations, it can be shown
\begin{equation}\label{42}
\left\{
\begin{aligned}
&\mathscr{A}=\frac{i\sqrt{\pi}}{2}n^{-\frac{3}{2}-2\epsilon-\delta},\\
&\mathscr{B}=-\frac{1}{\sqrt{\pi}}n^{-\frac{3}{2}-2\epsilon-\delta}.
\end{aligned}
\right.
\end{equation}
Thus
\begin{align}\label{43}
&\mathcal{V}_\mathbf{n}\left(\tau \right) =\sqrt{\left|\sin\Theta\right|}n^{-\mu}\Bigg\{\frac{i\sqrt{\pi}}{2} P^\mu_\nu\left(\cos\Theta \right) -\frac{1}{\sqrt{\pi}} Q^\mu_\nu\left(\cos\Theta \right)- \frac{\epsilon}{n^2-4}\frac{\Gamma\left(\nu-\mu+1 \right) }{\Gamma\left(\nu+\mu+1 \right)}\times\nonumber\\
& \Big[\frac{i\sqrt{\pi}}{2}P^\mu_\nu\left(\cos\Theta \right) +\frac{1}{\sqrt{\pi}}Q^\mu_\nu\left(\cos\Theta \right) \Big]\int_0^{\left( 1-\epsilon \right) \tau}\sin ^3\Upsilon\left(2\frac{n^2-6}{\cos ^2 \Upsilon}+\frac{3}{\cos^4\Upsilon} \right)P^\mu_\nu\left(\cos\Upsilon \right)\times\nonumber\\
& Q^\mu_\nu\left(\cos\Upsilon \right)d\eta+\frac{i\sqrt{\pi}}{2}\frac{\epsilon}{n^2-4}\frac{\Gamma\left(\nu-\mu+1 \right) }{\Gamma\left(\nu+\mu+1 \right)}Q^\mu_\nu\left(\cos\Theta \right)\int_0^{\left( 1-\epsilon \right) \tau} \sin^3 \Upsilon\left(2\frac{n^2-6}{\cos^2\Upsilon}+\frac{3}{\cos^4\Upsilon} \right)\times\nonumber\\
&\Big[ P^\mu_\nu\left(\cos\Upsilon \right)\Big]^2d\eta +\frac{1}{\sqrt{\pi}}\frac{\epsilon}{n^2-4}\frac{\Gamma\left(\nu-\mu+1 \right) }{\Gamma\left(\nu+\mu+1 \right)}P^\mu_\nu\left(\cos\Theta \right) \int_0^{\left( 1-\epsilon \right) \tau}\sin^3 \Upsilon\left(2\frac{n^2-6}{\cos^2\Upsilon}+\frac{3}{\cos^4\Upsilon} \right)\times \nonumber\\
&\Big[ Q^\mu_\nu\left(\cos\Upsilon \right)\Big] ^2 d\eta \Bigg\} .
\end{align}
Furthermore, by ignoring the non-linear terms, $ \mathcal{R}_\mathbf{n} $ takes the form
\begin{align}\label{44}
&\mathcal{R}_\mathbf{n}\left(\tau \right) =\sqrt{4\pi G}\mathfrak{H}\left( \mathfrak{RH}\right) ^{\epsilon +\delta}\left|\frac{\sin \Xi}{n} \right|^\mu\left|\cos \Xi \right| \Big[ \frac{i\sqrt{\pi}}{2} P^\mu_\nu\left(\cos \Xi\right)-\frac{1}{\sqrt{\pi}} Q^\mu_\nu\left(\cos \Xi \right)\Big]\nonumber\\
&+\epsilon \sqrt{4\pi G}\mathfrak{H}\left( \mathfrak{RH}\right) ^{\epsilon +\delta}\left|\frac{\sin \Xi}{n} \right|^\mu\left|\cos \Xi \right| \Bigg\{ -\frac{i\sqrt{\pi}}{2}\tau\frac{dP^\mu_\nu\left(\cos\Xi \right)}{d\tau}+\frac{1}{\sqrt{\pi}}\tau\frac{dQ^\mu_\nu\left(\cos\Xi \right)}{d\tau}\nonumber\\
&-\Big[2\tau\cot 2\Xi +\frac{1}{2}\tau\cot \Xi-\frac{1}{2\left(n^2-4 \right)\cos ^2\Xi } \Big] \Big[ \frac{i\sqrt{\pi}}{2}P^\mu_\nu\left( \cos\Xi\right) -\frac{1}{\sqrt{\pi}}Q^\mu _\nu\left( \cos\Xi\right) \Big]\nonumber\\
&-\frac{1}{n^2-4}\frac{\Gamma\left(\nu-\mu+1 \right) }{\Gamma\left(\nu+\mu+1 \right)}\Big[\frac{i\sqrt{\pi}}{2}P^\mu_\nu\left(\cos\Xi \right) +\frac{1}{\sqrt{\pi}}Q^\mu_\nu\left(\cos\Xi \right) \Big]\nonumber\\
&\times \int_0^\tau\sin ^3\Upsilon\left(2\frac{n^2-6}{\cos ^2 \Upsilon}+\frac{3}{\cos^4\Upsilon} \right)P^\mu_\nu\left(\cos\Upsilon \right)Q^\mu_\nu\left(\cos\Upsilon \right)d\eta\nonumber\\
&+\frac{i\sqrt{\pi}}{2}\frac{1}{n^2-4}\frac{\Gamma\left(\nu-\mu+1 \right) }{\Gamma\left(\nu+\mu+1 \right)}Q^\mu_\nu\left(\cos\Xi \right)\int_0^\tau \sin^3 \Upsilon\left(2\frac{n^2-6}{\cos^2\Upsilon}+\frac{3}{\cos^4\Upsilon} \right)\Big[ P^\mu_\nu\left(\cos\Upsilon \right)\Big]^2d\eta \nonumber\\
&+\frac{1}{\sqrt{\pi}}\frac{1}{n^2-4}\frac{\Gamma\left(\nu-\mu+1 \right) }{\Gamma\left(\nu+\mu+1 \right)}P^\mu_\nu\left(\cos\Xi \right) \int_0^\tau\sin^3 \Upsilon\left(2\frac{n^2-6}{\cos^2\Upsilon}+\frac{3}{\cos^4\Upsilon} \right)\Big[ Q^\mu_\nu\left(\cos\Upsilon \right)\Big] ^2 d\eta \Bigg\} ,
\end{align}
where $ \Xi= \tau-\cot^{-1}n $ and $ \Upsilon=\eta-\cot^{-1}n $.\\
It is important to evaluate the comoving curvature perturbation at the horizon exit time $ \tau =0 $ i.e. when the quantum fluctuations of inflaton came to be classical perturbations. Besides, by inserting $ \tau =0 $ in equation (\ref{44}) the arguments of $ P^\mu_\nu $ and $ Q^\mu_\nu $ become $ \cos\left(\cot^{-1}n \right)=\frac{n}{\sqrt{n^2+1}} $ which for $ n\geq 3 $ , $ 0.94 \leq \frac{n}{\sqrt{n^2+1}}<1 $, so it may be plausible to use asymptotic formulas of the associated Legendre functions near one i.e. \cite{ref47}
\begin{align}
&\theta\longrightarrow 0\quad :\quad P^\mu_\nu\left(\cos\theta \right) \sim\frac{1}{\pi}\Gamma\left(\mu\right) \sin\mu\pi \left( \frac{2}{1-\cos\theta}\right)^{\frac{\mu}{2}} ,\label{45} \\
&\theta\longrightarrow 0\quad :\quad Q^\mu_\nu\left(\cos\theta \right) \sim\frac{1}{2}\Gamma\left(\mu\right) \cos\mu\pi \left( \frac{2}{1-\cos\theta}\right)^{\frac{\mu}{2}} .\label{46}
\end{align}
So by doing some straightforward calculation, it can be shown
\begin{multline}\label{47}
\mathcal{R}^o_\mathbf{n}=-\sqrt{G} \mathfrak{H}\left( \mathfrak{RH}\right) ^{\epsilon +\delta} \Gamma\left(\mu\right) \exp\left(-i\mu \pi \right)\frac{n^{1-\mu}}{\sqrt{n^2+1}}\left(2+\frac{2n}{\sqrt{n^2+1}} \right) ^{\frac{\mu}{2}}\\
\times\left(1+\frac{n^2+1}{2n^2\left(n^2-4 \right) }\epsilon\right) .
\end{multline}
Let’s approximate $ \frac{n}{\sqrt{n^2+1}}\sim 1 $, thus equation (\ref{47}) takes the form
\begin{equation}\label{48}
\mathcal{R}^o_\mathbf{n}\simeq-\sqrt{G} \mathfrak{H}\left(\mathfrak{RH}\right) ^{\epsilon +\delta}2^{\frac{3}{2}+2\epsilon +\delta} \Gamma\left(\mu\right) \exp\left(-i\mu\pi \right)n^{-\mu}\left(1+\frac{\epsilon}{2\left(n^2-4 \right) }\right) .
\end{equation}
Consequently, the curvature power spectrum in the maximally extended inflationary universe with single field reduces to
\begin{equation}\label{49}
\mathcal{P}^o_{\mathcal{R}}\left(n\right) \propto n^{-3-4\epsilon-2\delta}\left( 1+\frac{\epsilon}{n^2-4}\right) .
\end{equation}
Except the additional factor $ 1+\frac{\epsilon}{n^2-4} $, spectrum (\ref{49}) is similar to the nearly flat spectrum which can be deduced from the slow-rolling inflationary scenario with spatially flat background\cite{ref13}. By definition of the \textit{curvature spectral index} as
\begin{equation}
\mathcal{P}^o_{\mathcal{R}}\left(n\right) \propto n^{\mathfrak{N}_s\left(n\right)-4} ,
\end{equation}
one can show that
\begin{equation}\label{50}
\mathfrak{N}_s\left(n\right)=1 - 4\epsilon - 2\delta + \frac{2\epsilon}{\left( {{n^2} - 4} \right)\ln n} .
\end{equation}
Because $ n\geq 3 $ so
\begin{equation}
1 - 4\epsilon - 2\delta <\mathfrak{N}_s\left(n\right)\lesssim 1-3.64\epsilon-2\delta .
\end{equation}
It means the curvature spectral index in the maximally extended universe shall be a bit larger than the $ K=0 $ corresponding model (For the $ K=0 $ case, $ \mathfrak{N}_s\left(n\right)=1 - 4\epsilon - 2\delta $). Moreover, $ \mathfrak{N}_s $ directly depends on the comoving wave number ($ n $) and so the spectrum is running. In other words, \textit{running parameter} of $ \mathfrak{N}_s $ doesn’t vanish
\begin{equation}
\mathfrak{N}_r\left( n \right) = n\frac{\partial \mathfrak{N}_s}{\partial n} =- 2\epsilon \frac{\left(n^2 - 4\right) + 2n^2\ln n}{\left( n^2 - 4\right)^2\ln ^2n} < 0 .
\end{equation}
It is remarkable that the sign of $ \mathfrak{N}_r $ coincides with experimental data. Moreover, running parameter in the maximally extended background inflationary model is proportional to $ \epsilon $ i.e. $ \mathfrak{N}_r $ is of the first order slow-roll parameters accordant with the reports \cite{ref11}, in spite of the spatially flat case in which against the Planck reports is roughly zero.
\section{Primordial gravitational waves power spectrum in the positively curved universe}
\label{sec:3}
The primordial gravitational waves during inflationary epoch can be treated in the same way as the comoving curvature perturbation considered in previous section. In fact, quantum fluctuations of the inflaton may result in tensorial perturbations described by a symmetric traceless divergenceless tensor field $ D_{ij}\left(t,\bf{x} \right) $ which perturbs the FLRW metric as\cite{ref13}
\begin{equation}
ds^2=-dt^2+a^2\left(\tilde{g}_{ij}+D_{ij}\right)dx^idx^j .
\end{equation}
Propagation of $ D_{ij} $ in the positively curved FLRW universe is described by\cite{ref41}
\begin{equation}\label{51}
a^2\nabla ^2D_{ij}-3a\dot{a}\dot{D}_{ij}-a^2\ddot{D}_{ij}-2D_{ij}=-16\pi Ga^2\Pi ^T_{ij} .
\end{equation}
Here $ \Pi ^T_{ij}\left(t,\bf{x} \right) $ is the anisotropic inertia tensor which vanishes for the scalar fields, so
\begin{equation}\label{52}
a^2\nabla ^2D_{ij}-3a\dot{a}\dot{D}_{ij}-a^2\ddot{D}_{ij}-2D_{ij}=0 .
\end{equation}
One may expand $ D_{ij} $ in terms of the t-t tensor spherical harmonics on $ \mathbb{S}^3\left(a \right) $ \cite{ref41}
\begin{equation}\label{53}
D_{ij}\left(t ,\mathbf{x} \right) = \sum\limits_{nlm} \left[ \mathcal{D}_{nlm}^{\mathbb{O}}\left(t \right)\big( T_{ij}^{\mathbb{O}} \big)_{nlm}+ \mathcal{D}_{nlm}^{\mathbb{E}}\left( t \right)\big( T_{ij}^{\mathbb{E}} \big)_{nlm} \right] ,
\end{equation}
where $ \mathcal{D}_\mathbf{n}^\mathbb{O} $ and $ \mathcal{D}_\mathbf{n}^\mathbb{E} $ correspond to two different polarizations of the gravitational waves. Notice that $ \left\lbrace \big( T_{ij}^{\mathbb{O}} \big)_{nlm} ,\big( T_{ij}^{\mathbb{E}} \big)_{nlm} \right\rbrace $ constitutes a complete orthonormal basis for the expansion of any symmetric traceless divergence-free covariant tensor field of rank 2 on $ \mathbb{S}^3\left(a \right) $. Furthermore \cite{ref41},
\begin{align}
& \nabla ^2 \big( T_{ij}^\mathbb{O}\big)_{nlm} = \frac{3 - n^2}{a^2 }\big(T_{ij}^\mathbb{O}\big)_{nlm}, \quad n = 3,4,... \\ &
\nabla ^2 \big(T_{ij}^\mathbb{E} \big)_{nlm} = \frac{3 - n^2 }{a^2}\big(T_{ij}^\mathbb{E} \big)_{nlm}.\quad n = 3,4,...
\end{align}
Thus equation (\ref{53}) reduces to two independent equations
\begin{equation}
\left\{\begin{aligned}
&\ddot{\mathcal{D}}_\mathbf{n}^\mathbb{O}\left(t \right) + 3H\dot{\mathcal{D}}_\mathbf{n}^\mathbb{O}\left(t \right) +\frac{n^2 - 1}{a^2} \mathcal{D}_\mathbf{n}^\mathbb{O}\left(t \right) = 0 ,\\
&\ddot{\mathcal{D}}_\mathbf{n}^\mathbb{E}\left(t \right) + 3H\dot{\mathcal{D}}_\mathbf{n}^\mathbb{E}\left(t \right) +\frac{n^2 - 1}{a^2} \mathcal{D}_\mathbf{n}^\mathbb{E}\left(t \right) = 0 .
\end{aligned}\right.
\end{equation}
Hereafter we omit the superscripts $ \mathbb{O} $ and $ \mathbb{E} $ because both of $ \mathcal{D}_\mathbf{n}^\mathbb{O} $ and $ \mathcal{D}_\mathbf{n}^\mathbb{E} $ satisfy the same equation
\begin{equation}\label{54}
\ddot{\mathcal{D}}_\mathbf{n}\left(t \right) + 3H\dot{\mathcal{D}}_\mathbf{n}\left(t \right) +\frac{n^2 - 1}{a^2} \mathcal{D}_\mathbf{n}\left(t \right) = 0 .
\end{equation}
$ \mathcal{D}_\mathbf{n}\left(t \right) $ is amplitude of the gravitational wave $ D_{ij}\left(t, \bf{x} \right) $ as well as a tensor random field on $ \mathbb{S}^3\left(a \right) $. By converting the cosmic time to the conformal time equation (\ref{54}) takes the form
\begin{equation}\label{55}
\mathcal{D}''_\mathbf{n}\left(\tau \right) + 2\mathcal{H}\mathcal{D}'_\mathbf{n}\left(\tau \right) +\left( n^2 - 1\right) \mathcal{D}_\mathbf{n}\left(\tau \right) = 0 .
\end{equation}
During the slow-rolling inflationary epoch we can write
\begin{equation}\label{56}
\mathcal{D}''_\mathbf{n}\left(\tau \right) - 2\cot\Theta\mathcal{D}'_\mathbf{n}\left(\tau \right) +\left( n^2 - 1\right) \mathcal{D}_\mathbf{n}\left(\tau \right) = 0 .
\end{equation}
By assumption $ x=\cos\Theta $ equation (\ref{56}) can be written as
\begin{equation}
\left(1-x^2 \right)\frac{d^2\mathcal{D}_\mathbf{n}}{dx^2}+\left(1+2\epsilon \right)x\frac{d\mathcal{D}_\mathbf{n}}{dx}+\left( 1+2\epsilon\right) \left( n^2-1\right) \mathcal{D}_\mathbf{n}\left(x \right)=0 ,
\end{equation}
which has the general solution as
\begin{equation}
\mathcal{D}_\mathbf{n}\left(x \right)=\left(1-x^2 \right) ^{\frac{2\epsilon +3}{4}}\left[\mathscr{P}P^{\epsilon +\frac{3}{2}}_{n\left( 1+\epsilon\right) -\frac{1}{2}}\left(x \right)+ \mathscr{Q}Q^{\epsilon +\frac{3}{2}}_{n\left( 1+\epsilon\right) -\frac{1}{2}}\left(x \right) \right] .
\end{equation}
Here $ \mathscr{P} $ and $ \mathscr{Q} $ are two arbitrary constants. So the solution of equation (\ref{56}) is
\begin{equation}\label{57}
\mathcal{D}_\mathbf{n}\left(\tau \right)=\left|\sin\Theta \right| ^\iota\Big[\mathscr{P}P^\iota_\kappa\left(\cos\Theta \right)+ \mathscr{Q}Q^\iota_\kappa\left(\cos\Theta \right) \Big] ,
\end{equation}
where
\begin{equation}
\left\{
\begin{aligned}
&\iota :=\epsilon +\frac{3}{2} ,\\
&\kappa :=n\left( 1+\epsilon\right) -\frac{1}{2} .
\end{aligned}
\right.
\end{equation}
Besides, the initial condition must be satisfied by $ \mathcal{D}_\mathbf{n} $ is very similar to the Bunch-Davies initial condition exerted to the Sasaki-Mukhanov variable\cite{ref13}
\begin{equation}\label{58}
\mathop {\lim }\limits_{n \to +\infty }\mathcal{D}_\mathbf{n}=\frac{\sqrt{16\pi G}}{a\left(t \right) }\frac{1}{\sqrt{2n}}\exp\left( -in\tau\right) ,
\end{equation}
which is applicable for both polarization modes distinctly. By considering the asymptotic formulas (\ref{40}) and (\ref{41}) and equation (\ref{25}) as well, one can obtain
\begin{equation}
\left\{
\begin{aligned}
&\mathscr{P}=2\pi i \mathfrak{H}\sqrt{G}n^{-\iota} ,\\
&\mathscr{Q}=-4\mathfrak{H}\sqrt{G}n^{-\iota} .
\end{aligned}
\right.
\end{equation}
Thus
\begin{equation}\label{59}
\mathcal{D}_\mathbf{n}\left( \tau \right) = 2\sqrt{G} \mathfrak{H}\left| \frac{\sin \Theta }{n}\right|^\iota \Big[\pi i P_\kappa ^\iota \left(\cos \Theta\right) - 2Q_\kappa ^\iota \left(\cos \Theta\right)\Big] .
\end{equation}
$ \mathcal{D}_\mathbf{n}^o $ may be determined by considering $ \mathcal{D}_\mathbf{n} $ at the time of horizon crossing ($ \tau=0 $)
\begin{equation}\label{60}
\mathcal{D}_\mathbf{n}^o=-2\sqrt{G}\mathfrak{H}\Gamma\left( \iota\right) \exp\left(-i \iota \pi \right) n^{-\iota}\left(2+\frac{2n}{\sqrt{n^2+1}} \right) ^{\frac{\iota}{2}} .
\end{equation}
Here again we use the asymptotic relations (\ref{45}) and (\ref{46}). By approximation $ \frac{n}{\sqrt{n^2+1}}\sim 1 $, equation (\ref{60}) acquires a simpler form which results in
\begin{equation}
\mathcal{P}^o_\mathcal{D}\left(n \right) \propto n^{-3-2\epsilon},
\end{equation}
So by definition of the \textit{tensor spectral index} as
\begin{equation}
\mathcal{P}^o_\mathcal{D}\propto n^{\mathfrak{N}_T-3},
\end{equation}
one can obtain
\begin{equation}
\mathfrak{N}_T=-2\epsilon ,
\end{equation}
which is perfectly analogous to tensor spectral index derived in the classical slow-rolling inflationary theory \cite{ref13}.
\section{Tensor-scalar ratio in the positively curved universe}
\label{sec:4}
\textit{Tensor-scalar ratio} in the positively curved FLRW universe may be defined as \cite{ref13}
\begin{equation}
r_n:=4\frac{\mathcal{P}_\mathcal{D}^o\left( n \right)}{\mathcal{P}_\mathcal{R}^o\left( n \right)} = 4\left|\frac{\mathcal{D}_\mathbf{n}^o}{\mathcal{R}_\mathbf{n}^o} \right| ^2 .
\end{equation}
Here the factor 4 refers to two different polarization modes of the gravitations waves. Significance of $ r_n $ come from its measurability, indeed tensor-scalar ratio can provide an assay for the inflationary scenarios and some inflation theories may be crossed out due to the contradiction with observational value of $ r_n $. According to the standard slow-rolling inflationary theory $ r_q=16\epsilon $ ($ q $ stands for the comoving wave number of perturbations in the spatially flat universe) \cite{ref13}, so if one suppose $ \epsilon=0.008 $ \cite{ref11} then $ r=0.128 $ which is less than BICEP2 released data ($ r=0.20_{-0.05}^{+0.07} $) \cite{ref22} and now a question dawns on the mind: “ Is it possible to eliminate this flaw by considering curvature factor?” In order to answer, lets calculate $ r_n $ using equation (\ref{47}) and (\ref{60}) which results in
\begin{multline}\label{61}
r_n=16\left(\mathfrak{RH} \right) ^{-2\left(\epsilon+\delta \right) }\left[\frac{\Gamma\left(\epsilon +\frac{3}{2} \right) }{\Gamma\left( 2\epsilon +\delta +\frac{3}{2}\right) } \right]^2n^{2\left( \epsilon +\delta\right) }\left(2+\frac{2n}{\sqrt{n^2+1}} \right) ^{-\left(\epsilon +\delta \right) }\\
\times\left(1+\frac{1}{n^2} \right) \left(1-\frac{n^2+1}{n^2\left(n^2-4\right) }\epsilon \right) .
\end{multline}
Besides, one can write
\begin{multline}\label{62}
n^{2\left( \epsilon +\delta\right) }=\left( \mathcal{H}^2\vert_{\tau=0}\right) ^{\epsilon +\delta}=\left( \cot^2\Theta\vert_{\tau=0}\right) ^{\epsilon +\delta}=\left( \cos^2\Theta\vert_{\tau=0}\right) ^{\epsilon +\delta}\left( \sin^2\Theta\vert_{\tau=0}\right) ^{-\left( \epsilon +\delta\right) }\\
=\left(1+ \frac{1}{n^2}\right) ^{-\left( \epsilon +\delta\right) }\left(\mathfrak{RH} \right) ^{2\left(\epsilon+\delta \right) }\epsilon .
\end{multline}
On the other hand, it isn’t hard to show that
\begin{equation}\label{63}
\frac{\Gamma\left( 2\epsilon +\delta +\frac{3}{2}\right) } {\Gamma\left(\epsilon +\frac{3}{2} \right)}= 1+\left(2-\gamma-\ln 2 \right) \left(\epsilon +\delta \right) \simeq \exp \left[\left(2-\gamma-\ln 2 \right) \left(\epsilon +\delta \right) \right] \simeq \left(2.074 \right) ^{\epsilon +\delta} ,
\end{equation}
where $ \gamma\simeq 0.577 $ is the Euler-Mascheroni constant. In order to derive equation (\ref{63}) one can use the following relation\cite{ref36}
\begin{equation}
\frac{\Gamma\left(x+\epsilon +1 \right) }{\Gamma\left(x+1 \right)}=1+\epsilon \left[ -\gamma +\sum _{n=1}^\infty \left(\frac{1}{n}-\frac{1}{x+n} \right) \right] .
\end{equation}
By inserting equations (\ref{62}) and (\ref{63}) in equation (\ref{61}) we can obtain
\begin{multline}\label{64}
r_n=16\epsilon\,e^{\left(-4+2\gamma +2\ln 2 \right) \left(\epsilon +\delta \right)} \left(1+\frac{1}{n^2} \right)^{1-\left( \epsilon +\delta\right) }\left( 2+\frac{2n}{\sqrt{n^2+1}}\right) ^{-\left( \epsilon +\delta\right) }\\
\times\left(1-\frac{n^2+1}{n^2\left(n^2-4 \right) }\epsilon \right) .
\end{multline}
For $ n\gg 1 $ equation (\ref{64}) reduces to
\begin{equation}\label{65}
r_{n\gg 1} \simeq 16\epsilon\, e^{\left( - 4 + 2\gamma \right)\left(\epsilon + \delta\right)}\simeq 16\epsilon \left( 0.58\right) ^{\epsilon + \delta } .
\end{equation}
By considering $ n_\ast=3 $ as the pivot comoving wave number and $ \epsilon=0.008 $ ,one can show $ r_\ast>16\epsilon $ provided that $ \delta\lesssim 0.03 $ (which is laid in the permitted Planck data\footnote{In the Planck collaboration paper, the slow-roll parameters are $ \epsilon _V $ and $ \eta_V $ which in comparison with our definition $ \epsilon_V=\epsilon $ and $ \eta_V=\epsilon-\delta $.} \cite{ref11}) , so it may reduce discrepancy between BICEP2 results and slow-rolling inflationary theory to some extent, but it isn't statistically significant to eliminate the flaw completely.
\section{Conclusion and summary}
\label{sec:5}
In this article we investigated an inflationary model with positive curvature index and calculated the scalar and tensor perturbations power spectra associated to it. For the severe super-Hubble scales (i.e. $ n\gg 1 $) it seems both spectra are completely similar to the spatially flat corresponding case. It is shown that this model yields a natural resolving of the running number problem. We also calculated the tensor-scalar ratio and showed it depends on the wave number of the perturbative modes directly. Furthermore, we showed that it doesn’t seems to mitigate the discrepancy between BICEP2 released results and anticipation of the slow-rolling inflationary model by entering the curvature factor.
|
\section{Introduction}
The main goal of this paper is to describe the quantum fluctuations of the gravitational field at finite temperature in the linearized gravity. This is done with the use of the Wigner function; a tool that has never been used in this context before. In order to reach this goal I shall introduce the quantization procedure that has the advantage of involving only the geometrical, gauge invariant elements. This is accomplished in the framework of the spinorial formulation providing another evidence for the special role played by spinors and justifying once again the praise of spinors expressed by Andrzej Trautman \cite{amt}: ``One is tempted to say that the world around us, and life in particular, are so rich because Nature found it convenient to use, among its building blocks, entities requiring spinors in their description.''
The original Wigner function is a function of positions and momenta of quantum particles. However, there is natural generalization of this concept. Namely, we may replace the canonically conjugate particle variables by their field-theoretic counterparts. This was done for the scalar field in \cite{mm} and for the electromagnetic field in \cite{ibb}. Upon this generalization, the Wigner function becomes the {\em Wigner functional} whose arguments are the field configurations.
Quantum properties of fields manifest themselves, in particular, in the field fluctuations present even in the vacuum state. In quantum electrodynamics these fluctuations lead to observable effects (Welton explanation of the Lamb shift, Casimir effect, photon shot noise). In classical physics statistical field fluctuations are described by the probability distribution of various field configurations. In quantum physics this simple description fails, due to the uncertainty relations. However, when the Wigner function is positive it may serve as a very good substitute for the standard classical distribution function.
Statistical properties of the vacuum fluctuations of the gravitational field are embodied in the probability distribution that assigns relative weights to different geometries. Exact solution of this problem is a hopeless task since it would require full-fledged quantum theory of gravity. However, an approximate solution can be obtained in linearized gravity---after all, our Universe is mostly flat. In this approach I assume that the linearized gravitational field can be quantized just like any other free field. Every {\em free field} can be viewed as a collection of uncoupled harmonic oscillators. Since the Wigner function for the thermal state of harmonic oscillators is Gaussian, it can serve as a {\em bona fide} probability distribution.
I do realize that linearization is a crude and questionable procedure characterized long time ago by Roger Penrose in highly critical terms \cite{rp} ``\dots if we remove the life from Einstein's beautiful theory by steam-rollering it first to flatness and linearity, then we shall learn nothing from attempting to wave the magic wand of quantum theory over the resulting corpse''. Nevertheless, I did make full use of an important feature of the Penrose approach---the correspondence (fully underscored in the spinorial formulation) between the electromagnetic field and the Riemann tensor.
In a recent paper \cite{fd} Freeman Dyson questioned altogether the widely accepted view that the gravitational field is just another field waiting to be quantized. He raised the possibility that ``the gravitational field is a statistical concept like entropy or temperature, only defined for gravitational effects of matter in bulk and not for effects of individual elementary particles''. If this indeed would be so and the gravitons would not exist, my analysis will loose its foundation. However, if you cannot do what you would like, you should like what you can do.
The analysis of the gravitational Wigner functional is greatly simplified if we take full advantage of the analogy between the electromagnetic field tensor $f_{\mu\nu}$ and the Riemann tensor $R_{\mu\nu\lambda\rho}$ in linearized gravity. Despite this close analogy, however, the differences between the electromagnetic and gravitational cases are substantial. The most striking difference is the divergence of the gravitational Wigner functional in the limit of zero temperature which might signify some problems with the ground state. I will show that the source of these problems is found in the singular behavior of Riemann correlators.
In order to make the analogy between the electromagnetic field and the gravitational field unambiguous we need an anchor point that will enable us to convert the loose analogy into precise mathematical formulas. I choose as this anchor point the universal formula for the energy valid for any free field expressed in terms of annihilation and creation operators. For all massless particles this formula can be written in the form:
\begin{eqnarray}\label{anch}
E=\sum_\lambda\int\!\frac{d^3k}{k}\hbar\omega\,a^\dagger_\lambda({\bm k})a_\lambda({\bm k}),
\end{eqnarray}
where $k=|{\bm k}|$ and $\lambda=\pm$ is the sign of helicity. I have chosen the relativistic normalization of the operators so that their commutation relations have the form:
\begin{eqnarray}\label{cr}
\left[a_\lambda({\bm k}),a_{\lambda'}^\dagger({\bm k'})\right]=\delta_{\lambda\lambda'}\,k\,\delta^{(3)}({\bm k}-{\bm k'}).
\end{eqnarray}
This normalization leads to the relativistically invariant volume element in (\ref{anch}) with $k$ in the denominator.
Quantum statistical properties (Planck spectrum, Einstein A and B coefficients, etc.) are, of course, the same for the gas of photons and (linear) gravitons. Both particles are massless bosons with two helicity states. Large differences appear only in the description in terms of fields localized in space.
The analogy between electromagnetism and gravity will be build on the universal formula (\ref{anch}) but the details will be most clearly seen in the spinorial formalism of relativity theory \cite{pr}. In this formalism only the true degrees of freedom will appear and there is no need for any gauge-fixing conditions.
In the next section I briefly summarize the properties of this formalism needed for my analysis. In Sections \ref{max} and \ref{lin} I describe the quantized electromagnetic and the linearized gravitational field within the spinorial formalism. In Sections \ref{wig1} and \ref{wig2} I evaluate the Wigner functionals for the thermal states of the electromagnetic and gravitational fields and I discuss the physical properties of these states. In the last Section I discuss the problems with the ground state of the quantized gravitational field. In \ref{aa} I prove an important equation that relates the number of particles operators to the spinorial field operators. Finally, in \ref{bb} I use the Bel-Robinson tensor to fix the numerical coefficient in the normalization factor for the gravitational field.
\section{Spinorial formalism}
The most convenient representation of the fields describing massless particles is in terms of {\em symmetric} spinors. The connection between the spinors and the electromagnetic field tensor and the Riemann tensor will be given in Sec. \ref{max} and \ref{lin}, respectively. Spinor indices will be denoted by capital letters and those for conjugate spinors by dotted letters. We shall need the following $2\times 2$ matrices:
\numparts\label{sp}
\begin{eqnarray}
\left\{g^{\mu{\dot A}B}\right\}&=\left\{I,{\bm\sigma}\right\}^{{\dot A}B},\quad\epsilon^{AB}
=\left(\begin{array}{cc}0&1\\-1&0\end{array}\right)=\epsilon_{AB},\\
S^{\mu\nu AB}&=\textstyle{\frac{1}{2}}\left(g^{\mu{\dot C}A}\epsilon_{{\dot C}{\dot D}}g^{\nu{\dot D}B}-
g^{\nu{\dot C}A}\epsilon_{{\dot C}{\dot D}}g^{\mu{\dot D}B}\right),
\end{eqnarray}
\endnumparts
where ${\sigma}$'s are the Pauli matrices. My spinor conventions are those of Ref.~\cite{cors}. They differ slightly from the conventions of Ref.~\cite{pr}. The spinor indices take on the values $(0,1)$ and they are raised and lowered as follows:
\begin{eqnarray}
\phi^A&=\epsilon^{AB}\phi_B,\quad \phi_B=\phi^A\epsilon_{AB}.
\end{eqnarray}
The spinorial wave function $\phi_{AB\dots L}(x)$ obeys the wave equation of the same form for all massless particles \cite{pr},
\begin{eqnarray}\label{fe}
g^{\mu{\dot A}A}\partial_\mu\phi_{AB\dots L}(x)=0.
\end{eqnarray}
The number of indices is equal to $2H$; twice the absolute value of the helicity.
The Fourier representation of the general solution of this equation, \begin{eqnarray}\label{sup}
\fl\qquad{\phi}_{AB\dots L}(x)
=\int\!\frac{d^3k}{(2\pi)^{3/2}k}\kappa_A\kappa_B\dots\kappa_L
\left[f_+({\bm k})e^{-ik\cdot x}+f_-^*({\bm k})e^{ik\cdot x}\right],
\end{eqnarray}
expresses the decomposition of the field into harmonic oscillators. The presence of the volume element $d^3k/k$ underscores the relativistic content of this formula. The spinors $\kappa_A$ are related to the integration variables ${\bm k}$ through the formulae:
\begin{eqnarray}\label{f1}
\kappa_{\dot{A}}g^{\mu{\dot A}A}\kappa_A=k^\mu,\quad \kappa^{\dot A}\kappa^A=\textstyle\frac{1}{2}g^{\mu{\dot A}A}k_\mu.
\end{eqnarray}
The equations (\ref{f1}) do not determine the overall phase of $\kappa_A$. However, this phase is not significant because it can be absorbed by a change of the phases of the amplitudes $f_\pm(\bm k)$ in classical theory or the phases of the annihilation and creation operators in quantum theory. A convenient choice of the spinor $\kappa_A$ is:
\begin{eqnarray}\label{eta}
\{\kappa_A\}=\frac{1}{\sqrt{2(k-k_z)}}\left\{k_x-ik_y,k-k_z\right\}.
\end{eqnarray}
The wave equation (\ref{fe}) is satisfied due to the relations:
\begin{eqnarray}\label{id}
g^{\mu{\dot A}A}k_\mu\kappa_A=g^{\mu{\dot A}A}\kappa_{\dot{B}}g_\mu^{{\dot B}B}\kappa_B\kappa_A=2\epsilon^{{\dot A}{\dot B}}\kappa_{\dot{B}}\epsilon^{AB}\kappa_A\kappa_B=0.
\end{eqnarray}
The field equations retain the same form (\ref{fe}) also for the field operators ${\hat\phi}_{AB\dots L}(x)$ in quantum field theory. In this case the amplitudes $f_\pm(\bm k)$ in the expansion (\ref{sup}) into plane waves must be replaced by the annihilation and creation operators of particles with positive and negative helicity,
\begin{eqnarray}\label{sup1}
\fl\qquad{\hat\phi}_{AB\dots L}(x)
=\gamma\int\!\frac{d^3k}{(2\pi)^{3/2}k}\kappa_A\kappa_B\dots\kappa_L
\left[a_+({\bm k})e^{-ik\cdot x}+a_-^\dagger({\bm k})e^{ik\cdot x}\right].
\end{eqnarray}
The prefactor $\gamma$ is essential because the field operator ${\hat\phi}_{AB\dots L}(x)$ carries the dimensionality of the corresponding physical field while the dimensionality of the integral on the right-hand side is {\em fixed} by the canonical commutation relations (\ref{cr}). It follows from these relations that the annihilation and creation operators have the dimension of length. Therefore, the dimension of the integral is $1/{\rm length}^{H+1}$.
In what follows I shall need the following equality proved in \ref{aa}:
\begin{eqnarray}\label{form}
\fl\qquad{\int\hspace{-0.5cm}\sum}\!h({\bm k})a^\dagger_\lambda({\bm k})a_\lambda({\bm k})&=\int\!d^3r\int\!d^3r'\,{\tilde h}({\bm r}-{\bm r}')\nonumber\\
&\times{\hat\phi}_{\dot{A}\dot{B}\dots\dot{L}}({\bm r},0)g^{0\dot{A}A}g^{0\dot{B}B}\!\dots g^{0\dot{L}L}{\hat\phi}_{AB\dots L}({\bm r'},0),
\end{eqnarray}
where the symbol ${\int\hspace{-0.3cm}\Sigma}$ stands for $\Sigma_\lambda\!\int\!d^3k/k$, the function $h({\bm k})$ must be invariant under reflections, $h({\bm k})=h(-{\bm k})$, and
\begin{eqnarray}\label{ht}
{\tilde h}({\bm r})=\int\!\frac{d^3k}{(2\pi)^3}\,e^{i{\bm k}\cdot{\bm r}}\frac{h({\bm k})}{\gamma^2k^{2H-1}}.
\end{eqnarray}
For odd functions, $h({\bm k})=-h(-{\bm k})$, the contribution from negative helicity on the left hand side enters with the negative sign. In the formula (\ref{form}) and also in (\ref{emt}), (\ref{energy}), (\ref{energy1}), and (\ref{spr}) normal ordering of annihilation and creation operators is implied.
The general formulation will now be applied to the electromagnetic field and then to the gravitational field. The corresponding annihilation and creation operators will be denoted by $(c,c^\dagger)$ and $(g,g^\dagger)$, respectively, while $(a,a^\dagger)$ will be used in the general case. The well established electromagnetic case will serve as a guide in the construction in the gravitational case. I shall often use interchangeably the classical fields and their quantum counterparts but it should be clear from the context what is meant.
\section{Quantized Maxwell theory}\label{max}
Classical electromagnetic field may be described by the second-rank symmetric spinor $\phi_{AB}(x)$. The corresponding field operator ${\hat\phi}_{AB}(x)$ is connected with the electro\-magnetic field operator ${\hat f}_{\mu\nu}$ through the formulas (dotted indices for operators imply Hermitian conjugation):
\begin{eqnarray}\label{ftophi}
{\hat\phi}_{AB}(x)=\textstyle{\frac{1}{4\sqrt{2}}}
S^{\mu\nu}_{\;\;\;AB}{\hat f}_{\mu\nu}(x),\\
{\hat f}_{\mu\nu}(x)
=\frac{1}{\sqrt{2}}\left(S_{\mu\nu}^{\;\;\;AB}{\hat\phi}_{AB}(x)
+S_{\mu\nu}^{\;\;\;{{\dot A}{\dot B}}}{\hat\phi}_{{\dot A}{\dot B}}(x)\right).
\end{eqnarray}
The numerical coefficients in these formulas follows from the convention adopted in \cite{bb}. The value of the electromagnetic prefactor $\gamma_E$ appearing in the formula:
\begin{eqnarray}\label{sup2}
\fl\qquad{\hat\phi}_{AB}(x)
=\gamma_E\int\!\frac{d^3k}{(2\pi)^{3/2}k}\kappa_A\kappa_B
\left[c_+({\bm k})e^{-ik\cdot x}+c_-^\dagger({\bm k})e^{ik\cdot x}\right],
\end{eqnarray}
can be found by comparing the expression (\ref{anch}) for the energy operator of the electromagnetic field with the one constructed from the 00-component of the energy-momentum tensor. In the spinorial representation this tensor has the form:
\begin{eqnarray}\label{emt}
{\hat T}^{\mu\nu}=2\epsilon_0{\hat\phi}_{{\dot A}{\dot B}}g^{\mu{\dot A}A}g^{\nu{\dot B}B}{\hat\phi}_{AB},
\end{eqnarray}
where $\epsilon_0$ is the permittivity of free space. Thus, the following two expressions must be equal:
\begin{eqnarray}\label{energy}
{\int\hspace{-0.5cm}\sum}\hbar\omega\,c^\dagger_\lambda({\bf k})c_\lambda({\bm k})=2\epsilon_0\!\int\!d^3r\,{\hat\phi}_{{\dot A}{\dot B}}g^{0{\dot A}A}g^{0{\dot B}B}{\hat\phi}_{AB}.
\end{eqnarray}
The left hand side with the use of (\ref{form}) can be expressed in terms of spinors as follows:
\begin{eqnarray}\label{energy1}
{\int\hspace{-0.5cm}\sum}\hbar\omega\,c^\dagger_\lambda({\bf k})c_\lambda({\bm k})=\frac{\hbar c}{\gamma_E^2}\int\!d^3r\,{\hat\phi}_{{\dot A}{\dot B}}g^{0{\dot A}A}g^{0{\dot B}B}{\hat\phi}_{AB}.
\end{eqnarray}
Comparing this result with (\ref{energy}) we must choose
$\gamma_E=\sqrt{\hbar c/2\epsilon_0}$ and we obtain finally:
\begin{eqnarray}\label{fem}
\fl\qquad{\hat\phi}_{AB}(x)=\sqrt{\hbar c/2\epsilon_0}\int\!\frac{d^3k}{(2\pi)^{3/2}k}\kappa_A\kappa_B
\left[c_+({\bm k})e^{-ik\cdot x}+c_-^\dagger({\bm k})e^{ik\cdot x}\right].
\end{eqnarray}
\section{Quantized linearized gravity}\label{lin}
In the standard approach to linearized gravity (see, for example, \cite{mtw}) one starts from the decomposition of the metric tensor into the background metric (usually Minkowskian) and a small addition, $g_{\mu\nu}=\eta_{\mu\nu}+h_{\mu\nu}$. Note that the smallness of $h_{\mu\nu}$ does not necessarily imply the smallness of its derivatives so that the linearized Riemann tensor does not have to be small. Next, one proceeds to express the linearized Riemann tensor (I shall keep referring to this tensor as the Riemann tensor even though in this case it effectively reduces to the Weyl tensor) in terms of $h_{\mu\nu}$ and its derivatives. I will bypass all these intermediate steps and following \cite{pr} I will connect directly $R_{\mu\nu\lambda\rho}$ to its representation by the symmetric fourth-rank spinor,
\begin{eqnarray}
\phi_{ABCD}=\textstyle{\frac{1}{16}}
S^{\mu\nu}_{\;\;\;AB}S^{\lambda\rho}_{\;\;\;CD}R_{\mu\nu\lambda\rho},\label{rela}\\
R_{\mu\nu\lambda\rho}=\frac{1}{4}\left(S_{\mu\nu}^{\;\;\;AB}S_{\lambda\rho}^{\;\;\;CD}\phi_{ABCD}
+S_{\mu\nu}^{\;\;\;{\dot A}{\dot B}}S_{\lambda\rho}^{\;\;\;{\dot C}{\dot D}}\phi_{{\dot A}{\dot B}{\dot C}{\dot D}}\right).\label{relb}
\end{eqnarray}
In order to express the field operator $\hat{R}_{\mu\nu\lambda\rho}$ in terms of annihilation and creation operators of gravitons we need the normalization factor $\gamma_G$ in the formula:
\begin{eqnarray}\label{fg}
\fl\qquad{\hat\phi}_{ABCD}(x)=\gamma_G\int\!\frac{d^3k}{(2\pi)^{3/2}k}
\kappa_A\kappa_B\kappa_C\kappa_D
\left[g_+({\bm k})e^{-ik\cdot x}+g_-^\dagger({\bm k}) e^{ik\cdot x}\right].
\end{eqnarray}
The calculation of $\gamma_G$ is a nontrivial task since the expression for the energy of the gravitational field in terms of the Riemann tensor is yet to be determined. Therefore, at this stage I will use a very good substitute --- the Bel-Robinson tensor. The time component of this tensor $T^{0000}$ is positive and it may play sometimes the role of the energy. It has even been called the {\em superenergy} \cite{bel,bel1,bel2}. As is shown in \ref{bb}, the calculation of $\gamma_G$ in this way gives:
\begin{eqnarray}\label{gam}
\gamma_G=\sqrt{8\pi}\,\ell_P,
\end{eqnarray}
where $\ell_P=\sqrt{\hbar G/c^3}$ is the Planck length. Of course, from pure dimensional analysis we can deduce that $\gamma_G$ has the dimension of length so that it must be proportional to the Planck length but the determination of the numerical coefficient requires the calculations in \ref{bb}. Having determined $\gamma_G$ we may freely pass from the description in terms of oscillators labeled by ${\bm k}$ and $\lambda$ to the description in terms fields living in space-time with the use of the formula:
\begin{eqnarray}\label{fg1}
\fl\quad{\hat\phi}_{ABCD}(x)=\sqrt{8\pi\hbar G/c^3}\int\!\frac{d^3k}{(2\pi)^{3/2}k}
\kappa_A\kappa_B\kappa_C\kappa_D
\left[g_+({\bm k})e^{-ik\cdot x}+g_-^\dagger({\bm k}) e^{ik\cdot x}\right].
\end{eqnarray}
\section{Wigner functional of the electromagnetic field}\label{wig1}
The main tool in the study of the fluctuations will be here the Wigner functional at finite temperature. This functional for the electromagnetic field was obtained in \cite{ibb} but I shall sketch here the main steps of the derivation to continue easily to the gravitational case. I start with the Wigner function at finite temperature for the one-dimensional harmonic oscillator \cite{davies,hosw},
\begin{eqnarray}\label{wt}
W_T(x,p)=C\exp\left[-2\tanh\!\left(\!\frac{\hbar\omega}{2k_BT}\!\right)
\frac{H(p,x)}{\hbar\omega}\right],
\end{eqnarray}
where $H(p,x)=p^2/2m+m\omega^2x^2/2$ is the Hamiltonian of the harmonic oscillator. The ratio $N=H(p,x)/\hbar\omega$ is the number of quanta (the energy divided by the energy of one quantum) expressed in terms of classical variables $(p,x)$. At $T=0$, i.e. in the ground state, $W_G=C\exp(-2N)$. The normalization constant $C$ is unimportant since for the infinite number of oscillators only the relative probabilities can be determined.
The Wigner functional of the thermal state of the electro\-magnetic field is constructed by replacing a single oscillator with the whole collection of oscillators labeled by $\bm k$ and $\lambda$ \cite{ibb},
\begin{eqnarray}\label{wigel}
\fl W_{EM}^T[{\bm E},{\bm B}]=\exp\left[-\int\!\!d^3r\!\!\int\!\!d^3r'f_E(\vert{\bm r}-{\bm r}\,'\vert)\left({\bm E}({\bm r}\,)
\!\cdot\!{\bm E}({\bm r}\,')+c^2{\bm B}({\bm r}\,)\!\cdot\!{\bm B}({\bm r}\,')\right)\right].
\end{eqnarray}
The electromagnetic correlation function $f_E(r)$ is the three-dimensional Fourier transform of the function appearing in the Wigner function for the one-dimensional oscillator,
\begin{eqnarray}\label{kerel}
\fl\qquad f_E(r)=\frac{1}{2\gamma_E^2}\int\!\frac{d^3k}{(2\pi)^3}\frac{\tanh
\left(\ell_Qk/2\right)}{k}e^{i{\bm k}\cdot{\bm r}}
=\frac{\epsilon_0}{2\pi\hbar c\ell_Q r\sinh(\pi r/\ell_Q)},
\end{eqnarray}
where the quantum thermal length is $\ell_Q=\hbar c/k_BT$=0.0023m/T[K].
In the derivation of (\ref{wigel}) I used (\ref{form}) and the following relation:
\begin{eqnarray}\label{rel1}
\fl\qquad\phi_{\dot{A}\dot{B}}({\bm r})g^{0\dot{A}A}g^{0\dot{B}B}\phi_{AB}({\bm r'})=\frac{1}{4}\left({\bm E}({\bm r})\!\cdot\!{\bm E}({\bm r}\,')+c^2{\bm B}({\bm r})\!\cdot\!{\bm B}({\bm r}\,')\right).
\end{eqnarray}
From the integral representation (\ref{kerel}) of $f_E(r)$ we obtain in the classical limit,
\begin{eqnarray}\label{lim}
\lim_{\hbar\to 0}f_E(r)=\frac{\epsilon_0}{2}\,\delta(\bm r).
\end{eqnarray}
Therefore, in this limit the Wigner functional becomes the Boltzmann distribution,
\begin{eqnarray}\label{boltz}
W_{EM}^{\rm cl}=\exp(-E_{EM}/k_BT).
\end{eqnarray}
At the other end, at $T=0$, the Wigner functional is equal to:
\begin{eqnarray}\label{zeld}
\fl\qquad W_{EM}^0[{\bm E},{\bm B}]=\exp\left[-\epsilon_0\int\!\!d^3r\!\!\int\!\!d^3r'\frac{{\bm E}({\bm r}\,)
\!\cdot\!{\bm E}({\bm r}\,')+c^2{\bm B}({\bm r}\,)\!\cdot\!{\bm B}({\bm r}\,')}{2\pi^2\hbar c\vert{\bm r}-{\bm r}\,'\vert^2}\right],
\end{eqnarray}
i.e. equal to $\exp(-2N)$ where $N$ is the total number of photons as calculated by Zeldovich \cite{zeld,ibb}.
\section{Wigner functional of the gravitational field}\label{wig2}
The thermal Wigner functional of the gravitational field can be constructed in the same way as for the electromagnetic field. The resulting formula is:
\begin{eqnarray}\label{wigg}
\fl W_G^T(R)=\exp\left[- \int\!\!d^3r\!\!\int\!\!d^3r'f_G(\vert{\bm r}-{\bm r}\,'\vert)\sum\limits_{ij}\left({\cal E}_{ij}({\bm r}\,)
{\cal E}_{ij}({\bm r}\,')+{\cal B}_{ij}({\bm r}\,){\cal B}_{ij}({\bm r}\,')\right)\right],
\end{eqnarray}
where ${\cal E}_{ij}=R_{i0j0}$ and ${\cal B}_{ij}=\textstyle{\frac{1}{2}}\epsilon_{ikl}R_{j0}^{\;\;\;kl}$ are the so called electric and magnetic parts of the curvature tensor \cite{mb,kt}. In the derivation of this formula I used the following gravitational counterpart of (\ref{rel1}) derived with the help of (\ref{form}):
\begin{eqnarray}\label{rep}
\fl\phi_{\dot{A}\dot{B}\dot{C}\dot{D}}({\bm r})g^{0\dot{A}A}g^{0\dot{B}B}g^{0\dot{C}C} g^{0\dot{D}D}
\phi_{ABCD}({\bm r'})=\sum_{ij}\left({\cal E}_{ij}({\bm r}){\cal E}_{ij}({\bm r'})+{\cal B}_{ij}({\bm r}){\cal B}_{ij}({\bm r'})\right).
\end{eqnarray}
The gravitational correlation function $f_G(r)$ can be explicitly evaluated and is given by the following counterpart of the formula (\ref{kerel}),
\begin{eqnarray}\label{kerg}
\fl\qquad\qquad f_G(r)&=\frac{1}{\gamma_G^2}\int\!\frac{d^3k}{(2\pi)^3}\frac{2\tanh \left(\ell_Qk/2\right)}{k^3}e^{i{\bm k}\cdot{\bm r}}\nonumber\\
&=\frac{1}{8\pi^4\ell_Gr}\left[\frac{\pi^2}{3}+\ln(\zeta)\ln(1+\zeta)
+{\rm Li}_2(1-\zeta)+{\rm Li}_2(-\zeta)\right],
\end{eqnarray}
where $\ell_G=Gk_BT/c^4=1.14\times 10^{-67}$mT[K] is the gravitational thermal length, $\zeta=\coth(\pi r/2\ell_Q)$, and ${\rm Li}_2$ is the dilogarithm function.
There is a great similarity between the probability distributions of various field configurations for the electromagnetic field and the gravitational field. In both cases the smaller the distance between the points, the more likely it is that the electromagnetic field or the curvature tensor at these points will have opposite signs. Thus, the formula (\ref{wigg}) may be viewed as a realization of the Wheeler concept of the virtual gravitational foam \cite{jaw}. Of course, in the linearized version of gravity there is no room for fluctuations of topology envisaged by Wheeler but the rapid changes of geometry at small distances do show a certain similarity with Wheeler's foam.
In the classical limit in the electromagnetic case we obtained the standard Boltzmann distribution (\ref{boltz}) and the same result holds in the gravitational case, $W_G^{\rm cl}(R)=\exp(-E_G/k_BT)$, where
\begin{eqnarray}\label{ham}
E_G= c^4\int\!\!d^3r\!\!\int\!\!d^3r'
\sum\limits_{ij}\frac{{\cal E}_{ij}({\bm r}\,)
{\cal E}_{ij}({\bm r}\,')+{\cal B}_{ij}({\bm r}\,){\cal B}_{ij}({\bm r}\,')}{32\pi^2G\vert{\bm r}-{\bm r}\,'\vert}.
\end{eqnarray}
One may check that $E_G$ is indeed equal to (\ref{anch}) by expressing (\ref{ham}), with the use of (\ref{form}) and (\ref{rep}), in terms of annihilation and creation operators,
\begin{eqnarray}\label{genergy}
E_G=\int\hspace{-0.5cm}\sum\hbar\omega\,g^\dagger_\lambda({\bf k})g_\lambda({\bm k}).
\end{eqnarray}
The equality of these two expressions proves the self-consistency of the whole procedure. The classical (high-temperature) limit gives the correct expression for the energy both in the electromagnetic and in the gravitational case. The nonlocal form of $E_G$ fully confirms the belief that there is no ``local gravitational energy-momentum'' \cite{mtw}.
There are, however, substantial differences between the gravitational and electro\-magnetic correlation functions. At large distances the gravitational correlation function $f_G(r)$ does not fall-off exponentially, as in electromagnetism, but has a {\em long tail} equal to its classical limit $f_G(r)\approx f^{\rm cl}_G(r)= 1/(32\pi^2\ell_Gr)$. There is also a striking difference in the behavior at $T\to 0$, as described in the next Section.
\section{Problems with the ground state}\label{grnd}
A puzzling phenomenon is the logarithmic divergence of the gravitational correlation function $f_G(r)$ in the limit, when $T\to 0$. This fact indicates that there might be a problem with the gravitational ground state.
The ground state of the quantized gravitational field has been studied before by Kucha{\v{r}} \cite{kk} and Hartle \cite{jbh}. They arrived at a formula for the ground state functional but their analogy between the electromagnetism and gravity is based on a different paradigm. They saw the analogy between the electromagnetic potentials and the metric tensor while in my approach, based on the spinorial description, there is a correspondence between the electromagnetic field tensor and the Riemann tensor.
The roots of the problem are in the expression (\ref{ham}) for the energy. To see the connection between the energy and the ground state let us consider the wave functions of a one-dimensional harmonic oscillator in the position representation and the momentum representation:
\begin{eqnarray}\label{gs}
\psi(x)=N_x\exp\left(-\frac{m\omega x^2}{2\hbar\omega}\right),\quad
{\tilde\psi}(p)=N_p\exp\left(-\frac{p^2}{2m\hbar\omega}\right).
\end{eqnarray}
In the position representation the exponent is equal to the potential energy divided by $\hbar\omega$ and in the momentum representation it is the kinetic energy divided by $\hbar\omega$. Applying the same prescription to the electromagnetic field we obtain the Wheeler expression \cite{jaw} (also rederived in \cite{kk} and \cite{jbh}). This follows directly from the observation that the division by $\hbar\omega$ in quantum mechanics is represented in quantum field theory by the following kernel
\begin{eqnarray}\label{ker}
\int\!\frac{d^3k}{\hbar\omega_k}e^{i{\bm k}\cdot({\bm r}-{\bm r}')} =
\frac{4\pi}{\hbar c\vert{\bm r} - {\bm r}'\vert^2}.
\end{eqnarray}
However, in the gravitational case we do not reproduce the Kucha{\v{r}}-Hartle result but we encounter the same logarithmic divergence that appeared in the Wigner functional. The convolution of this kernel with the already present kernel $1/\vert{\bm r} - {\bm r}'\vert$ in the formula (\ref{energy}) for the energy leads to the logarithmic divergence at large values of $r$. The same divergence appears already in the formula for the number of gravitons obtained by setting $h({\bm k})=1$ and $H=2$ in (\ref{ht}).
The divergence of the gravitational probability at $T=0$ can also be related to the behavior of the two-point function (sometimes called Riemann correlator \cite{frv}) at small distances. With the use of the formulas (\ref{sup1}) and (\ref{relb}) we can evaluate the vacuum expectation value
\begin{eqnarray}\label{rc}
\langle 0|{\hat\phi}_{AB\dots L}(x){\hat\phi}_{{\dot{A}}{\dot{B}}\dots{\dot{L}}}(\bm x')|0\rangle
\end{eqnarray}
for the spinor fields of rank $2H$. The resulting expression will have $2H$ derivatives acting on $\Delta^{(+)}(x-x')$---the invariant (positive frequency) solution of the wave equation. For a Gaussian distribution there is a simple relation between the matrix of the second moments $\langle x_ix_j\rangle$ and the matrix in the probability distribution. Namely, one is the inverse of the other. There exists an analogy between this discrete case and our field-theoretic continuous case. The two-point function is the analog of the covariance matrix and the correlation functions $f_E$ and $f_G$ in (\ref{wigel}) and (\ref{wigg}) are the analogs of the inverse of the covariance matrix. The divergence of $f_G$ at zero temperature can be traced to the nonexistence of the inverse of the gravitational two-point function. The presence of four derivatives in this function results in the fourth power of $k$ in the Fourier transform and its inverse is not integrable at small $k$. In the electromagnetic case we have the second derivatives and this leads to an integrable Fourier transform.
A technical cure for the problem with non-integrability could be to disallow the interchange of the integrations over ${\bm r}$ and ${\bm r}'$ in the formula (\ref{wigg}) and the evaluation of the limit when $T\to 0$. Unfortunately, the limit would then depend on the behavior of the Riemann tensor at large distances. Since the divergence is of the ``infrared type'' it can presumably be cured by assuming a finite radius of the Universe but this is outside the scope of this paper. The problem with the ground state might also mean, however, that there is some truth in Dyson's hypothesis.
\section{Conclusions}
I have shown that the quantization of the linearized gravitational field that employs only the Riemann tensor (and not the metric tensor) can be achieved without any reference to the canonical formalism. In this approach the complications arising in the process of extracting true degrees of freedom never appear. The main result is the evaluation of the probability distribution of various geometries in the thermal state of the gravitational field. An encountered puzzle is the divergence of this probability distribution at $T\to 0$ that might signify some problems with the ground state.
\ack
I would like to thank Zofia Bialynicka-Birula for very fruitful criticism and Albert Roura and Jerzy Kijowski for discussions and references. This research was financed by the Polish National Science Center Grant No. 2012/07/B/ST1/03347.
|
\section{\label{sec:intro}Introduction}
Combinatorial optimization appears in many important fields such as computer science \cite{karp1972complexity, mezard1987spin}, drug discovery and life-science \cite{kitchen2004docking}, and information processing technology \cite{nishimori2001statistical}.
One of the example of such problems is an Ising problem to minimize the Ising Hamiltonian, which is a function of a spin configuration $\sigma = (\sigma_i)$ defined as
\begin{eqnarray}\label{eq:ising}
\mathcal{H}(\sigma) = - \sum_{i < j}J_{ij}\sigma_i\sigma_j\quad (1 \leq i,j \leq N),
\end{eqnarray}
where each spin takes binary values $ \sigma_i \in \{\pm 1\}$, a real number symmetric matrix $J_{ij}$ denotes a coupling constant, and $N$ is the total number of spins.
Despite its simple statement, it belongs to the non-deterministic polynomial-time (NP)-hard class to find the ground state of the Ising model on the three-dimensional lattice \cite{barahona1982computational}.
Similarly, a maximum cut (MAX-CUT) problem in the graph theory is to find the size of the largest cut in a given undirected graph. Here, a cut is a partition of the vertices $V$ into two disjoint subsets $\{S_1, S_2\}$ and the size of the cut is the total weight of edges $w_{ij}$ with one vertex $i$ in $S_1$ and the other $j$ in $S_2$. The size of the cut can be counted by assigning the binary spin values to express which subset the vertex $i$ belongs to $\sigma_i \in \{\pm 1\}$ \cite{johnson1979computers}:
\begin{eqnarray}\label{eq:cut}
C(\sigma) &=& \sum_{i \in S_1, j \in S_2}w_{ij} = \sum_{i<j}w_{ij} \frac{(1 - \sigma_i \sigma_j)}{2} \nonumber\\
&=& \frac{1}{2}\sum_{i<j}w_{ij} - \frac{1}{2}\mathcal{H}(\sigma),
\end{eqnarray}
where $\mathcal{H}$ is an Ising Hamiltonian defined in Eq.~(\ref{eq:ising}) with $J_{ij} = -w_{ij}$. It indicates that the MAX-CUT problem is equivalent to the Ising problem except for the constant factor.
The MAX-CUT problem belongs to the NP-hard class in general, even though there are graph topologies which can be solved in polynomial time \cite{barahona1982computational, orlova1972finding, hadlock1975finding, grotschel1981weakly, grotschel1984polynomial, galluccio2001optimization}. Many attempts have been made to approximately solve NP-hard MAX-CUT problems, but the probabilistically checkable proof (PCP) theorem states that no polynomial time algorithms can approximate MAX-CUT problems better than $0.94118$ \cite{arora1998proof, haastad2001some}.
Currently, the approximation ratio of $0.87856$ achieved by the Goemans-Williamson algorithm (GW) based on semidefinite programming (SDP) is the best value for performance guarantee \cite{goemans1995improved}. This algorithm is a well-established benchmark to evaluate any new algorithms or computing methods.
Besides, there exist several heuristic algorithms to tackle these NP-hard MAX-CUT problems. The simulated annealing (SA) was designed by mimicking the thermal annealing procedure in metallurgy \cite{kirkpatrick1983optimization}. A quantum annealing technique was also formulated and was shown to have competitive performance against SA \cite{kadowaki1998quantum, santoro2002theory, farhi2001quantum, van2001powerful, aharonov2008adiabatic}.
Independently, novel algorithms which are superior either in its speed or its accuracy are proposed \cite{sahni1976p, kahruman2007greedy, benlic2013breakout}.
We recently proposed a novel computing system to implement the NP-hard Ising problems using the criticality of laser \cite{utsunomiya2011mapping, takata2012transient, takata2014data, utsunomiya2015binary} and degenerate optical parametric oscillator (DOPO) phase transition \cite{wang2013coherent, marandi2014network}. The architecture of this machine is motivated by the principle of laser and DOPO in which the mode with the minimum loss rate is most likely to be excited first. The energy of the Ising Hamiltonian can be mapped onto the total loss rate of the laser or DOPO network. The selected oscillation mode in the laser or DOPO network corresponds to the ground state of a given Ising Hamiltonian, while the gain accessible to all other possible modes is depleted due to the cross-gain saturation. This means that a mode with the lowest loss rate reaches a threshold condition first and clumps the gain at its loss rate, so that all the other modes with higher loss rates stay at sub-threshold conditions.
Moreover, the DOPO is in the linear superposition of $0$-phase state and $\pi$-phase state at its oscillation threshold \cite{drummond1980}. The coupled DOPOs form quantum entanglement in spite of their inherent dissipative natures \cite{takata2015quantum}, so that some form of quantum parallel search could be embedded in a DOPO network.
In this article, the validity of the CIM for MAX-CUT problems is tested against the representative approximation algorithms. The DOPO signal pulse amplitudes in CIM, which are interpreted as the solution, are described by the c-number stochastic differential equations (CSDE) as presented in Section \ref{sec:MF}. Then we conduct numerical simulations for MAX-CUT problems in Section \ref{sec:maxcut} with the number of vertices up to $N = 20000$.
It is, of course, difficult to compare the performance of the proposed system as a MAX-CUT solver with the representative approximation algorithms which can be run on current digital computers mainly because the unit of ``clock'' cannot be uniquely defined. Thus we defined the feasible system clock which dominates the computing process in CIM as mentioned later. Moreover, here we evaluated the computational ability under either time or accuracy was fixed, while a preliminary benchmark study done in the previous paper is focused on the performance after physical convergence \cite{marandi2014network}.
\section{\label{sec:MF}Multiple-pulse DOPO with mutual coupling}
\subsection{\label{sec:spec}A proposed machine}
A standard CIM based on multiple-pulse DOPO with all-optical mutual coupling circuits is shown in Fig. \ref{fig:CIM}. The system starts with a pulsed master laser at a wavelength of $1.56\ \mu \mathrm{m}$. A second harmonic generation (SHG) crystal produces the pulse trains at a wavelength of $0.78\ \mu \mathrm{m}$ which in turn generate multiple DOPO pulses at a wavelength of $1.56\ \mu \mathrm{m}$ inside a fiber ring resonator. If the round trip time of the fiber ring resonator is properly adjusted to $N$ times the pump pulse interval, we can simultaneously generate $N$ independent DOPO pulses inside the resonator. Each of these pulses is either in $0$-phase state or $\pi$-phase state at well above the oscillation threshold and represents an Ising spin of up or down.
\begin{figure*}
\includegraphics[width=4.in]{Figure1.eps}
\caption{\label{fig:CIM}A coherent Ising machine based on the time-division multiplexed DOPO with mutual coupling implemented by optical delay lines. A part of each pulse is picked off from the main cavity by the output coupler followed by an optical phase sensitive amplifier (PSA) which amplifies the in-phase amplitude $\tilde{c}_i$ of each DOPO pulse. The feedback pulses, which are produced by combining the outputs from $N-1$ intensity and phase modulators, are injected back to the main cavity by the injection coupler.}
\end{figure*}
In order to implement an Ising coupling $J_{ij}$ in Eq. (\ref{eq:ising}), a part of each DOPO pulse in the fiber ring resonator is picked-off and fed into an optical phase sensitive amplifier (PSA), followed by optical delay lines with intensity and phase modulators. Using such $N-1$ optical delay lines, (arbitrary) $i$-th pulse can be coupled to (arbitrary) $j$-th pulse with a coupling coefficient $J_{ij}$.
Such an all-optical coupling scheme has been demonstrated for $N = 4$ and $N = 16$ CIMs \cite{marandi2014network, kenta2015}.
In Sections \ref{sec:SDP} and \ref{sec:SA}, we assume a CIM with a fiber length of $2\ \mathrm{km}$ (or cavity round trip time of 10 $\mu$s) and pulse spacing of 10 cm (or pulse repetition frequency of $2\ \mathrm{GHz}$), thus $2\times10^{4}$ independent DOPO pulses can be prepared for computation. The system clock frequency for the CIM should be defined by the cavity circulation frequency (inverse of cavity round trip time). One clock cycle (round trip) includes every elements of computation, such as parametric amplification, out-coupling port, and coherent feedback. Thus the clock frequency of the CIM assumed for the present benchmark study is 100 kHz since the round trip time of 2 km fiber ring is $10\ \mu\mathrm{s}$. We fixed this system clock frequency, just like any digital computer has a fixed clock frequency and chose the appropriate pulse interval to pack the desired number of pulses in the $2 \mathrm{km}$ fiber.
\subsection{c-number stochastic differential equations for multiple-pulse DOPO with mutual coupling}
The in-phase and quadrature-phase amplitudes of a single isolated DOPO pulse obey the following c-number stochastic differential equations (CSDE) \cite{kinsler1991quantum}:
\begin{eqnarray}
dc&=& (-1+p-c^2-s^2)c\, dt+\frac{1}{A_\mathrm{s}}\sqrt{c^2\!+\!s^2\!+\!\frac{1}{2}}dW_1\label{eq:1 CSDE} \\
ds&=& (-1-p-c^2-s^2)s\, dt+\frac{1}{A_\mathrm{s}}\sqrt{c^2\!+\!s^2\!+\!\frac{1}{2}}dW_2.\label{eq:2 CSDE}
\end{eqnarray}
The above CSDE are derived by expanding the DOPO field density operator with the truncated Wigner distribution functions.
An alternative approach is to use two coherent states in the generalized (off-diagonal) $P(\alpha_\mathrm{s}, \beta_\mathrm{s})$-representation for the field density matrix \cite{drumond1981non}. The two approaches by the truncated Wigner function and the generalised P-representation are equivalent for highly dissipative systems such as ours. The pump field is adiabatically eliminated in (\ref{eq:1 CSDE}) and (\ref{eq:2 CSDE}) by assuming that the pump photon decay rate $\gamma_\mathrm{p}$ is much larger than the signal photon decay rate $\gamma_\mathrm{s}$. The term $A_\mathrm{s}=(\gamma_\mathrm{s} \gamma_\mathrm{p} /2 \kappa^2)^{1/2}$ is the DOPO field amplitude at a normalized pump rate $p = F_\mathrm{p}/F_{\mathrm{th}}=2$, and $\kappa$ is the second order nonlinear coefficient associated with the degenerate optical parametric amplification. The variable $t=\gamma_\mathrm{s}\tau / 2$ is a normalized time, while $\tau$ is a real time in seconds. The term $F_\mathrm{p}$ is the pump field amplitude and $F_{\mathrm{th}}=\gamma_\mathrm{s} \gamma_\mathrm{p}/4\kappa$ is the threshold pump field amplitude. Finally, $dW_1$ and $dW_2$ are two independent Gaussian noise processes that represent the incident vacuum fluctuations from the open port of the output coupler and the pump field fluctuation for in-phase and quadrature-phase components, respectively. The vacuum fluctuation of the signal channel contributes to the 1/2 term and the quantum noise of the pump field contributes to $c^2+s^2$ in the square-root bracket in (\ref{eq:1 CSDE}) and (\ref{eq:2 CSDE}).
When the $i$-th signal pulse is incident upon the output coupler,
the output-coupled field and remaining field inside a cavity are written as
\begin{eqnarray}
c_{i, \mathrm{out}}&=&\sqrt{T}c_i-\sqrt{1-T}\frac{f_i}{A_\mathrm{s}}\label{eq:3 coupler 1}\\
c_{i, \mathrm{cavity}}&=&\sqrt{1-T}c_i+\sqrt{T}\frac{f_i}{A_\mathrm{s}}\label{eq:4 coupler 1},
\end{eqnarray}
where $T$ is the power transmission coefficient of the output coupler and $f_i$ is the incident vacuum fluctuation from the open port of the coupler. The out-coupled field and the signal amplitude after PSA can be
\begin{eqnarray}
\tilde{c}_{i} \equiv \frac{c_{i, \mathrm{out}}}{\sqrt{T}} = c_i-\sqrt{\frac{1-T}{T}}\frac{f_i}{A_\mathrm{s}}.\label{eq:5 coupler 1_}
\end{eqnarray}
From these out-coupled pulse stream, the intensity and phase modulators placed in the $N-1$ delay lines produce the mutual coupling pulse $\sum_j \xi_{ij}\tilde{c}_j$, which is actually added to the $i$-th signal pulse by an injection coupler. Here, $\xi_{ij}$ is the effective coupling coefficient from the $j$-th pulse to the $i$-th pulse, determined by the transmission coefficient $\sqrt{T'}$ of the injection coupler. In the highly dissipative limit of a mutual coupling circuit, such as in our scheme, we can use the CSDE supplemented with the noisy coupling term. Since the transmission coefficient $\sqrt{T'}$ of the injection coupler should be much smaller than one, we do not need to consider any additional noise in the injected feedback pulse. The CSDE (\ref{eq:1 CSDE}) can be now rewritten to include the mutual coupling terms
\begin{eqnarray}
dc_{i}=[(-1+p-c_i^2-s_i^2)c_i + \sum_{j}\xi_{ij}\tilde{c}_j]\, dt \nonumber\\
+\frac{1}{A_\mathrm{s}}\sqrt{c_i^2+s_i^2+\frac{1}{2}}dW_i.\label{eq:6 coupler 1,2}
\end{eqnarray}
The summation in Eq. (\ref{eq:6 coupler 1,2}) represents the quantum measurement-feedback term including the measurement error given by Eq. (\ref{eq:5 coupler 1_}). The vacuum fluctuation coupled to the $i$-th pulse in the output coupler is already taken into account in the last term of right-hand side of Eq. (\ref{eq:6 coupler 1,2}) together with the pump noise.
We conducted the numerical simulation of the coupled CSDE (\ref{eq:6 coupler 1,2}) to evaluate the performance of the CIM.
\section{\label{sec:maxcut}Benchmark studies on MAX-CUT problems}
\subsection{MAX-CUT problems on cubic graphs}
The MAX-CUT problem on cubic graphs, in which each vertex has exactly three edges, is called MAX-CUT-3 problem and also belongs to NP-hard class \cite{halperin2002max}. The smallest simple MAX-CUT-3 problem is defined on the complete graph $K_4$ with four vertices and six edges with identical weight $J_{ij} = -1$, where anti-ferromagnetic couplings have frustration so that the ground states are highly degenerate. The solution to this problem are the set of two-by-two cuts, which contains six degenerate ground states of the Ising Hamiltonian, i.e., $\{\left|\uparrow \uparrow \downarrow \downarrow \right\rangle, \left|\uparrow \downarrow \uparrow \downarrow \right\rangle, \left|\uparrow \downarrow \downarrow \uparrow \right\rangle, \left|\downarrow \uparrow \uparrow \downarrow \right\rangle, \left|\downarrow \uparrow \downarrow \uparrow \right\rangle, \left|\downarrow \downarrow \uparrow \uparrow \right\rangle\}$.
Figure \ref{fig:maxcut4} shows the time evolution of $c_i$ $(i=1,\dots, 4)$ when $p = 1.1$ and $\xi = -0.1$. A correct solution spontaneously emerges after several tens of round trips. The statistics of obtaining different states against 1000 sessions of such a numerical simulation are shown in Fig. \ref{fig:4mcut}, in which six degenerate ground states appear with almost equal probabilities with no errors found.
\begin{figure}[htbp]
\centering
\includegraphics[width=3.1in]{Figure2.eps}
\caption{\label{fig:maxcut4}Normalized DOPO signal amplitudes as a function of normalized time (in unit of cavity round trip numbers) for a $N=4$ simple MAX-CUT-3 problem. Each color corresponds to the four different DOPOs indexed with $i = 1, \dots, 4$. Small window is enlarged to indicate the status of signal amplitude inside a cavity at three components (as in Fig. \ref{fig:CIM}); A: OPA gain medium (PPLN waveguide), B: out-coupler, and C: injection coupler of the mutual coupling pulse. The two flat regions between B and C and between C and A are the passive propagation in a fiber.}
\end{figure}
\begin{figure}[htbp]
\centering
\includegraphics[width=3.1in]{Figure3.eps}
\caption{\label{fig:4mcut}Distribution of output spin configurations in 1000 trials of numerical simulations against a simple MAX-CUT-3 problem of graph order $N=4$. All trials were successful to find one of the six degenerate ground states.}
\end{figure}
\subsection{Many-body interaction problem}
If the interaction is not a standard two-body Ising interaction type but rather a four-body interaction such as
\begin{eqnarray}\label{eq:4body}
\mathcal{H}=-J_{1234}\sigma_1 \sigma_2 \sigma_3 \sigma_4,
\end{eqnarray}
where $J_{1234}\in \mathbb{R}$, the coupled field into the $i$-th pulse is no longer given by $\sum_j \xi_{ij} \tilde{c}_j$ but by $\xi \tilde{c_j} \tilde{c_k} \tilde{c_l} \ (j, k, l \neq i)$. In this case, the CSDE (\ref{eq:6 coupler 1,2}) can be rewritten to include the four-body coupling term
\begin{eqnarray}\label{eq:4bodyCSDE}
dc_{i}=[(-1+p-c_i^2-s_i^2)c_i+\xi \tilde{c}_j \tilde{c}_k \tilde{c}_l]\, dt \nonumber\\+\frac{1}{A_\mathrm{s}}\sqrt{c_i^2+s_i^2+\frac{1}{2}}dW_i.
\end{eqnarray}
When the four-body coupling coefficient $J_{1234}$ is $-1$ (multi-body anti-ferromagnetic coupling), there are eight degenerate ground states, i.e., $\left|\uparrow \uparrow \uparrow \downarrow \right\rangle, \left|\uparrow \uparrow \downarrow \uparrow \right\rangle, \left|\uparrow \downarrow \uparrow \uparrow \right\rangle, \left|\downarrow \uparrow \uparrow \uparrow \right\rangle$ and their inverse spin configurations. Figure \ref{fig:amp4} shows the time evolution of $c_i$ $(i=1 ,\dots, 4)$ when $p = 1.1$ and $\xi = -0.1$. One of the eight degenerate ground states emerges spontaneously after several tens of round trips. The statistics of observing different states in 1000 independent sessions of the numerical simulation of Eq. (\ref{eq:4bodyCSDE}) are shown in Fig. \ref{fig:4body}, in which eight degenerate ground states are obtained with no errors found.
\begin{figure}[htbp]
\centering
\includegraphics[width=3.1in]{Figure4.eps}
\caption{\label{fig:amp4}Normalized DOPO pulse amplitudes $c_i$ $(i=1, \dots, 4)$ under the interaction between four-body Ising coupling expressed by Eq. (\ref{eq:4body}).}
\end{figure}
\begin{figure}[htbp]
\centering
\includegraphics[width=3.1in]{Figure5.eps}
\caption{\label{fig:4body}Distribution of output spin configurations in 1000 trials of numerical simulation against a four-body Ising model of $N=4$. All trials were successful to find one of the eight degenerate ground states.}
\end{figure}
\subsection{\label{sec:algorithm}Algorithm description}
In this subsection, we will review the four representative approximation algorithms for MAX-CUT problems.
The Goemans-Williamson algorithm (GW) based on SDP has a $0.87856$-performance guarantee for NP-hard MAX-CUT problems \cite{goemans1995improved}. It achieves the optimal approximation ratio for MAX-CUT problems under the assumptions of $\mathrm{P}\neq \mathrm{NP}$ and the unique games conjecture \cite{khot2007optimal}. The SDP relaxation of the original MAX-CUT problem is a vector-valued optimization problem as maximizing $\ \frac{1}{2}\sum_{i < j}w_{ij}(1 - \vec v_i \cdot \vec v_j), \ \vec v_i \in S^{k-1}$, where $S^{k-1}$ is a unit sphere in $\mathbb{R}^k$ and $k \leq N$ (or $\#V$: number of vertices). There exist polynomial time algorithms to find the optimal solution of this relaxation problem (with error $\varepsilon > 0$), and its value is commonly called the SDP upper bound. A final solution to the original MAX-CUT problem is obtained by projecting the solution vector sets to randomly chosen one-dimensional Euclidean spaces (i.e., dividing the sphere by random hyperplanes).
There are three types of computational complexities of the best-known algorithms for solving the SDP relaxation problem. If a graph with $N$ vertices and $m$ edges is regular, the SDP problem can be approximately solved in almost linear time as $\tilde O (m) = O(m (\log N)^2 \varepsilon^{-4})$ using the matrix multiplicative weights method \cite{arora2007combinatorial}, where $\varepsilon$ represents the accuracy of the obtained solution. However, slower algorithms are required for general graphs. If the edge weights of the graph are all non-negative, the fastest algorithm runs in $\tilde O(Nm) = O(Nm (\log N)^2 \varepsilon^{-3})$ time based on the Lagrangian relaxation method \cite{klein1996efficient}. For graphs with both positive and negative edge weights, the SDP problem is commonly solved using the interior-point method, which scales as $\tilde O(N^{3.5}) = O(N^{3.5} \log(1/\varepsilon))$ \cite{alizadeh1995interior}. Besides, low rank formulation of SDP is effective when the graph is sparse \cite{yamashita2012latest, grippo2012speedp, fujisawa2014petascale}. In our computational experiments, the COPL\_SDP based on the interior point method was used as a solver for MAX-CUT problems \cite{ye1999copl}. The SDP upper bound $U_{\mathrm{SDP}}$ and the solution $C_{\mathrm{GW}}$ were obtained using the following parameters: interior point method was used until the relative gap $r_{\mathrm{gap}} = 1 - P_{\mathrm{obj}}/D_{\mathrm{obj}}$ reached $10^{-3}$, where $P_{\mathrm{obj}}$ and $D_{\mathrm{obj}}$ are the objective functions of the primal and dual of the SDP problem, respectively \cite{benson2000solving}. Random projection onto one-dimensional space was executed $N$ times.
For many practical applications, heuristic algorithms are more convenient to use, since the GW algorithm generally requires long computation time $\tilde O(N^{3.5})$. Metropolis et al. introduced a simple algorithm that can be used to provide an efficient simulation of a collection of atoms in equilibrium at a given temperature \cite{metropolis1953equation}. Kirkpatrick et al. applied the algorithm to optimization problems by replacing the energy of the atomic system to the cost function of optimization problems and using spin configurations $\sigma$, which is called the simulated annealing algorithm (SA) \cite{kirkpatrick1983optimization}. In each step of this algorithm, a system is given with a random spin flip and the resulting change $\Delta E$ in the energy is computed. If $\Delta E \leq 0$, the spin flip is always accepted, and the configuration with the flipped spin is used as the starting point of the next step. If $\Delta E > 0$, the spin is treated probabilistically, i.e., the probability that the new configuration will be accepted is $P(\Delta E) = \exp(-\Delta E / k_\mathrm{B} T)$ with a control parameter of system temperature $T$. This choice of $P(\Delta E)$ results in the system evolving into an equilibrium Boltzmann distribution. Repeating this procedure, with the temperature $T$ gradually lowered to zero in sufficiently long time, leads spins $\sigma$ to convergence to the lowest energy state. In practical case, with the finite time, the annealing schedule affects the quality of output values. Here in our numerical simulations, the temperature was lowered according to the logarithmic function \cite{hajek1988cooling}. Note that 1 Monte Carlo step corresponds to $N$ trials of spin flip.
Sahni and Gonzalez constructed a greedy algorithm for MAX-CUT problems, which has 1/2-performance guarantee \cite{sahni1976p}, and SG3 is a modified version of it \cite{kahruman2007greedy}. In this algorithm, nodes $V$ are divided into two disjoint subsets $\{S_1, S_2\}$ sequentially. For each iterative process, the node with the maximum score is selected, and it is put into either set $S_1$ or $S_2$ so as to earn larger cuts. Here, the score function of SG3 is defined as $x_i = |\sum_{j \in S_1} w_{ij} - \sum_{j \in S_2} w_{ij}|\ (i = 1, \dots, N)$. It stops when all the edges are evaluated to calculate the score function, thus SG3 scales as $O(m)$.
The power of breakout local search (BLS) appears in the benchmark result for G-set graphs \cite{benlic2013breakout}. It updated almost half of the best solutions in G-set with the specialized data structure for sorting and dedicated procedure to escape from local minima. The algorithm is combination of steepest descent and forced spin flipping: after being trapped by a local minima as a result of steepest descent procedure, three types of forced spin flipping (single, pair, and random) are probabilistically executed according to the vertex influence list (i.e., which vertex will increase the number of cut most when it's flipped) on each subset of partition.
These algorithms are coded in C/C++ and run on a single thread of a single core on a Linux machine with two 6-core Intel Xeon X5650 (2.67 GHz) processors and 94 GB RAM.
The CIM is simulated based on the coupled CSDE (\ref{eq:6 coupler 1,2}) on the same machine.
Note that the computation time of CIM does not mean the simulation time on the Linux machine but corresponds to the actual evolution time of a physical CIM.
\subsection{\label{sec:SDP}Computational accuracy on G-set instances}
The performance of a CIM with DOPO network was tested on the NP-hard MAX-CUT problems on sparse graphs, so-called G-set \cite{helmberg2000spectral}. These test instances were randomly constructed using a machine-independent graph generator written by G. Rinaldi, with the number of vertices ranging from 800 to 20000, edge density from $0.02\%$ to $6\%$, and topology from random, almost planar, to toroidal.
The output cut values of running the CIM, SA, GW, and the best known solutions so far we could find \cite{benlic2013breakout, ikuta2015max} for some of G-set graphs are summarized in Table \ref{OPO_SDP}. The results for CIM are obtained in 50 ms, which correspond to the performance of an experimental system after 5000 DOPO cavity round trips. The best result and ensemble average value for 100 trials are shown. Here, the parameters are set to be $p=1.6$, $\xi=-0.06$ and the coupling constant $\xi_{ij} = \xi w_{ij} / \sqrt{\langle k \rangle}$ is normalized by the square root of the graph average degree $\langle k \rangle$.
The hysteretic optimization method, in which the swinging and decaying Zeeman term that flips the signal amplitude (spin) back and forth, is implemented four times after 10 ms initial free evolution \cite{zarand2002using}. Each hysteretic optimization takes 10 ms so that the total search takes 50 ms.
The result of SA is also obtained in 50 ms for each graph. For GW, the computation time ranged between 2.3 s and $1.1\times 10^5$ s, depending on $N$. The best outputs of the CIM were $1.62 \pm 0.58\ \%$ better than GW but $0.38\pm 0.40\ \%$ worse than SA, and CIM found better cut against GW except for a toroidal graph (g50) and a disconnected random graph (g70).
\begin{table*}
\caption{\label{OPO_SDP}Performance of the coherent Ising machine, simulated annealing and Goemans-Williamson SDP algorithm in solving the MAX-CUT problems on sparse G-set graphs. $\#V$ is the number $N$ of vertices in the graph, $\#E$ is the number $m$ of edges, $U_{\mathrm{SDP}}$ is the optimal solution to the semidefinite relaxation of the MAX-CUT problem when the duality gap reaches to $10^{-3}$, and $C_{\mathrm{best}}$ is the best known result so far we could find. $C_{\mathrm{GW}}$ is the best solution obtained by $N$ projections after SDP. $C_{\mathrm{SA}}$ and $\langle{C_{\mathrm{SA}}}\rangle$ are the best and average values obtained by SA in 100 trials of 50 ms. $C_{\mathrm{CIM}}$ and $\langle{C_{\mathrm{CIM}}}\rangle$ are the best and average values in CIM in $100$ runs of 50 ms $(=$ 5000 DOPO cavity round trips), respectively. To make comparisons with each other, every cut value $C$ generated from each algorithm is normalized according to $(C+E_{\mathrm{neg}})/(U_{\mathrm{SDP}} + E_{\mathrm{neg}})$, where $E_{\mathrm{neg}}\geq 0$ is the number of negative edges. In the bottom of this table, the average and worst values of all 71 G-set graphs are shown.}
\begin{tabular}{l r r r c c c c c c}
\hline
Graph & $\#V$ & $\#E$ & $U_{\mathrm{SDP}}$ & $C_{\mathrm{best}}$ &$C_{\mathrm{GW}}$ & $C_{\mathrm{SA}}$ & $\langle{C_{\mathrm{SA}}}\rangle$ & $C_{\mathrm{CIM}}$ & $\langle{C_{\mathrm{CIM}}}\rangle$ \\
\hline
g1 & 800 & 19176 & 12083 & 0.9620 & 0.9457 & 0.9620 & 0.9597 & 0.9614 & 0.9570 \\
g6 & 800 & 19176 & 2656 & 0.9607 & 0.9448 & 0.9606 & 0.9592 & 0.9601 & 0.9559 \\
g11 & 800 & 1600 & 629 & 0.9540 & 0.9327 & 0.9526 & 0.9478 & 0.9455 & 0.9370 \\
g14 & 800 & 4694 & 3191 & 0.9602 & 0.9336 & 0.9580 & 0.9544 & 0.9514 & 0.9472 \\
g18 & 800 & 4694 & 1166 & 0.9500 & 0.9282 & 0.9492 & 0.9439 & 0.9434 & 0.9372 \\
g22 & 2000 & 19990 & 14136 & 0.9450 & 0.9191 & 0.9445 & 0.9409 & 0.9405 & 0.9361 \\
g27 & 2000 & 19990 & 4141 & 0.9435 & 0.9174 & 0.9422 & 0.9400 & 0.9390 & 0.9356 \\
g32 & 2000 & 4000 & 1567 & 0.9559 & 0.9272 & 0.9508 & 0.9478 & 0.9424 & 0.9384 \\
g35 & 2000 & 11778 & 8014 & 0.9588 & 0.9292 & 0.9551 & 0.9523 & 0.9471 & 0.9438 \\
g39 & 2000 & 11778 & 2877 & 0.9464 & 0.9226 & 0.9431 & 0.9399 & 0.9364 & 0.9318 \\
g43 & 1000 & 9990 & 7032 & 0.9471 & 0.9292 & 0.9471 & 0.9439 & 0.9458 & 0.9396 \\
g48 & 3000 & 6000 & 6000 & 1.0000 & 1.0000 & 1.0000 & 0.9919 & 1.0000 & 0.9747 \\
g51 & 1000 & 5909 & 4006 & 0.9606 & 0.9333 & 0.9583 & 0.9544 & 0.9506 & 0.9468 \\
g55 & 5000 & 12498 & 11039 & 0.9325 & 0.9006 & 0.9264 & 0.9215 & 0.9193 & 0.9160 \\
g57 & 5000 & 10000 & 3885 & 0.9561 & 0.9237 & 0.9496 & 0.9473 & 0.9419 & 0.9384 \\
g59 & 5000 & 29570 & 7312 & 0.9440 & 0.9148 & 0.9376 & 0.9356 & 0.9308 & 0.9288 \\
g60 & 7000 & 17148 & 15222 & 0.9313 & 0.8989 & 0.9231 & 0.9201 & 0.9191 & 0.9152 \\
g64 & 7000 & 41459 & 10466 & 0.9440 & 0.9143 & 0.9347 & 0.9324 & 0.9320 & 0.9299 \\
g67 & 10000 & 20000 & 7744 & 0.9554 & 0.9215 & 0.9480 & 0.9459 & 0.9411 & 0.9388 \\
g70 & 10000 & 9999 & 9863 & 0.9674 & 0.9633 & 0.9523 & 0.9479 & 0.9515 & 0.9482 \\
g81 & 20000 & 40000 & 15656 & 0.9551 & 0.9195 & 0.9187 & 0.9125 & 0.9393 & 0.9376 \\\hline
Average & & & & 0.9540 & 0.9303 & 0.9502 & 0.9469 & 0.9464 & 0.9415 \\
Worst & & & & 0.9313 & 0.8989 & 0.9187 & 0.9125 & 0.9185 & 0.9149 \\
\hline
\end{tabular}
\end{table*}
As the size of optimization problems increases, the average accuracy is important for practical applications. Table \ref{OPO_SDP} shows that for all G-set graphs, the average accuracy in 100 trials is 0.94148 to the SDP upper bound $U_{\mathrm{SDP}}$, i.e., the CIM can find a cut value larger than 0.94148 of the optimal value for the MAX-CUT problems on average, whereas the average accuracy of the GW is 0.93025 and that of the SA is 0.94692. Note that $U_{\mathrm{SDP}}$ is always greater than or equal to the optimal value for each MAX-CUT problem.
\subsection{\label{sec:SA}Computation time on MAX-CUT problems}
In the previous section, the running time of CIM and SA are fixed to estimate the computational accuracy.
These two algorithms explored the solutions as good as possible in 50 ms.
Although, if we finish the computation at a certain accuracy, more reasonable computation time can be defined.
Here, the GW solution was used as the mark of sufficient accuracy because it ensures the 87.856\% of the ground states.
The CIM and SA then competed the computation time to reach the same values obtained by GW.
The time and temperature scheduling parameters of the SA were set as follows: Inverse temperature increased with logarithmic function. The number of spin flipping was optimized to be $10^{l}$ times for some $l \in \mathbb{N}$, which requires the minimum computation time to achieve the same accuracy as with the GW.
\begin{figure*}[htbp]
\includegraphics[width=6.1in]{Figure6.eps}
\caption{\label{fig:energy}Performance comparison of CIM, SA, and GW in solving complete graphs (a) $K_{800}$ and (b) $K_{4000}$, where each edge was randomly weighted $\pm 1$. Each time, bundle of curves depicted average energy (solid line) $\pm$ standard deviations (dashed line) in 100 runs. Dotted line is the accuracy obtained by GW algorithm, whose computation time is shown with dot. The number of spin flip in SA algorithm were $10^5$ for $K_{800}$ and $10^6$ for $K_{4000}$, respectively, to optimize the computation time.
SA and GW are running on 2.67 GHz Intel Xeon without parallelization and CIM has the cavity circulation frequency of 100 kHz.}
\end{figure*}
Computational experiments were conducted on fully connected complete graphs, denoted by $K_N$, where the number of vertices $N$ ranging from 40 to 20000 and the edges are randomly weighted $\pm 1$. Figure \ref{fig:energy} shows the Ising energy in Eq. (\ref{eq:ising}) as a function of running time. Both CIM and SA run stochastically due to quantum and thermal noise, respectively, the ensemble average of energies are calculated as follows: For the CIM, the energy of all 100 runs was averaged at each round trip. For the SA, the averaged energy was calculated at each point on the time axis with an interpolated value from real time sampling. The parameters for CIM are chosen to be $p=0.2$, $\xi=-0.03$ (for $N=800$), and $\xi=-0.003$ (for $N=4000$). In Fig. \ref{fig:energy} (a), where $N = 800$, the GW achieved an energy equal to $-15624$ in $22.96\ \mbox{s}$, while the CIM and SA reached the same energy in $6.1 \times 10^{-4}\ \mbox{s}$ and in $0.671\ \mbox{s}$.
In Fig. \ref{fig:energy} (b), the GW achieved an energy of $-168160$ in $6646.25\ \mbox{s}$, while the CIM reached the same energy in $5.5 \times 10^{-4}\ \mbox{s}$ and the SA did so in $10.1\ \mbox{s}$.
Note that this result of SA and GW comes from a specific computer configuration as mentioned in Sec. \ref{sec:algorithm}. There is room for an improvement in the computation time in constant factor due to cases like using faster CPUs or parallelized codes.
Similarly, the computation time of CIM also depends on the system configuration and can be made faster when we use the higher clock frequency.
In this sense, the ratios between time of CIM and that of the other algorithms are arbitrary.
Thus we should study the computation time scaling as a function of the problem size.
\begin{figure*}[htbp]
\centering
\includegraphics[width=6.1in]{Figure7.eps}
\caption{\label{fig:time_scaling}(a) Computation time of coherent Ising machine (CIM) empirically scales as $O(1)$, while that of simulated annealing (SA), and Goemans-Williamson SDP (GW) fitted well to lines indicating $O(N^2)$ and $O(N^{3.5})$, respectively. Computation time of SG3 and BLS also fit to $O(N^2)$ with a difference of constant factor. The computation time of CIM, SA, SG3, and BLS is defined as the time to reach the same accuracy achieved by GW (without I/O time). Data points of CIM, SA, SG3, and BLS were calculated by averaging 100 runs. Only for $N = 800$ case, 100 randomly $\pm 1$ weighted complete graphs are generated. The results for $N = 800$ show the average computation time for 100 graphs with error bars indicating standard deviations (which are distributed as in Fig. \ref{fig:tod} (b)).
Note that the computation time of CIM is evaluated by (number of round trips) $\times$ (10 $\mu$s) as in Sec. \ref{sec:spec}. (b) The computation time of CIM when the DOPO cavity circulation frequency \{10 kHz, 100 kHz, 1 MHz\} and pulse interval \{10 cm (solid line), 1 cm (dashed line), 1 mm (dotted line)\} are changed. Computation time of CIM scales as $O(N)$ if we vary the fiber length proportional to $N$ with fixed pulse interval.
}
\end{figure*}
Figure \ref{fig:time_scaling} (a) shows the computation time versus problem size (number of vertices). The computation time is defined as the CPU time to solve a given MAX-CUT problem in complete graph for GW; as the CPU time to reach the same accuracy as GW for SA, SG3, and BLS; and as the time estimated by the (number of round trips) $\times$ (cavity round trip time) to obtain the same accuracy as GW for CIM. The preparation time needed to input $J_{ij}$ into the computing system, i.e., the graph I/O time, is not included.
For complete graphs of $N \leq 20000$, the CIM exhibits a problem-size independent computation time of less than $10^{-3}\ \mathrm{s}$ if we assume the fixed cavity circulation frequency of 100 kHz and pulse interval of 10 cm.
This means the target accuracy is obtained in the constant number of round trips. It indicate that the computation time of CIM is determined by the turn-on delay time of the DOPO network oscillation, which in turn depends on the round trip time and the pump rate.
In Figure \ref{fig:time_scaling} (b), the computation time of the CIM with different system clock frequency
and pulse spacing are shown (see the Sec. \ref{sec:spec} for the definition).
Since the solutions are obtained in a constant number of cavity round trips, the computation time is pulse spacing independent but linearly depends on the clock frequency, i.e., cavity circulation frequency.
The number of pulses accommodated in the fiber can be changed to vary the pulse spacing under the fixed clock frequency.
On the other hand, when the pulse spacing is fixed and the fiber length is varied, the maximum number of pulses should be increased in proportional to the fiber length.
Going back to the Figure \ref{fig:time_scaling} (a), the time complexity $O(N^{3.5})$ for the GW is dominated by the interior-point method in the Goemans-Williamson algorithm. The SA seems to scale in $O(N^{2})$, which indicates that it requires the number of spin flips to be proportional to $N$ (i.e., constant Monte Carlo steps) to achieve the optimal performance. Each spin flip costs a computation time proportional to the degree $k_i$, where $k_i$ is equal to $N-1$ for all $i \in V$ in the case of complete graphs. Thus, the computation time scales as $O(N \langle k \rangle) = O(N^2)$ for the SA in the complete graphs. Note that CIM and SA didn't always reach the energy obtained by GW for the graph of $N = 40$, half of the 100 runs of stochastic algorithms were post-selected to reach that value. SG3 scales as $O(m) = O(N^2)$ in Fig. \ref{fig:time_scaling} (a), but the values for $N = 40, 160, 800$ are not shown because it didn't reach the accuracy reached by the GW solution. BLS exhibits competitive performance against SA.
Besides, the DOPO amplitudes in CIM evolve as in Fig. \ref{fig:tod} (a) when $N = 800$. The distribution of computation time for 100 randomly weighted complete graphs of $N = 800$ is also shown in Fig. \ref{fig:tod} (b).
\begin{figure*}
\centering
\includegraphics[width=6.1in]{Figure8.eps}
\caption{\label{fig:tod}Detailed data for $N = 800$ randomly weighted complete graphs (appeared in Fig. \ref{fig:energy} (a)). (a) Time evolution of DOPO amplitudes of the graph of $N = 800$.
Only 100 pulses out of $N = 800$ signal pulses are shown. (b)Histogram of computation times for solving 100 randomly weighted complete graphs of size $N=800$ by the CIM, SA, and GW.}
\end{figure*}
\begin{figure}[htbp]
\centering
\includegraphics[width=3.1in]{Figure9.eps}
\caption{\label{fig:gset}Computation time of coherent Ising machine, simulated annealing algorithm, and Goemans-Williamson SDP algorithm on random graphs in G-set instances. The computation time for CIM and SA is defined as the time to reach the same accuracy achieved by GW (without I/O time).
Note that the computation time of CIM is evaluated by (number of round trips) $\times$ (10 $\mu$s) as in Sec. \ref{sec:spec}.}
\end{figure}
Computation time for the random graphs in G-set instances is also studied. Here the subset of graphs in which MAX-CUT problems can be solved in polynomial time (i.e., planar graphs, weakly bipartite graphs, positive weighted graphs without a long odd cycle, and graphs with integer edge weight bounded by $N$ and fixed genus) are excluded.
The execution time of CIM is evaluated under the machine spec described in Sec. \ref{sec:spec} with $p = 0.2$, $\xi = -0.06$, and $\xi_{ij} = \xi w_{ij} / \sqrt{\langle k \rangle}$. Again, the computation time of SA and CIM is the actual time (without graph file I/O) to obtain the same accuracy of solution as GW.
Figure \ref{fig:gset} shows the computation time as functions of the problem size $N$. The computational cost of interior point method dominates the GW algorithm. (Note that G-set contains graphs with both positive and negative edge weights so that we must use the slowest interior point method.)
Then the computation time is almost constant for both SA and CIM. The computation time of SA with constant Monte Carlo step is expected to scale $O(N \langle k \rangle) \sim O(1)$ (here for the random graphs in G-set, $\langle k \rangle \sim O(N^{-1.09}$)). The computation time of CIM here is governed by a turn-on delay time of the DOPO network to reach a steady state oscillation condition, which is constant for varying values of $N$ as mentioned above \cite{takata2012transient}.
\section{\label{sec:outro}Summary and discussion}
The potential for solving NP-hard problems using a CIM was numerically studied by conducting computational experiments using the MAX-CUT problems on sparse G-set graphs and fully connected complete graphs of order up to $2 \times 10^4$. With the normalized pump rate and coupling coefficient $p=0.2$ and $\xi=-0.06$, the CIM achieved a good approximation rate of 0.94148 on average and found better cut compared to the GW for 69 out of 71 graphs in G-set. The computation time for this sparse graph set, including few sessions of hysteretic optimization, is estimated as $50$ ms.
The time scaling was also tested on complete graphs of number of vertices up to $2 \times 10^4$ and number of edges up to $10^8$. The results imply that CIM achieves empirically constant time scaling in a fixed system clock frequency, i.e., the fixed cavity circulation frequency (fiber length),
while SA, SG3, and BLS scale as $O(N^2)$ and GW scales as $O(N^{3.5})$.
Those results suggest that CIM may find applications in high-speed computation for various combinatorial optimization problems, in particular for temporal networks.
The present simulation results do not mean that the CIM can get a reasonably accurate solution by a constant time for arbitrary large problem size. As mentioned already, in the CIM based on a fiber ring resonator, the number of DOPO pulses is determined by the the fiber length and the pulse spacing. In order to implement $2 \times 10^5$ and $2 \times 10^6$ DOPO pulses in the $20\ \mathrm{km}$ fiber ring cavity, we must use a pulse repetition frequency to 2 GHz and 20 GHz, respectively. This is a challenge for both optical components and electronic components of CIM, but certainly within a reach in current technologies.
\begin{acknowledgments}
The authors would like to thank K. Inoue, H. Takesue, K. Aihara, A. Marandi, P. McMahon, T. Leleu, S. Tamate, K. Yan, Z. Wang, and K. Takata for their useful discussions.
This project is supported by the ImPACT program of the Japanese Cabinet Office.
\end{acknowledgments}
|
\section{Introduction}
In this paper we study equilibrium configurations of thin nematic liquid crystalline films. Nematic materials are typically composed of rod- or disk-like molecules and can flow like fluids, yet they retain a degree of molecular orientational order similar to crystalline solids. There are several mathematical frameworks to study the nematics, leading to different, but related variational models. The focus of the present work is on rigorous dimensional reduction of the Landau-de Gennes $Q$-tensor model from three to two dimensions.
We begin by briefly reviewing the basic theory of nematics. The local orientations of molecules in {\em uniaxial nematics} are described by a director---a unit vector in a direction preferred by the molecules at a given point. The director field underlies the Oseen-Frank theory \cite{virga} based on an energy penalizing for spatial variations of the director. This theory incorporates various elastic modes (splay, bend, twist) and interactions with electromagnetic fields and has generally been very successful in predicting equilibrium nematic configurations. However the Oseen-Frank approach is limited in that it prohibits certain types of topological defects, e.g., disclinations, as the constraint that the director must have a unit length is too rigid. A possible remedy was proposed by Ericksen \cite{ericksen} who introduced a scalar parameter intended to describe the quality of local molecular orientational order.
Despite the fact that Ericksen's theory is capable of handling line defects, it still assumes that a preferred direction is specified by the director, excluding a possibility that the nematic can be biaxial. Here a biaxial state differs from a uniaxial state in that it has no rotational symmetry; instead it possesses reflection symmetries with respect to each of three orthogonal axes (only two of which need to be specified). Biaxial configurations are conjectured to exist, e.g., at the core of a nematic defect \cite{kralj2001universal}. Further, certain nematic configurations cannot even be orientable, that is, they cannot be described by a continuous director field \cite{Ball_Zarn}. These deficiencies can be circumvented within the Landau-de Gennes theory in three dimensions that we will discuss in subsequent sections (see also \cite{Ball_Zarn}, \cite{apala_zarnescu_01}, and \cite{Mottram_Newton}). Briefly, the Landau-de Gennes theory is based on the $Q$-tensor order parameter field that is related to the second moment of local orientational probability distribution. The relevant variational model involves minimization of an energy functional consisting of elastic, bulk and surface contributions. Recently there has been considerable activity
on modeling with surface energy terms using both $Q$-tensor and director theories, including for example, \cite{apala_zarnescu_01,canevari13,Lamy14,GM,segatti14,ball2010nematic}.
We are interested in proper reduction of the Landau-de Gennes model to two dimensions in the thin film limit. In this asymptotic regime, surface energy plays a greater role and we take particular care in understanding its influence on the structure of the minimizers of the derived two-dimensional energy. To achieve this goal we employ the tool of $\Gamma$-convergence that has proved successful in tackling problems of dimension reduction in other settings, such as elasticity \cite{anzellotti1994dimension} and Ginzburg-Landau theory \cite{contreras2010gamma}. We work in the domain $\Omega\times(0,h)$ where $0<h\ll1$ and $\Omega\subset\mathbb R^2$ is bounded and Lipschitz. In a subsequent publication we extend this analysis to the case of a small neighborhood of an arbitrary smooth surface, either with or without boundary, as a rigorous analog of the dimension reduction procedure in \cite{Napoli_Vergori} (see also \cite{virga_talk}).
In Section \ref{s:surf} we introduce and analyze the general expression for the surface energy and then, in Section \ref{s:model} combine it with the bulk and elastic terms to form the full non-dimensionalized three dimensional energy functional. In Section \ref{s:conv} we derive the expression for the limiting energy $F_0$ and in Section \ref{s:reg} we analyze minimizers of $F_0$ in different parameter regimes. In all regimes that we consider, it is crucial that the space of competing $Q$-tensors is constrained to accommodate the requirement that each tensor has a normal to the surface of the film as an eigenvector. This condition is forced in the limit by the surface energy and justifies the reduced $Q$-tensor ansatz imposed in \cite{PhysRevLett.59.2582,bauman_phillips_park} in relation to experiments in \cite{PhysRevE.66.030701}.
\section{The $Q$-tensor}
In the three-dimensional setting, one describes the nematic liquid crystal by a $2$-tensor $Q$ which takes the form of a $3\times 3$ symmetric, traceless matrix. Here $Q(x)$ models the second moment of the orientational distribution of the rod-like molecules near $x$. The tensor $Q$ has three real eigenvalues satisfying $\lambda_1+\lambda_2+\lambda_3=0$ and a mutually orthonormal eigenframe $\left\{\mathbf{l},\mathbf{m},\mathbf{n}\right\}$.
Suppose that $\lambda_1=\lambda_2=-\lambda_3/2.$ Then the liquid crystal is in a {\em uniaxial nematic} state and \begin{equation}Q=-\frac{\lambda_3}{2}\mathbf{l}\otimes\mathbf{l}-\frac{\lambda_3}{2}\mathbf{m}\otimes\mathbf{m}+
\lambda_3\mathbf{n}\otimes\mathbf{n}=S\left(\mathbf{n}\otimes\mathbf{n}-\frac{1}{3}\mathbf{I}\right),\label{uniaxial}
\end{equation}
where $S:=\frac{3\lambda_3}{2}$ is the uniaxial nematic order parameter and $\mathbf{n}\in\mathbb{S}^2$ is the nematic director. If there are no repeated eigenvalues, the liquid crystal is in a {\em biaxial nematic} state and
\begin{multline}
Q=\lambda_1\mathbf{l}\otimes\mathbf{l}+\lambda_3\mathbf{n}\otimes\mathbf{n}-\left(\lambda_1+\lambda_3\right)\left(\mathbf{I}-\mathbf{l}\otimes\mathbf{l}-\mathbf{n}\otimes\mathbf{n}\right)\\=S_1\left(\mathbf{l}\otimes\mathbf{l}-\frac{1}{3}\mathbf{I}\right)+S_2\left(\mathbf{n}\otimes\mathbf{n}-\frac{1}{3}\mathbf{I}\right),
\label{biaxial}\end{multline}
where $S_1:=2\lambda_1+\lambda_3$ and $S_2=\lambda_1+2\lambda_3$ are biaxial order parameters.
Note that uniaxiality can also be described in terms of $S_1$ and $S_2$, that is one of the following three cases occurs: $S_1=0$ but $S_2\not=0$, $S_2=0$ but $S_1\not =0$ or $S_1=S_2\not=0.$ When $S_1=S_2=0$ so that ${\bf Q}=0$ the nematic liquid crystal is said to be in an isotropic state associated, for instance, with a high
temperature regime.
From the modeling perspective it turns out that the eigenvalues of $Q$ must satisfy the constraints \cite{ball2010nematic,sonnet2012dissipative}:
\begin{equation}
\label{eq:bnds}
\lambda_i\in[-1/3,2/3],\ \mathrm{for}\ i=1,2,3.
\end{equation}
\section{Surface energy}
\label{s:surf}
In this section we discuss the behavior of the nematic on the boundary of the sample. Here two alternatives are possible. First, the Dirichlet boundary conditions on $Q$ are referred to as strong anchoring conditions in the physics literature: they impose specific preferred orientations on nematic molecules on surfaces bounding the liquid crystal. In the sequel we impose these conditions on the lateral part of the cylindrical sample $\partial\Omega\times(0,h)$. An alternative is to specify the surface energy on the boundary of the sample; then orientations of the molecules on the boundary are determined as a part of the minimization procedure. We adopt this approach, referred to as {\em weak anchoring}, on the top and the bottom surfaces of the film.
We seek a general expression for surface energy that has a family of surface-energy-minimizing tensors with the normal to the surface of the liquid crystal as their eigenvector. The requirement that the normal to the film is also an eigenvector of the $Q$ tensor is motivated by the desire to model both homeotropic and parallel anchoring---corresponding to the nematic molecules oriented perpendicular and parallel to the surface of the film, respectively \cite{virga}. In the former case the uniaxial nematic tensor is prescribed on the boundary (up to the multiplicative order parameter) with the director being perpendicular to the surface of the film. In the latter case, the director orientation is perpendicular to the normal to the film but otherwise may be arbitrary.
Consider the "bare" surface energy (Eq. 7 in \cite{OH}, see also \cite{sen1987landau} and \cite{sluckin1986fluid})
\begin{equation}
\label{bare}
f_s(Q,\nu):=c_1(Q\nu\cdot\nu)+c_2Q\cdot Q+c_3(Q\nu\cdot\nu)^2+c_4{|Q\nu|}^2,
\end{equation}
where $c_i,\ i=1,\ldots,4$ are constants, $A\cdot B=\mathrm{tr}\,\left(B^TA\right)$ for any two $n\times n$ matrices $A$ and $B$, and $\nu\in\mathbb{S}^2$ is a normal to the surface of the liquid crystal. This expression can be supplemented, in principle, by the surface Landau-de Gennes-type expression (Eq. 75 in \cite{Mottram_Newton}) in order to control eigenvalues and to relax constraints on the constants $c_i$ in \eqref{bare} that will be imposed below. This would amount essentially to augmenting \eqref{bare} with an expression of the form \eqref{eq:LdG} below.
Fix $\nu\in\mathbb{S}^1$ and let
\begin{equation}
\label{eq:cala}
\mathcal A:=\left\{Q\in M^{3\times3}_{sym}:\mathrm{tr}\,{Q}=0\right\}.
\end{equation}
We now explore different parameter regimes associated with minimization of the surface energy over the set $\mathcal A$. Some comments along these lines can be found in \cite{fournier2005modeling}.
\begin{theorem}
\label{thm:surf}
The minimum of $f_s(Q,\nu)$ over $Q\in\mathcal{A}$ is achieved in the following five cases as characterized below in terms of the parameters $c_1,\ldots,c_4$:
\begin{enumerate}[(i)]
\item If $\min\{c_2,2c_2+c_4,3c_2+2c_3+2c_4\}>0$, then the minimum of $f_s(Q,\nu)$ is achieved at any $Q$ that is uniaxial and homeotropic with the eigenvalue $-\frac{c_1}{3c_2+2c_3+2c_4}$ corresponding to the eigenvector $\nu$.
\item If $\min\{c_2,2c_2+c_4\}>0$ and $3c_2+2c_3+2c_4=c_1=0$, then the minimum of $f_s(Q,\nu)$ is achieved at any $Q$ that is uniaxial and homeotropic, i.e., with the nematic director parallel to $\nu$.
\item If $\min\{c_2,2c_3-c_2\}>0$ and $2c_2+c_4=0,$ then the minimum of $f_s(Q,\nu)$ is achieved at any $Q$ that satisfies $Q\nu\cdot\nu=\frac{c_1}{c_2-2c_3}$ and has one eigenvector orthogonal to $\nu$ with eigenvalue $\frac{c_1}{4c_3-2c_2}$.
\item If $\min\{2c_2+c_4,c_3+c_4\}>0$ and $c_2=0,$ then the minimum of $f_s(Q,\nu)$ is achieved at any uniaxial or biaxial $Q$ of the form
\begin{equation}
\label{eq:mu}
Q=\mu\mathbf{m}\otimes\mathbf{m}+\left(\frac{c_1}{2(c_3+c_4)}-\mu\right)\mathbf{n}\otimes\mathbf{n}-\frac{c_1}{2(c_3+c_4)}\nu\otimes\nu,
\end{equation}
where $\left\{\mathbf{m}, \mathbf{n}\right\}$ is an arbitrary orthonormal frame in the plane tangent to the surface of the liquid crystal, and $\mu\in\mathbb{R}$ is arbitrary.
\item If $c_3>0$ and $c_2=c_4=0,$ then the minimum of $f_s(Q,\nu)$ is achieved and any minimizing $Q$ must satisfy $Q\nu\cdot\nu=-\frac{c_1}{2c_3}$.
\end{enumerate}
In all other cases, $\inf_{Q\in\mathcal{A}}f_s(Q,\nu)=-\infty$.
\end{theorem}
\begin{proof}
Without loss of generality we may assume that $\nu=(0,0,1):=\hat z$. In order to find $Q\in\mathcal A$ that minimizes $f_s$, observe that
\[Q\cdot Q=2{|Q\nu|}^2-{(Q\nu\cdot\nu)}^2+Q_2\cdot Q_2,\]
where $Q_2\in M^{2\times2}_{sym}$ is a nonzero square block of $\left(\mathbf{I}-\nu\otimes\nu\right)Q\left(\mathbf{I}-\nu\otimes\nu\right).$ Then the expression for $f_s$ can be written as
\begin{equation}
\label{bare_m}
f_s(Q,\nu)=c_1(x\cdot\nu)+c_2Q_2\cdot Q_2+\left(c_3-c_2\right)(x\cdot\nu)^2+\left(2c_2+c_4\right){|x|}^2,
\end{equation}
where $x:=Q\nu\in\mathbb{R}^3$. The traceless condition for $Q$ can be reformulated in terms of $Q_2$ and $x$ as
\begin{equation}
\label{trcond}
\mathrm{tr}\,Q_2+x\cdot\nu=0\,.
\end{equation}
Thus, we are looking for the minimum of \eqref{bare_m} among all $Q_2\in M^{2\times2}_{sym}$ and $x\in\mathbb{R}^3$ that satisfy the condition \eqref{trcond}.
Suppose first that $c_2<0$. Then, setting $x=0$ and observing that
\[Q_2=
\left(
\begin{array}{cc}
\alpha & 0 \\
0 & -\alpha \\
\end{array}
\right)
\]
satisfies the constraint \eqref{trcond} for any $\alpha\in\mathbb{R},$ we conclude that $\inf_{x,Q_2}f_s=-\infty.$ We leave the verification of other parameter regimes which result in $\inf_{x,Q_2}f_s=-\infty$ to the reader and instead concentrate on the five cases laid out in the statement of the theorem.
To this end, let $c_2>0$. Minimizing $f_s$ with respect to $Q_2$ subject to the constraint \eqref{trcond} is equivalent to minimizing $Q_2\cdot Q_2$ subject to \eqref{trcond}. We have
\begin{equation}
\label{eq:1}
2Q_2+\Lambda\mathbf{I}_2=0,
\end{equation}
where $\Lambda$ is the Lagrange multiplier. Combining \eqref{trcond} and \eqref{eq:1}, we find that $\Lambda=(x\cdot\nu)$ and
\begin{equation}
\label{eq:Q2M}
Q_2=-\frac{1}{2}(x\cdot\nu)\mathbf{I}_2.
\end{equation}
Substituting this expression back into \eqref{bare_m}, we have
\[\inf_{\mathcal{A}} f_s=\inf_{\mathbb{R}^3}\tilde{f}_s,\]
where
\begin{equation}
\label{eq:exp1}
\tilde{f}_s(x,\nu):=c_1(x\cdot\nu)+\left(c_3-\frac{c_2}{2}\right)(x\cdot\nu)^2+\left(2c_2+c_4\right){|x|}^2,
\end{equation}
or, equivalently,
\begin{equation}
\label{eq:exp2}
\tilde{f}_s(x,\nu):=c_1(x\cdot\nu)+\left(\frac{3}{2}c_2+c_3+c_4\right)(x\cdot\nu)^2+\left(2c_2+c_4\right){\left|\left({\mathbf I}-\nu\otimes\nu\right)x\right|}^2,
\end{equation}
Observe that if $\min\left\{3c_2+2c_3+2c_4,2c_2+c_4\right\}<0$, then $\inf_{x}\tilde f_s=-\infty$. Suppose now that $\min\left\{3c_2+2c_3+2c_4,2c_2+c_4\right\}>0$ so that there is a critical point $x$ of $\tilde f_s$ that satisfies
\[
\frac{\partial \tilde{f}_s}{\partial x}=c_1\nu+\left(2c_3-c_2\right)(x\cdot\nu)\nu+2\left(2c_2+c_4\right)x=0.
\]
It follows that $x$ is parallel to $\nu$ and
\begin{equation}
\label{eq:3}
x=-\frac{c_1}{3c_2+2c_3+2c_4}\nu
\end{equation}
is a minimum. Consequently,
\begin{equation}
\label{eq:5}
Q\nu=\lambda\nu,
\end{equation}
where $\lambda=-\frac{c_1}{3c_2+2c_3+2c_4}$. Combining \eqref{eq:5} with the expression for $Q_2$, we find the single minimum
\begin{equation}
\label{eq:4}
Q=\frac{3\lambda}{2}\left(\nu\otimes\nu-\frac{1}{3}\mathbf{I}\right),
\end{equation}
of the surface energy corresponding to a fixed uniaxial nematic state with the order parameter $S=\frac{3\lambda}{2}$ and the nematic director $\nu$. This is the case of so-called homeotropic (perpendicular) anchoring. This establishes (i).
Proceeding with the proof of (ii), if $\min\{c_2,2c_2+c_4\}>0$ and $3c_2+2c_3+2c_4=c_1=0$ then the expression \eqref{eq:exp2} for $\tilde f_s$ reduces to its last term. Hence minimization simply requires $x$ to be parallel to $\nu$ meaning that $\nu$ is an eigenvector of $Q$. Further, $Q$ is uniaxial since the corresponding eigenvalue is $x\cdot\nu$ and $Q_2$ is given by \eqref{eq:Q2M}.
Next, if $\min\{c_2,2c_3-c_2\}>0$ and $2c_2+c_4=0,$ then \eqref{eq:exp1} reveals that $\tilde f_s$ is a quadratic function of $x\cdot\nu$ that is minimized when $x\cdot\nu=\frac{c}{c_2-2c_3}=:\sigma$. Combining this observation with \eqref{eq:Q2M} we obtain that the minimizing $Q$ must be of the form
\[
Q=\left(
\begin{array}{ccc}
-\sigma/2 & 0 & q_{13} \\
0 & -\sigma/2 & q_{23} \\
q_{13} & q_{23} & \sigma
\end{array}
\right).
\]
From this we readily see that $\left(-q_{23},q_{13},0\right)$ is an eigenvector of $Q$. This establishes (iii). Note that the minimizing set of $f_s$ consists of a family of biaxial tensors that contains a single uniaxial representative corresponding to a homeotropic boundary condition. Here the biaxial tensors {\it do not} have a normal to the surface of the film as an eigenvector.
Now we pursue the regime $c_2=0$. In this case, from \eqref{bare_m} we see that $f_s$ is independent of $Q_2$. Thus the minimizing matrix $Q_2$ is arbitrary as long as it satisfies the trace constraint \eqref{trcond}. If, in addition, $\min\{2c_2+c_4,c_3+c_4\}>0$ the energy is still minimized by $x$ given by \eqref{eq:3} with $c_2=0$, i.e.,
\[Q\nu=-\frac{c_1}{2(c_3+c_4)}\nu.\]
It also follows that if $\left\{\mathbf{m}, \mathbf{n}\right\}$ is an arbitrary orthonormal frame in the plane tangent to the surface of the liquid crystal, then any
\[
Q=\mu\mathbf{m}\otimes\mathbf{m}+\left(\frac{c_1}{2(c_3+c_4)}-\mu\right)\mathbf{n}\otimes\mathbf{n}-\frac{c_1}{2(c_3+c_4)}\nu\otimes\nu
\]
minimizes the surface energy. This verifies (iv).
Finally in case (v) the expression \eqref{bare_m} reduces to a quadratic function of $x\cdot\nu$ that is minimized at $x\cdot\nu=-c_1/2c_3$. Observe that in this case the minimizing $Q_2$ is again arbitrary up to the trace constraint \eqref{trcond}.
\end{proof}
\begin{remark}
Note that the eigenvalues determined in the cases (i), (iii), and (iv) must respect the bounds \eqref{eq:bnds} on eigenvalues of $Q$ thereby imposing additional restrictions on the parameters $c_1,\ldots,c_4$.
\end{remark}
Having explored all possible parameter regimes we now focus on case (iv) where $\min\{2c_2+c_4,c_3+c_4\}>0$ and $c_2=0.$ Here the degeneracy of the set of tensors minimizing the surface energy $f_s$ provides sufficient freedom for nontrivial reduction to two-dimensional limits that we will carry out in the next section. An alternative approach would be to extend the surface energy by including quartic terms \cite{fournier2005modeling} and even surface derivative terms \cite{Longa}.
By rearranging the terms in \eqref{bare}, the surface energy has the form
\begin{multline}
\label{fs}
f_s(Q,\nu)=c_1(Q\nu\cdot\nu)+c_3(Q\nu\cdot\nu)^2+c_4{|Q\nu|}^2 \\ =\left(c_3+c_4\right)\left[(Q\nu\cdot\nu)+\frac{c_1}{2(c_3+c_4)}\right]^2+c_4{\left|\left(\mathbf{I}-\nu\otimes\nu\right)Q\nu\right|}^2 \\ =\alpha\left[(Q\nu\cdot\nu)-\beta\right]^2+\gamma{\left|\left(\mathbf{I}-\nu\otimes\nu\right)Q\nu\right|}^2,
\end{multline}
up to an additive constant. Here $\alpha=c_3+c_4>0$, $\beta=-\frac{c_1}{2(c_3+c_4)}$, and $\gamma=c_4>0.$ This form of the surface energy explicitly demonstrates that the minimizer has $\nu$ as one eigenvector with corresponding eigenvalue equal to $\beta$.
\section{Landau-de Gennes Energy Functional.\\ Non-dimensionalization}
\label{s:model}
Suppose that the bulk elastic energy density of a nematic liquid crystal is given by
\begin{equation}
\label{elastic}
f_e(\nabla Q):=\frac{L_1}{2}{|\nabla Q|}^2+\frac{L_2}{2}Q_{ij,j}Q_{ik,k}+\frac{L_3}{2}Q_{ik,j}Q_{ij,k}+\frac{L_4}{2}Q_{lk}Q_{ij,k}Q_{ij,l}
\end{equation}
and that the bulk Landau-de Gennes energy density is
\begin{equation}
\label{eq:LdG}
f_{LdG}(Q):=a\,\mathrm{tr}\left(Q^2\right)+\frac{2b}{3}\mathrm{tr}\left(Q^3\right)+\frac{c}{2}\left(\mathrm{tr}\left(Q^2\right)\right)^2,
\end{equation}
cf. \cite{Mottram_Newton}. Here the coefficient $a$ is temperature-dependent and in particular is negative for sufficiently low temperatures, and $c>0$.
One readily checks that the form \eqref{eq:LdG} of this potential implies that in fact $f_{LdG}$ depends only on the eigenvalues of $Q$, and due to the trace-free condition, therefore
depends only on two eigenvalues. Equivalently, one can view $f_{LdG}$ as a function of the two degrees of orientation $S_1$ and $S_2$ appearing in \eqref{biaxial}. Furthermore,
its form guarantees that the isotropic state $Q\equiv 0$ (or equivalently $S_1=S_2=0$) yields a global minimum at high temperatures while a uniaxial state of the form \eqref{uniaxial} where either $S_1=0,\;S_2=0$ or $S_1=S_2$ gives the minimum when temperature (i.e. the parameter $a$) is reduced below a certain critical value, cf. \cite{apala_zarnescu_01,Mottram_Newton}. In this paper we fix the temperature to be low enough so that the minimizers of $f_{LdG}$ are uniaxial. We also remark for future use that $f_{LdG}$ is bounded from below and can be made nonnegative by adding an appropriate constant. In light of this, we will henceforth assume a minimum value of zero for $f_{LdG}.$
\begin{figure}[htb]
\centering
\includegraphics[height=2in]{diagram.eps}
\caption{Geometry of the problem.}
\label{fig:1}
\end{figure}
Let $\Omega\subset\mathbb{R}^2$ be a bounded domain with a Lipschitz boundary and let $h>0$ be given (Figure \ref{fig:1}). Assume that the energy functional is
\begin{equation}
\label{energy}
E[Q]:=\int_{\Omega\times(0,h)}\left\{f_e(\nabla Q)+f_{LdG}(Q)\right\}\,dV+\int_{\Omega\times\{0,h\}}f_s(Q,\hat{z})\,dA\,,
\end{equation}
where $\hat{z}$ is a unit vector normal to the surface of the film. Given uniaxial data $g\in H^{1/2}(\partial\Omega;\mathcal{A})$ we prescribe the lateral boundary condition of the form
\begin{equation}
\label{eq:bd}
Q(x,y,z)=g(x,y)\ \mathrm{for}\ (x,y)\in\partial\Omega\ \mathrm{and}\ z\in (0,h).
\end{equation}
The admissible class of tensor-valued functions is then \[\mathcal{C}_h^g:=\left\{Q\in H^1\left(\Omega\times(0,h);\mathcal{A}\right):Q|_{\partial\Omega\times\{z\}}=g,\forall z\in(0,h)\right\},\] where $\mathcal{A}$ is the set of three-by-three symmetric traceless matrices defined in \eqref{eq:cala}. Throughout this work we assume that $g$ is uniaxial and is taken so that $\mathcal{C}_h^g$ is nonempty.
It has been shown, however, in \cite{ball2010nematic} that when $L_4\neq0$ minimizers of \eqref{energy} may fail to exist. On the other hand, $L_4=0$ precludes an appropriate reduction to the general form of the Oseen-Frank energy for nematics \cite{Mottram_Newton}. Since we are interested in a characterization of minimizers, we will set $L_4=0$.
We nondimensionalize the problem by scaling the spatial coordinates
\[\tilde{x}=\frac{x}{D},\ \tilde{y}=\frac{y}{D},\ \tilde{z}=\frac{z}{h},\ \]
where $D:=\mathrm{diam}(\Omega)$. Set $\xi=\frac{L_1}{2D^2},$ $\epsilon=\frac{h}{D}$ and introduce $\tilde{f}_e(Q,\nabla Q):=\frac{1}{\xi}f_e(Q,\nabla Q).$ Dropping tildes, we obtain
\begin{multline}
f_e(\nabla Q): =\left[Q_{im,j}Q_{im,j}+M_2Q_{ik,k}Q_{ij,j}+M_3Q_{ij,k}Q_{ik,j}\right]\\
+\frac{2}{\epsilon}\left[M_2Q_{ij,j}Q_{i3,3}+M_3Q_{i3,j}Q_{ij,3}\right] \\
+\frac{1}{\epsilon^2}\left[Q_{im,3}Q_{im,3}+(M_2+M_3)Q_{i3,3}Q_{i3,3}\right],
\end{multline}
where $M_2=\frac{L_2}{L_1},$ $M_3=\frac{L_3}{L_1},$ the indices $i,m=1,2,3,$ and $j,k,l=1,2.$ Rescaling the Landau-de Gennes potential via $\tilde{f}_{LdG}(Q):=\frac{1}{w_l\xi}f_{LdG}(Q)$ and dropping tildes gives
\begin{equation}
f_{LdG}(Q)=2A\,\mathrm{tr}\left(Q^2\right)+\frac{4}{3}B\,\mathrm{tr}\left(Q^3\right)+{\left(\mathrm{tr}\left(Q^2\right)\right)}^2,
\end{equation}
where $A=\frac{a}{c},$ $B=\frac{b}{c},$ and $w_l=\frac{c}{2\xi}.$ Letting $\tilde\alpha=\frac{\alpha}{\xi},\ \tilde{\gamma}=\frac{\gamma}{\xi}$, setting
\[\tilde{f}_s(Q,\hat z):=\frac{1}{{\tilde\alpha}^2D\xi}f_s(Q,\hat z),\]
and dropping tildes, the expression for the nondimensionalized surface energy is
\begin{equation}
\label{eq:baren}
f_s(Q,\hat z)=\alpha\left[(Q\hat z\cdot\hat z)-\beta\right]^2+\gamma{\left|\left(\mathbf{I}-\hat z\otimes\hat z\right)Q\hat z\right|}^2.
\end{equation}
We conclude that the total energy is
\begin{equation}
\label{eq:8}
E[Q]=\xi D^2h\int_{\Omega\times(0,1)}\left(f_e(\nabla Q)+w_lf_{LdG}(Q)\right)\,dV+\xi D^3\int_{\Omega\times\{0,1\}}f_s(Q,\hat z)\,dA,
\end{equation}
where the rescaled domain is denoted by the same letter $\Omega$.
Introducing the non-dimensional energy $F_\epsilon[Q]:=\frac{2}{L_1h}E[Q]$, we find using \eqref{eq:8} that
\begin{equation}
\label{nden}
F_\epsilon[Q]=\int_{\Omega\times(0,1)}\left(f_e(\nabla Q)+w_lf_{LdG}(Q)\right)\,dV+\frac{1}{\epsilon}\int_{\Omega\times\{0,1\}}f_s(Q,\hat z)\,dA.
\end{equation}
\section{Convergence of minimizers of $F_\epsilon[Q]$ when $\epsilon\to 0$}
\label{s:conv}
Assume that an appropriate constant has been added to the Landau-de Gennes energy to guarantee that $F_\epsilon[Q]\geq 0$. We wish to consider a range of asymptotic regimes corresponding to different magnitudes of $\alpha$ and $\gamma$. To this end, we will assume that $\alpha=\alpha_0+\varepsilon\alpha_1$ and $\gamma=\gamma_0+\varepsilon\gamma_1$ for some nonnegative constants $\alpha_0, \alpha_1,\gamma_0,\gamma_1$. Then the surface energy density \eqref{eq:baren} can be written as
\begin{equation}
\label{eq:se}
f_s(Q,\hat z)=f_s^{(0)}(Q,\hat z)+\varepsilon f_s^{(1)}(Q,\hat z),
\end{equation}
where
\begin{equation}
\label{eq:fso}
f_s^{(0)}:=\alpha_0\left[(Q\hat z\cdot\hat z)-\beta\right]^2+\gamma_0{\left|\left(\mathbf{I}-\hat z\otimes\hat z\right)Q\hat z\right|}^2,
\end{equation}
and
\begin{equation}
\label{eq:fsi}
f_s^{(1)}:=\alpha_1\left[(Q\hat z\cdot\hat z)-\beta\right]^2+\gamma_1{\left|\left(\mathbf{I}-\hat z\otimes\hat z\right)Q\hat z\right|}^2.
\end{equation}
As it will become evident below, we can assume that $\alpha_0\alpha_1=\gamma_0\gamma_1=0$.
Let
\begin{equation}
\label{eq:f0}
F_0[Q]:=\left\{
\begin{array}{ll}
\int_{\Omega}\left\{f_{e}^0(\nabla Q)+w_lf_{LdG}(Q)+2f_s^{(1)}(Q,\hat z)\right\}\,dA & \mbox{ if }Q\in H^1_g, \\
+\infty & \mbox{ otherwise. }
\end{array}
\right.
\end{equation}
Here \[f_{e}^0(\nabla Q)=Q_{im,j}Q_{im,j}+M_2Q_{ij,j}Q_{ik,k}+M_3Q_{ik,j}Q_{ij,k},\] the space \[H^1_g:=\left\{Q\in H^1(\Omega;\mathcal{D}):Q|_{\partial\Omega}=g\right\}\] and \[\mathcal{D}:=\left\{Q\in \mathcal{A}:f_s^{(0)}(Q)=0\right\},\] for some boundary data $g:\partial\Omega\to\mathcal{D}.$
We now state our main theorem on dimension reduction via $\Gamma$-convergence. For those unfamiliar with the notion, we refer, for example, to \cite{DalMaso}. Note that whenever necessary we will view $H^1_g$ as a subset of $\mathcal{C}^g_1$ via a trivial extension to three dimensions.
\begin{theorem}
\label{t1}
Fix $g:\partial\Omega\to\mathcal{D}$ such that $H^1_g$ is nonempty. Assume that $-1< M_3<2$, and $-\frac{3}{5}-\frac{1}{10}M_3< M_2$. Then $\Gamma$-$\lim_\varepsilon{F_\varepsilon}=F_0$ weakly in $\mathcal{C}_1^g$. Furthermore, if a sequence $\left\{Q_\varepsilon\right\}_{\varepsilon>0}\subset\mathcal{C}_1^g$ satisfies a uniform energy bound $F_\varepsilon[Q_\varepsilon]<C_0$ then there is a subsequence weakly convergent in $\mathcal{C}_1^g$ to a map in $H^1_g$.
\end{theorem}
\begin{proof}First, we demonstrate that one can always choose a trivial recovery sequence. Indeed, if $Q_\varepsilon\equiv Q\in\mathcal{C}_1^g\backslash H^1_g$ then $\lim_{\varepsilon\to0}F_\varepsilon[Q_\varepsilon]=+\infty=F_0[Q]$ and when $Q_\varepsilon\equiv Q\in H^1_g$ then $F_\varepsilon[Q_\varepsilon]=F_0[Q_\varepsilon]=F_0[Q]$ for all $\varepsilon$.
For the lower semicontinuity part of the proof, consider an arbitrary sequence $\left\{Q_\varepsilon\right\}_{\varepsilon>0}\subset\mathcal{C}_1^g$ converging weakly in $\mathcal{C}_1^g$ to some $Q_0\in H^1_g.$ It has been established in \cite{Gartland_Davis} (Lemma 4.2) and \cite{Longa} that when the elastic constants satisfy the conditions $-1< M_3<2$, and $-\frac{3}{5}-\frac{1}{10}M_3< M_2$ the integral of $f_e$ is weakly lower semicontinuous in $H^1(\Omega\times(0,1))$ and, in fact,
\begin{equation}
\label{eq:coer}
f_e(\nabla Q)\geq C{|\nabla Q|}^2
\end{equation}
pointwise for all admissible $Q$, where $C>0$. Then using Sobolev embedding, one finds
\[\liminf_{\varepsilon\to0}F_\varepsilon[Q_\varepsilon]\geq F_0[Q_0].\]
On the other hand, if $\left\{Q_\varepsilon\right\}_{\varepsilon>0}\subset\mathcal{C}_1^g$ converges weakly in $\mathcal{C}_1^g$ to some $Q_0\in \mathcal{C}_1^g\backslash H^1_g,$ then $Q_0$ depends on $z$ and/or $Q_0$ is not $\mathcal{D}$-valued. In the first case, invoking \eqref{eq:coer}, we have
\[\liminf_{\varepsilon\to0}F_\varepsilon[Q_\varepsilon]\geq C\liminf_{\varepsilon\to0}\frac{1}{\epsilon^2}\int_{\Omega\times(0,1)}{|Q_{\varepsilon,z}|}^2\,dV=+\infty.\]
In the second case,
\[\int_{\Omega\times\{0,1\}}f_s^{(0)}(Q_0,\hat z)\,dA\neq0,\]
(cf. \eqref{fs}), thus by strong $L^2$-convergence of traces on $\Omega\times\{0,1\}$,
\[\lim_{\varepsilon\to0}{\frac{1}{\varepsilon}\int_{\Omega\times\{0,1\}}f_s^{(0)}(Q_\varepsilon,\hat z)\,dA=+\infty.}\]
Finally, since the uniform energy bound implies a uniform $H^1$-bound, there exists a subsequence weakly convergent in $H^1(\Omega\times(0,1);\mathcal{A})$ to a limit $Q_0$ that is independent of $z$. Further, strong convergence of traces in $L^2$ implies that $Q_0\in H^1_g$.
\end{proof}
\begin{remark}
{\em When $M_2=M_3=0$, one can easily argue that the convergence of the subsequence is, in fact, strong. Indeed, $F_\varepsilon[Q]=F_0[Q]$ for every $Q\in H^1_g$, hence $F_\varepsilon[Q_\varepsilon]\leq F_0[Q_0]$ for all $\varepsilon>0$. Since
\begin{multline*}
\int_{\Omega\times(0,1)}f_{LdG}(Q_\varepsilon)\,dV+\int_{\Omega\times\{0,1\}}f_s^{(1)}(Q_\varepsilon,\hat z)\,dA \\ \to\int_{\Omega}\left(f_{LdG}(Q_0)+2f_s^{(1)}(Q_0,\hat z)\right)\,dA\mbox{ as }\varepsilon\to0
\end{multline*}
we have
\[\limsup_{\varepsilon\to0}\int_{\Omega\times(0,1)}\left({|Q_{\varepsilon,x}|}^2+{|Q_{\varepsilon,y}|}^2+\frac{1}{\epsilon^2}{|Q_{\varepsilon,z}|}^2\right)\,dV\leq\int_{\Omega}\left({|Q_{0,x}|}^2+{|Q_{0,y}|}^2\right)\,dA.\]
Combining this with the lower semicontinuity of the $L^2$-norm of the derivative, strong convergence in $\mathcal{C}_1^g$ follows.}
\end{remark}
\section{Minimizers of the $\Gamma$-limit in different regimes}
\label{s:reg}
In this section we explore minimization of the $\Gamma$-limit for different parameter regimes. The added penalty terms that originate from the three-dimensional surface term have a potential to disconnect uniaxial states whenever the Landau-de Gennes term becomes dominant. As we will show, this may result in formation of singular structures such as boundary layers of Allen-Cahn-type or vortices of Ginzburg-Landau-type with possible emergence of biaxiality (cf. \cite{biscari1997topological}).
In order to apply Theorem \ref{t1}, we note first that a minimizer $Q_\varepsilon$ of $F_\varepsilon$ exists for every $\varepsilon>0$ by the direct method. Furthermore, selecting any $G\in H^1_g$ we observe that $F_\varepsilon[Q_\varepsilon]\leq F_\varepsilon[G]=F_0[G]$ and so we can extract a weakly convergent subsequence $\{Q_\varepsilon\}_{\varepsilon>0}$ such that $Q_\varepsilon\rightharpoonup Q_0$ for some $Q_0$ in $H^1_g$. Since by the basic properties of $\Gamma$-convergence any limit of minimizers is a minimizer of the limiting functional, we conclude that $Q_0$ minimizes $F_0$.
We will assume throughout this section that $\Omega$ is simply connected with a sufficiently smooth boundary. In what follows we will take $\gamma>0$ to be independent of $\varepsilon$ (i.e., $\gamma_1=0$ in \eqref{eq:fsi}) and consider two distinct parameter regimes for $\alpha$ under several boundary conditions.
\subsection{Regime $f_s^{(1)}\equiv0$}
\label{regime1}
First, let $\alpha_1=0$ so that $f_s^{(1)}\equiv0$ and $f_s^{(0)}$ given by \eqref{eq:fso}. Then $\mathcal{D}$-valued maps have $\hat z$ as one eigenvector with corresponding eigenvalue $\beta$. Note that these maps are not necessarily uniaxial, although they are required to be uniaxial on the boundary.
There are two types of uniaxial $\mathcal D$-valued maps: those corresponding to $\mu=-\beta/2$ and $\mu=\beta$ in \eqref{eq:mu}, respectively. When $\mu=-\beta/2$ we have
\[Q=\frac{3\beta}{2}\left(\hat z\otimes\hat z-\frac{1}{3}\mathbf{I}\right)\]
and the uniaxial Dirichlet condition is completely rigid as $Q$ is equal to a constant. Alternatively, when $\mu=\beta$, one finds that \[Q=-3\beta\left(\mathbf{n}\otimes\mathbf{n}-\frac{1}{3}\mathbf{I}\right),\]
where $\mathbf{n}$ is an arbitrary unit vector field on the plane and one has the freedom to choose uniaxial Dirichlet boundary data of any degree. Here by the degree of $Q$ we understand the winding number of the planar $\mathbb S^1$-valued vector field
\(e^{2i\psi}\) on $\partial\Omega$ where $\mathbf n=e^{i\psi}$. Note that in this definition, it is assumed that $Q\in H^{1/2}(\partial\Omega)$ but this does not preclude the possibility of the phase $\psi$ of $\mathbf n$ jumping by an odd multiple of $\pi$ after one circulation around $\partial\Omega$. When this happens, the vector field $\mathbf n$ is discontinuous, but the field $e^{2i\psi}$ is smooth.
\medskip
\noindent{{\bf Case 1.} We begin by characterizing $Q_0$ in the first, topologically simpler case
\begin{equation}
\label{dobc}
Q|_{\partial\Omega\times(0,1)}=g:=\frac{3}{2}\beta\left(\hat z\otimes\hat z-\frac{1}{3}\mathbf{I}\right).
\end{equation}
Unless specified otherwise, we find it preferable from this point on to use the following representation of $Q\in H^1_g$ invoked, for example, in \cite{bauman_phillips_park} and motivated by simulations in \cite{PhysRevLett.59.2582}:
\begin{equation}
\label{eq:pr}
Q=\left(
\begin{array}{ccc}
p_1-\frac{\beta}{2} & p_2 & 0 \\
p_2 & -p_1-\frac{\beta}{2} & 0 \\
0 & 0 & \beta
\end{array}
\right).
\end{equation}
It is a convenient change of variables in the setting when one eigenvector of the $Q$-tensor is parallel to the $z$-axis. Note that the boundary condition in the new coordinates is
\begin{equation}
\label{eq:bcpq}
\mathbf{p}|_{\partial\Omega}=(p_1,p_2)|_{\partial\Omega}=\mathbf{0}.
\end{equation}
Applying the identity
\begin{multline}
\label{eq:bpp}
f_e^0(Q)=\frac{1}{2}\left(2+M_2+M_3\right)|\nabla {\mathbf p}|^2+\frac{1}{8}\left(6+M_2+M_3\right)|\nabla\beta|^2 \\
+\frac{M_2+M_3}{2}\left(p_{1x}\beta_x-p_{1y}\beta_y+p_{2y}\beta_x+p_{2x}\beta_y\right)+\left|M_2+M_3\right|\left(p_{1x}p_{2y}-p_{1y}p_{2x}\right)
\end{multline}
from \cite{bauman_phillips_park}, the expression for $F_0[Q]$ takes the form
\begin{equation}
\label{eq:f0pq}
\frac{1}{M}F_0[Q]=\tilde F_0[\mathbf{p}]:=\int_{\Omega}\left\{\frac{1}{2}{\left|\nabla\mathbf{p}\right|}^2+\frac{1}{\delta^2}W(|\mathbf{p}|)\right\}\,dV,
\end{equation}
where we have used that $\beta$ is constant and that the integral of the Jacobian of $\mathbf p$ vanishes due to \eqref{eq:bcpq}. Here $\mathbf{p}=\left(p_1,p_2\right),$ the parameters $M=2+M_2+M_3$, $\delta=\sqrt{M/w_l}$, and
\begin{equation}
\label{eq:tc}
W(t)=4t^4+\tilde{C}t^2+\tilde{D}
\end{equation}
with $\tilde{C}=6\beta^2-4B\beta+4A$ and $\tilde{D}\in\mathbb{R}$. Note that $\tilde C$, in particular, varies with temperature through its dependence on the coefficient $A$. It is plausible to assume that $\tilde C$ may change its sign in appropriate circumstances.
One easily observes that the minimizer of \eqref{eq:f0pq}, subject to the boundary condition \eqref{eq:bcpq} has a constant phase and so \eqref{eq:f0pq} reduces to a scalar minimization problem for the modulus $p:=|\mathbf{p}|$. The minimizers of \eqref{eq:f0pq} satisfy the Allen-Cahn type equation
\begin{equation}
\label{eq:whatever}
-\Delta p+\frac{1}{\delta^2}W^\prime(p)=0\mbox{ in }\Omega,\quad p=0\mbox{ on }\partial\Omega.
\end{equation}
The function $p\equiv0$ always solves this problem and, in fact, is the unique critical point and thus the minimizer if $\tilde{C}\geq0$. Therefore the minimizing $Q$-tensor in this regime corresponds to a constant uniaxial state
\begin{equation}
\label{eq:qbeta}
Q_\beta=\left(
\begin{array}{ccc}
-\frac{\beta}{2} & 0 & 0 \\
0 & -\frac{\beta}{2} & 0 \\
0 & 0 & \beta
\end{array}
\right).
\end{equation}
(cf. \eqref{eq:pr}) matching the boundary condition \eqref{dobc}. When $\tilde{C}<0$, however, the constant state $Q\equiv Q_\beta$ looses stability once $\lambda_1(\Omega)$ exceeds $-2\tilde{C}/\delta^2$, where $\lambda_1(\Omega)$ is the first eigenvalue of the Laplacian. A minimizing nonconstant solution emerges in this parameter regime when the value of $p$ enforced on $\partial\Omega$ by the surface energy does not minimize the bulk Landau-de Gennes energy. Indeed we expect a boundary layer to form in the vicinity of $\partial\Omega$, bridging $p=0$ to the minimum value of $W$ in the bulk when $\delta\ll1$ see e.g. \cite{Berger_70}.
Further, the nonconstant minimizing configuration cannot be uniaxial everywhere in $\Omega$ as can be seen from the measure of biaxiality introduced in \cite{kaiser1992stability}:
\begin{equation}
\label{eq:bia}
\xi(Q)^2:=1-6\frac{{\left(\mathrm{tr}\,{Q}^3\right)}^2}{{\left(\mathrm{tr}\,{Q}^2\right)}^3},
\end{equation}
where $\xi(Q)=0$ implies that $Q$ is uniaxial. If we express $\xi$ in terms of $p$ and $\beta$, we have
\begin{equation}
\label{eq:biap}
\xi(p,\beta)=1-27\,\frac{\beta^2\left(4p^2-\beta^2\right)^2}{(4p^2+3\beta^2)^3}.
\end{equation}
Since $\beta$ is fixed, there are finitely many values of $p$ where $\xi(p,\beta)$ vanishes. We conclude that if $p$ is a nontrivial solution of the elliptic boundary value problem \eqref{eq:whatever} then the minimizer is necessarily biaxial almost everywhere.
\medskip
\noindent{{\bf Case 2.} Now we turn to the case of the boundary condition
\begin{equation}
\label{nbc}
Q|_{\partial\Omega\times(0,1)}=g:=-3\beta\left(\mathbf{n}\otimes\mathbf{n} -\frac{1}{3}\mathbf{I}\right),
\end{equation}
where $\mathbf{n}:\partial\Omega\to\mathbb{S}^1$ is such that $\mathbf{n}\otimes\mathbf{n}$ is smooth and may have a nonzero degree in the sense described above. Then the tensor $Q_0$ minimizes the energy $F_0$ given by \eqref{eq:f0pq}, subject to the boundary condition
\begin{equation}
\label{nbc_0}
Q|_{\partial\Omega}=-3\beta\left(\mathbf{n}\otimes\mathbf{n} -\frac{1}{3}\mathbf{I}\right).
\end{equation}
Using the representation \eqref{eq:pr}, we have that $Q_0$ can be represented by the vector $\mathbf{p}_0$ that minimizes $\tilde{F}_0$ and satisfies
\begin{equation}
\label{eq:pbc}
\mathbf{p}=-3\beta\left(n_1^2-\frac{1}{2},n_1n_2\right)
\end{equation}
on $\partial\Omega$ where $|\mathbf{p}|=\frac{3\beta}{2}$. In fact, $\tilde{F}_0$ is altered by an additive constant depending on $g$ due to the presence of the null Lagrangian $p_{1x}p_{2y}-p_{1y}p_{2x}$ in \eqref{eq:bpp} that will not affect minimization.
In order to better understand the behavior of $Q_0$ let $\delta$ be small. Then, in general, when $\tilde{C}$ in \eqref{eq:tc} is negative, the corresponding variational problem is of Ginzburg-Landau-type, but with a boundary layer bridging the equilibrium value $\sqrt{-\tilde C/8}$ of $p$ in the bulk to that of $3\beta/2$ enforced by the surface energy on $\partial\Omega$. Furthermore, for topologically nontrivial boundary data for $\mathbf{p}$, the minimizing vector field $\mathbf{p}_0$ has to vanish somewhere within a vortex core structure of a characteristic size $\delta$ in $\Omega$. Recalling \eqref{eq:pr}, we have that $\mathbf{p}_0(x)=\mathbf{0}$ corresponds to a uniaxial state at $x$ with the director pointing along $\hat z$, namely
\[Q_0(x)=\frac{3}{2}\beta\left(\hat z\otimes\hat z-\frac{1}{3}\mathbf{I}\right).\]
In the case of $\tilde{C}\geq0$---making $W$ convex with a unique minimum at $p=0$---we expect again that a boundary layer would form along $\partial\Omega$ bridging the boundary value of $p$, namely $3\beta/2$, to zero in the interior of $\Omega$. In all cases discussed in this section, we expect that symmetry breaking caused by surface energy will induce biaxiality within small sets where $Q_0$ experiences large variations.
\begin{remark}
It is also possible to consider an intermediate asymptotic regime with non-zero degree Dirichlet data where minimality does not force biaxiality. Suppose the surface energy is taken to be substantially smaller than in those cases considered so far but where the relative strength of the Landau-de Gennes contribution is stronger. Suppose, for example,
that $\gamma=\varepsilon\gamma_1$ and $\alpha_0=\alpha_1=0$ for some $\gamma_1>0$ in expressions \eqref{eq:fso}, \eqref{eq:fsi} while say $w_l=\frac{w_0}{\varepsilon}$ for some $w_0>0.$ In view of \eqref{nden} one
sees that the new $\Gamma$-limit would now take the form
\begin{equation}
\label{eq:f0r}
F_0[Q]:=\left\{
\begin{array}{ll}
\int_{\Omega}\left\{f_{e}^0(\nabla Q)+2\gamma_0{\left|\left(\mathbf{I}-\hat z\otimes\hat z\right)Q\hat z\right|}^2\right\}\,dA & \mbox{ if }Q\in \tilde{H}^1_g, \\
+\infty & \mbox{ otherwise. }
\end{array}
\right.
\end{equation}
where $\tilde{H}^1_g:=\{Q\in H^1(\Omega;\mathcal{A}):\;Q|_{\partial\Omega}=g,\;f_{LdG}(Q)=0\}$.
Here we assume the uniaxial data $g$ takes the form
\[
g=S^*\left(\mathbf{n}\otimes\mathbf{n} -\frac{1}{3}\mathbf{I}\right)
\] for some planar vector field $\mathbf{n}$ where $S^*$
corresponds to the preferred value dictated by the requirement $f_{LdG}(g)=0$ and suppose $g$ has even degree so as to ensure that $\tilde{H}^1_g$ is nonempty. In this case,
say for $\gamma_0$ large, one expects the uniaxial minimizer to simply undergo an out of plane rotation of its director in a neighborhood of the boundary, keeping the degree of orientation fixed, thereby
smoothly bridging the boundary value ${\bf n}$ to $\hat{z}$ inside the domain so as to accommodate the cost of the remnant of the surface energy term residing in the $\Gamma$-limit.
\end{remark}
\subsection{Regime $f_s^{(1)}(Q,\hat z)=\alpha_1\left[(Q\hat z\cdot\hat z)-\beta\right]^2$}
Now consider the case when $\alpha=\varepsilon \alpha_1$ and $\gamma=\gamma_0$ for some $\gamma_0,\alpha_1>0$. Then $f_s^{(0)}(Q,\hat z)=\gamma_0{\left|\left(\mathbf{I}-\hat z\otimes\hat z\right)Q\hat z\right|}^2$ and $f_s^{(1)}(Q,\hat z)=\alpha_1\left[(Q\hat z\cdot\hat z)-\beta\right]^2$ and the set $\mathcal{D}$ consists of traceless symmetric tensors having $\hat z$ as one of its eigenvectors. When $\alpha_1=0$ the asymptotic behavior of the limiting functional $F_0$ was considered in \cite{bauman_phillips_park} as $\delta\to0$. They characterize the minimizer by identifying finitely many interacting vortices whose positions are determined via minimization of a renormalized energy of the type introduced in \cite{BBH} for the Ginzburg-Landau functional. This description is very similar to the already discussed regime above which formally corresponds to $\alpha_1=\infty$, i.e., an eigenvalue corresponding to $\hat z$ is identically equal to $\beta$ in $\Omega$. Hence we expect similar behavior for any finite value of $\alpha_1$.
\section{Structure of the singular region for topologically nontrivial boundary data}
In this section we provide further insight into the structure of the singular region that develops for topologically nontrivial boundary data in Case 2 of Section \ref{regime1} when $\tilde C<0$.
Suppose for the moment that the dimensional reduction was carried out as in Theorem \ref{t1} for a sequence of functionals without the surface energy term. In this case, the conclusion of the theorem would remain the same, with the exception that no additional constraints would have to be imposed on the target space of admissible tensors for the limiting problem. In particular, any tensor field described by a field of symmetric traceless $3\times3$ matrices in $H^1(\Omega)$ satisfying the appropriate boundary conditions on $\partial\Omega$ would be admissible.
Now recall that any tensor representing a uniaxial nematic as in \eqref{uniaxial} can be associated with a point on $\mathbb{RP}^2$, i.e., on $\mathbb{S}^2$ with antipodal points identified (since ${\bf n}$ and $-{\bf n}$ lead to the same $Q$-tensor). This means that the tensor field
$Q|_{\partial\Omega}$ corresponding to the $\mathbb{S}^1$-valued boundary data \eqref{nbc_0} can be associated with one or more semi-circular arcs along the equator of $\mathbb{S}^2$. If the degree of ${\bf p}$ in \eqref{eq:pbc} is an even integer, then in terms of the image in $\mathbb{RP}^2$ this corresponds to an even number of such half-equators. One can then smoothly deform the $Q$-tensor within $\Omega$ through a field of uniaxial nematics as indicated in Figure \ref{fig:even} to, say, a point in the interior of $\Omega$ where the director points either north or south. That is, the even number of semi-circles originally overlapping on the equator can gradually migrate towards the poles as a family of closed curves on $\mathbb{RP}^2$.
Note that a sufficiently smooth uniaxial nematic tensor field can be described by a smooth director field \cite{Ball_Zarn} with the boundary data of degree equal to the half of that for $Q|_{\partial\Omega}$. Hence the uniaxial nematic tensor field can be constructed even if the degree of the corresponding director field on the boundary is non-zero. Putting this another way, the director is $\mathbb{S}^2$-valued in $\Omega$ and its $\mathbb{S}^1$-valued topologically nontrivial restriction to $\partial\Omega$ can be bridged without forming a singularity in the interior of $\Omega$ by the director ``escaping into the third dimension".
\begin{figure}[htb]
\centering
\includegraphics[height=2.5in]{split_diagram_even.eps}
\caption{A uniaxial nematic tensor $Q$ with the director ${\bf n}$ can be identified with a projection matrix $\mathrm{P}={\bf n}\otimes{\bf n}$, up to a translation and dilation. Let $\Omega=B_1(0)\subset\mathbb{R}^2$ and let $s$ be some parametrization of a circle of radius $r\in(0,1]$. (a) The boundary data ${\bf n}|_{\partial\Omega}=e^{is}$ corresponds to an equator in $\mathbb{S}^2$; (b) The boundary data for $\mathrm{P}$ corresponds to two half-equators in $\mathbb{S}^2$; (c) For $0<r_0<1$, the half-circles migrate toward the north and south poles of $\mathbb{S}^2$, respectively; (d) At $r=0$ half-circles contract to the poles. This corresponds to the director pointing up or down at $r=0$.}
\label{fig:even}
\end{figure}
If the degree of ${\bf p}$ in \eqref{eq:pbc} is an odd integer, then the image of $\partial\Omega$ in $\mathbb{RP}^2$ corresponds to an odd number of half-equators of $\mathbb{S}^2$. This curve is not contractible in $\mathbb{RP}^2$, i.e., it cannot be smoothly deformed into a pair of points in $\mathbb{S}^2$. As Figure \ref{fig:odd} illustrates, the shortest smooth deformation of an odd number of half-equators of $\mathbb{S}^2$ is a single half-equator in $\mathbb{S}^2$, that is a closed geodesic in $\mathbb{RP}^2$. Then a uniaxial nematic $Q$-tensor field with non-contractible boundary data on $\partial\Omega$ cannot be smooth everywhere in $\Omega$ and must have a singularity at some $x\in\Omega$ as $Q$ will have to assume all values on a single half-equator in $\mathbb{S}^2$ in an arbitrarily small neighborhood of $x$.
\begin{figure}[htb]
\centering
\includegraphics[height=2.5in]{split_diagram_odd.eps}
\caption{Behavior of $Q$ for the boundary data of degree $3$. The setup is the same as in the caption to Figure \ref{fig:even} and $0<r_1\ll1$.}
\label{fig:odd}
\end{figure}
Staying with this broader class of admissible tensors for the moment, if we now let $w_l$ in \eqref{eq:f0} be large but finite (this is equivalent to setting $\delta$ in \eqref{eq:f0pq} small), any deviation of $Q$ from a uniaxial nematic state--a minimum of $f_{LdG}$-- will result in large energy penalty. The smooth energy-minimizing tensor field will then be approximately uniaxial nematic everywhere. That is, as described above, if the degree of $Q|_{\partial\Omega}$ is even then there is no topological obstruction preventing $Q_0$ from being almost uniaxial nematic everywhere in $\Omega$. Since $Q_0$ has a fixed set of eigenvalues that minimize $f_{LdG}$ and the director vector is an eigenvector of $Q$ corresponding to the eigenvalue with the largest magnitude, the variations of $Q$ throughout the domain $\Omega$ can be interpreted as rotations of the eigenframe of $Q$. On the other hand, if the degree of $Q|_{\partial\Omega}$, i.e. the degree of ${\bf p}$ in \eqref{eq:pbc}, is an odd integer, then no such smooth uniaxial deformation exists going into $\Omega$ and there will be a small region of the size $1/\sqrt{w_l}$ where the topological constraint will force $Q_0$ to be isotropic and/or biaxial (cf. \cite{GM,canevari13}).
Let us now contrast this more relaxed target space with what occurs in our investigation where the full energy functional contains the surface energy term. The effect of this is that the admissible tensor fields $Q$ for the limiting problem are now $\mathcal D$-valued.
\begin{figure}[htb]
\centering
\includegraphics[height=1.5in]{D-class.eps}
\caption{Geometry of the target space of uniaxial nematic $\mathcal D$-valued maps. Only the values along the equator and at the poles of $\mathbb{S}^2$ are permitted.}
\label{fig:dclass}
\end{figure}
Since the normal to the surface of the film is then one of the eigenvectors of $Q$, there are only two types of possible uniaxial nematic states for admissible tensors---those with the director either perpendicular or parallel to the surface of the film (Figure \ref{fig:dclass}). If $\delta$ in \eqref{eq:f0pq} is small, the potential term $W$ would force the liquid crystal to be uniaxial nematic throughout most of $\Omega$, thus making the director either perpendicular or parallel to the surface normal $\hat z$. This orientational constraint is due to the strong influence of the surface energy and it makes director escape through the sequence of nematic states extremely costly. Given any topologically nontrivial planar boundary data for the director on $\partial\Omega$, we expect the director to remain planar as much as possible thus leading to formation of a localized region or regions with large gradients inside $\Omega$ for {\it any} choice of topologically nontrivial ${\bf p}_{\partial\Omega}$.
At the core of any singular region that may develop, one would expect that the orientations of the molecules would be parallel to the surface of the film, while remaining random within that surface. The orientational distribution function then is axially symmetric with respect to the normal to the film and the corresponding $Q$-tensor is uniaxial with the director being parallel to the normal (cf. \cite{virga_talk}). Since only two types of uniaxial states are permitted for admissible tensors, the connection between the uniaxial states in the core and away from the core has to occur through a sequence of biaxial states via the mechanism of so-called eigenvalue exchange \cite{palffy1994new}.
\section{Conclusions}
In this paper we have considered a rigorous dimension reduction procedure for nematic films described within the Landau-de Gennes model. We have established a general $\Gamma$-convergence result for the sequence of non-dimensional Landau-de Gennes energies as a small parameter $\varepsilon$ related to the film thickness tends to zero. Although this result is applicable to any combination of coefficients in the expression \eqref{bare} for the surface energy, we only consider cases when a $Q$-tensor minimizing $f_s(Q,\nu)$ in \eqref{bare} has the normal to the surface of the film as one of its eigenvectors.
In the thin-film limit, the dominant contributions to the energy come from the bulk elastic energy terms containing the derivatives in the direction normal to the surface of the film as well as from the surface energy terms that are independent of $\varepsilon$. The energy penalty due to normal derivative terms forces the limiting energy-minimizing $Q$-tensor field to be independent of the spatial variable perpendicular to the surface of the film and reduces the domain of the problem from three to two dimensions. The presence of the surface energy terms that are independent of $\varepsilon$ imposes constraints on the target space of admissible three-by-three $Q$-tensors. The resulting limiting bulk energy defined over the constrained set of $Q$-tensors consists of lower order bulk and surface energy contributions.
Our results, along with those in \cite{bauman_phillips_park}, demonstrate that the limiting problems for various parameter regimes can be studied by using the techniques developed for scalar Allen-Cahn type problems or Ginzburg-Landau-type vector-valued problems. Depending on the relationship between the coefficients of the surface and bulk energy terms, minimizing $Q$-tensor fields can develop boundary layers. For topologically nontrivial boundary data and large $w_l$ in \eqref{eq:f0}, the $Q$-tensor fields are characterized by vortices with characteristic core size $\sim1/\sqrt{w_l}.$ Note that even though $1/\sqrt{w_l}$ in this case is small, it is still much larger than $\varepsilon$ and two-dimensional vortices correspond to disclination lines perpendicular to the surface of the film.
\section{Acknowledgements}
D.G. acknowledges support from NSF DMS-1434969. P.S. acknowledges support from NSF DMS-1101290 and NSF DMS-1362879.
\bibliographystyle{ieeetr}
|
\section{Introduction}
The purpose of this paper is to prove that the coated inclusions neutral to all uniform fields in the isotropic medium are concentric balls in three dimensions. The coated inclusion is depicted by $(D, \Omega)$ where $D$ and $\Omega$ are bounded domains with Lipschitz boundaries in $\mathbb{R}^d$ ($d=2,3$) such that $\overline{D} \subset \Omega$. Here, $D$ represents the core and $\Omega \setminus D$ the shell. The conductivity (or the dielectric constant) is $\sigma_c$ in the core and $\sigma_s$ in the shell ($\sigma_c \neq \sigma_s$). If the structure $(D, \Omega)$ is inserted into the free space $\mathbb{R}^d$ with conductivity $\sigma_m$ where there is a uniform field $-\nabla ({\bf a} \cdot {\bf x})=-{\bf a}$ for some constant vector ${\bf a}$, then the field is perturbed in general. But for certain inclusions the field is not perturbed, in other words, the field does not recognize the existence of the inclusion. For example, the coated inclusion is made of concentric balls with specially chosen conductivities (confocal ellipsoids if $\sigma_m$ is anisotropic), one can see the uniform field is not perturbed. The inclusion with this property is called a neutral inclusion (or neutrally coated inclusion) and the neutral inclusion problem is to show that the inclusions of concentric balls (or confocal ellipsoids) are the only coated inclusions neutral to all uniform fields.
Let $\sigma$ denote the conductivity distribution of the medium so that
\begin{equation}
\sigma =
\begin{cases}
\sigma_c \quad&\mbox{in } D, \\
\sigma_s \quad&\mbox{in } \Omega \setminus D, \\
\sigma_m \quad &\mbox{in } \mathbb{R}^d \setminus \Omega.
\end{cases}
\end{equation}
Here we assume that $\sigma_c$ and $\sigma_s$ are constants (or isotropic matrices), but $\sigma_m$ is allowed to be anisotropic symmetric matrix. We consider the following problem:
\begin{equation}\label{transgeneral}
\begin{cases}
\nabla \cdot \sigma \nabla u = 0 \quad \mbox{in } \mathbb{R}^d, \\
u({\bf x})- {\bf a} \cdot {\bf x} = O(|{\bf x}|^{-2}) \quad \mbox{as } |{\bf x}| \to \infty,
\end{cases}
\end{equation}
where ${\bf a}$ is a constant vector. The term $u({\bf x})- {\bf a} \cdot {\bf x}$ depicts the perturbation of the potential due to insertion of the coated inclusion $(D,\Omega)$. If the potential is not perturbed, namely,
\begin{equation}
u({\bf x})- {\bf a} \cdot {\bf x} \equiv 0 \quad\mbox{in } \mathbb{R}^d \setminus \Omega,
\end{equation}
the coated inclusion $(D, \Omega)$ is said to be neutral to the field ${\bf a}$. If $(D, \Omega)$ is neutral to the field ${\bf e}_j$ for $j=1,\ldots, d$, where ${\bf e}_j$ is the standard basis of $\mathbb{R}^d$, then $(D, \Omega)$ is neutral to all uniform fields.
Much interest in neutrally coated inclusions was aroused by the work of Hashin and Shtrikman \cite{HS} and Hashin \cite{hashine}. They showed that since insertion of neutral inclusions does not perturb the outside uniform field, the effective conductivity of the assemblage filled with coated inclusions of many different scales is $\sigma_m$. We refer to \cite{milton} for developments on neutral inclusions in relation to the theory of composites. Another interest in neutral inclusions has been aroused in relation to invisibility cloaking. The neutral inclusion is invisible from the probe by uniform fields as observed in \cite{kerker}. Recently, the idea of neutrally coated inclusions has been extended to construct multi-coated circular structures which are neutral not only to uniform fields but also to fields of higher order up to $N$ for a given integer $N$ \cite{AKLL1}. It was proved there that the multi-coated structure combined with a transformation dramatically enhances the near cloaking of \cite{kohn1}. Cloaking by transformation optics was proposed in \cite{pendry} (and \cite{glu}).
As mentioned before, concentric balls (or disks) are neutral to all uniform fields by choosing $\sigma_c$, $\sigma_s$ and $\sigma_m$ properly ($\sigma_m$ is isotropic). Confocal ellipsoids (or ellipses) are also neutral to all uniform fields if $\sigma_m$ is anisotropic \cite{kerker} (see also section \ref{sec:ellipsoid}). Then a question arises naturally: are there any other shapes which are neutral to all uniform fields? In two dimensions there are no other shapes: if a coated inclusion $(D,\Omega)$ is neutral to all uniform fields in two dimensions, then $D$ and $\Omega$ are concentric disks (confocal ellipses if $\sigma_m$ is anisotropic). This is proved when $\sigma_c=0$ or $\infty$ in \cite{MS} and when $\sigma_c$ is finite in \cite{kl14}. In this paper we consider the neutral inclusion problem in three dimensions. We emphasize that the methods in \cite{kl14, MS} use powerful tools from complex analysis such as conformal mappings and harmonic conjugates, which cannot be applied to three dimensions. It is worth mentioning that there are many different shapes of coated inclusions neutral to a single uniform field as shown in two dimensions in \cite{JM, MS}.
We first show that if $(D,\Omega)$ is neutral to all uniform fields in three dimensions and if $\sigma_c>\sigma_s$, then the following problem admits a solution:
\begin{equation}\label{free}
\left\{
\begin{array}{ll}
\displaystyle \Delta w= k \quad &\mbox{in } \Omega \setminus \overline{D}, \\
\noalign{\smallskip}
\displaystyle \nabla w = 0 \quad &\mbox{on } \partial \Omega , \\
\noalign{\smallskip}
\displaystyle \nabla w({\bf x})= {\bf A}{\bf x} +{\bf d} \quad &\mbox{on } \partial D,
\end{array}
\right.
\end{equation}
where $k(\not=0)$ is a constant, ${\bf A}$ is a symmetric matrix, and ${\bf d}$ is a constant vector. We emphasize that this is an over-determined problem because $\nabla w$ is prescribed on the boundaries. The problem, which is of independent interest, is to prove that if \eqnref{free} admits a solution in three dimensions, then $D$ and $\Omega$ are confocal ellipsoids. If $D$ and $\Omega$ are confocal ellipsoids, then \eqnref{free} admits a solution and ${\bf A}$ should be either positive or negative-definite depending on the sign of $k$ (see section \ref{sec:ellipsoid}). So a part of the problem is to show that ${\bf A}$ is either positive or negative-definite. In two dimensions it is proved in \cite{kl14} that if \eqnref{free} admits a solution then $D$ and $\Omega$ are confocal ellipses (concentric disks if ${\bf A}$ is isotropic). However, the proof there is based on the powerful result that there is a conformal mapping from $\Omega \setminus \overline{D}$ onto an annulus. So it cannot be extended to three dimensions. The condition $\sigma_c>\sigma_s$, which is not natural, is required because of a technical reason for the derivation of \eqnref{free} in subsection \ref{sub22}. Even though we do not know how to do so, it is likely that the condition can be removed.
In this paper we solve the problem partially as the following theorem shows.
\begin{theorem}\label{conBall}
Let $D$ and $\Omega$ be bounded domains with Lipschitz boundaries in $\mathbb{R}^3$ with $\overline{D} \subset \Omega$. Suppose that $\Omega \setminus \overline{D}$ is connected.
If {\rm \eqnref{free}} admits a solution for ${\bf A}=c{\bf I}$ for some constant $c$ where ${\bf I}$ is the identity matrix in three dimensions, then $D$ and $\Omega$ are concentric balls.
\end{theorem}
As a consequence, we obtain the following theorem.
\begin{theorem}\label{thm:main}
Let $D$ and $\Omega$ be bounded domains with Lipschitz boundaries in $\mathbb{R}^3$ with $\overline{D} \subset \Omega$. Suppose that $\partial D$ is connected and $\mathbb R^3 \setminus \overline{D}$ is simply connected. If $\sigma_m$ is isotropic, $\sigma_c>\sigma_s$ and $(D,\Omega)$ is neutral to all uniform fields, then $D$ and $\Omega$ are concentric balls.
\end{theorem}
This paper is organized as follows. In section \ref{sec:over} we show that if $(D,\Omega)$ is neutral to all uniform fields then \eqnref{free} admits a solution. In section \ref{sec:ellipsoid} we construct a solution to \eqnref{free} when $D$ and $\Omega$ are confocal ellipsoids. Section \ref{sec:proof} is to prove Theorem \ref{conBall}. In section \ref{sec:newton} we formulate the problem \eqnref{free} using Newtonian potentials and relate the problem with a known characterization of ellipsoids.
\section{Derivation of the over-determined problem}\label{sec:over}
In this section we derive \eqnref{free} out of the neutral inclusion problem. We will do so only in three dimensions since \eqnref{free} has been derived in two dimensions \cite{kl14}. We assume that $\partial D$ is connected and $\mathbb R^3 \setminus \overline{D}$ is simply connected.
Suppose, after diagonalization, that
\begin{equation}
\sigma_m = \mbox{diag} [\sigma_{m,1}, \sigma_{m,2}, \sigma_{m,3}].
\end{equation}
Let $u_j$, $j=1,2,3$, be the solution to
\begin{equation}\label{trans}
\left\{
\begin{array}{ll}
\nabla \cdot \sigma \nabla u_j = 0 \quad &\mbox{in } \mathbb{R}^3, \\
u_j(x)-x_j = O(|x|^{-2}) \quad &\mbox{as } |x| \to \infty.
\end{array}
\right.
\end{equation}
The structure being neutral to all three fields means that $u_j(x)-x_j=0$ in $\mathbb{R}^3 \setminus \Omega$ for $j=1,2,3$.
Let
\begin{equation}\label{wj}
w_j = \frac{1}{\beta_j} u_j
\end{equation}
where
$$
\beta_j := \frac{\sigma_{m,j}}{\sigma_s}-1, \quad j=1,2,3,
$$
and ${\bf w}=(w_1,w_2,w_3)^T$ ($T$ for transpose). Set also
\begin{equation}\label{BB}
{\bf B}= \mbox{diag} \, [1/\beta_1, 1/\beta_2, 1/\beta_3].
\end{equation}
We will show the following:
\begin{itemize}
\item[(i)] $\nabla {\bf w}$ is symmetric and $\div {\bf w}$ is constant, and hence there is a function $\psi$ in $\overline{\Omega}\setminus D$ such that
\begin{equation}\label{bwgrad}
{\bf w}=\nabla \psi\ \mbox{ and }\ \Delta \psi = \mbox{Tr}\, {\bf B} +1 \quad\mbox{in }\Omega \setminus \overline{D}.
\end{equation}
\item[(ii)] ${\bf w}({\bf x})=c_0{\bf x}+ {\bf d}$, ${\bf x} \in D$ for some constant $c_0$ and constant vector ${\bf d}$ (under the assumption that $\sigma_c > \sigma_s$).
\end{itemize}
We emphasize that it is in (ii) where the condition $\sigma_c > \sigma_s$ is required.
Once we have (i) and (ii), then we can show that \eqnref{free} has a solution. In fact, since $u_j=x_j$ on $\partial\Omega$, we have
$$
\nabla \psi({\bf x})= {\bf B} {\bf x} \quad\mbox{on } \partial\Omega.
$$
Note that $\nabla \psi({\bf x})= c_0{\bf x}+ {\bf d}$ on $\partial D$. Now define
\begin{equation}\label{definition_of_w}
w({\bf x}):= \psi({\bf x}) - \frac{1}{2} {\bf x} \cdot {\bf B} {\bf x}.
\end{equation}
Then $w$ satisfies \eqnref{free} with $k=1$ and ${\bf A}=c_0 {\bf I} - {\bf B}$. We emphasize that if $\sigma_m$ is isotropic, so are ${\bf B}$ and ${\bf A}$.
\subsection{Proof of (i)}\label{shell}
Let us first deal with the case when $0< \sigma_c < \infty$. Denote by $\nu = (n_{1}, n_{2}, n_{3})^T$ the outward unit normal vector field to $\partial\Omega$ or $\partial D$. Note that the solution $u_j$ ($j=1,2,3$) to \eqnref{trans} satisfies the following transmission conditions on two interfaces :
\begin{equation}\label{trans1}
u_j|_+ - u_j|_- =0, \quad \sigma_{m,j} \pd{u_j}{\nu}\Big|_+ - \sigma_s \pd{u_j}{\nu}\Big|_-=0 \quad \mbox{on } \partial\Omega
\end{equation}
and
\begin{equation}\label{trans2}
u_j|_+ - u_j|_- = 0, \quad \sigma_s \pd{u_j}{\nu}\Big|_+ - \sigma_c \pd{u_j}{\nu}\Big|_-=0 \quad \mbox{on } \partial D
\end{equation}
where $+$ denotes the limit from outside and $-$ that from inside of $\Omega$ or $D$. If $(D, \Omega)$ is neutral to $x_j$, then $u_j({\bf x})-x_j=0$ in $\mathbb{R}^3 \setminus \Omega$, so we see from \eqnref{trans1} that
\begin{equation}\label{trans3}
u_j|_- = x_j, \quad \sigma_s \pd{u_j}{\nu}\Big|_-= \sigma_{m,j} n_j \quad \mbox{on } \partial\Omega.
\end{equation}
In other words, $u_j$ is the solution to the following over-determined problem:
\begin{equation}\label{overdetermine}
\left\{
\begin{array}{ll}
\nabla \cdot \sigma \nabla u_j = 0 \quad &\mbox{in } \Omega, \\
u_j = x_j, \ \ \displaystyle \pd{u_j}{\nu} = \frac{\sigma_{m,j}}{\sigma_s} n_j \quad &\mbox{on } \partial\Omega.
\end{array}
\right.
\end{equation}
Let $v_j \in C^2(\overline{\Omega})$. Then we see from the divergence theorem and \eqnref{trans2} that
\begin{align*}
\int_{\partial\Omega} \pd{u_j}{\nu}\Big|_- v_j - u_j \pd{v_j}{\nu}
&= - \int_{\Omega \setminus D} u_j \Delta v_j + \int_{\partial D} \pd{u_j}{\nu}\Big|_+ v_j - u_j \pd{v_j}{\nu} \\
&= - \int_{\Omega \setminus D} u_j \Delta v_j + \left( \frac{\sigma_c}{\sigma_s} -1 \right) \int_{\partial D} \pd{u_j}{\nu}\Big|_- v_j + \int_{\partial D} \pd{u_j}{\nu}\Big|_- v_j - u_j \pd{v_j}{\nu} \\
&= - \int_{\Omega \setminus D} u_j \Delta v_j + \left( \frac{\sigma_c}{\sigma_s} -1 \right) \int_{D} \nabla u_j \cdot \nabla v_j - \int_{D} u_j \Delta v_j \\
&= - \int_{\Omega} u_j \Delta v_j + \left( \frac{\sigma_c}{\sigma_s} -1 \right) \int_{D} \nabla u_j \cdot \nabla v_j .
\end{align*}
On the other hand, we see from \eqnref{trans3} that
\begin{align*}
\int_{\partial\Omega} \pd{u_j}{\nu}\Big|_- v_j - u_j \pd{v_j}{\nu}
&= \int_{\partial\Omega} \frac{\sigma_{m,j}}{\sigma_s} n_j v_j - y_j \pd{v_j}{\nu} \\
&= \left( \frac{\sigma_{m,j}}{\sigma_s} -1 \right) \int_{\Omega} \pd{v_j}{y_j} - \int_{\Omega} y_j \Delta v_j.
\end{align*}
Equating two identities above we obtain
\begin{equation}\label{weakform}
\int_{\Omega} (y_j-u_j) \Delta v_j + \alpha \int_{D} \nabla u_j \cdot \nabla v_j - \beta_j \int_{\Omega} \pd{v_j}{y_j}=0 , \quad j=1,2,3
\end{equation}
for $v_j \in C^2(\overline{\Omega})$, where $\alpha$ and $\beta_j$ are defined for ease of notation to be
\begin{equation}
\alpha= \frac{\sigma_c}{\sigma_s} -1 \quad\mbox{and} \quad \beta_j= \frac{\sigma_{m,j}}{\sigma_s}-1.
\end{equation}
Let $w_j$ be defined by $w_j := \frac{1}{\beta_j} u_j$ as in \eqnref{wj}.
Then \eqnref{weakform} can be rephrased as
\begin{equation}\label{weakform11}
\int_{\Omega} (\frac{1}{\beta_j} y_j-w_j) \Delta v_j + \alpha \int_{D} \nabla w_j \cdot \nabla v_j - \int_{\Omega} \pd{v_j}{y_j}=0 , \quad j=1,2,3.
\end{equation}
Summing \eqnref{weakform11} over $j=1,2,3$ we have
$$
\int_{\Omega} \sum_{j=1}^3 (\frac{1}{\beta_j}y_j-w_j) \Delta v_j + \alpha \int_{D} \sum_{j=1}^3 \nabla w_j \cdot \nabla v_j - \int_{\Omega} \sum_{j=1}^3 \pd{v_j}{y_j}=0
$$
for $v_j \in C^2(\overline{\Omega})$. If we use vector notation ${\bf w}=(w_1,w_2, w_3)^T$ and ${\bf v}=(v_1,v_2,v_3)^T$ ($T$ for transpose), then the above identity can be rewritten as
\begin{equation}\label{weakform2}
\int_{\Omega} ({\bf B}{\bf y}-{\bf w}) \cdot \Delta {\bf v} + \alpha \int_{D} \nabla {\bf w} : \nabla {\bf v} - \int_{\Omega} \div {\bf v} =0.
\end{equation}
Here and afterwards ${\bf A}:{\bf B}$ denote the contraction of two matrices ${\bf A}$ and ${\bf B}$,
{\it i.e.}, ${\bf A}:{\bf B}=\sum a_{ij}b_{ij}=\textrm{Tr}({\bf A}^{T}{\bf B})$.
Let $\Gamma$ be the fundamental solution of the Laplace operator in $\mathbb{R}^3$, {\it i.e.},
\begin{equation}
\Gamma({\bf x}) := - \frac{1}{4 \pi |{\bf x}|}, \quad {\bf x} \neq 0.
\end{equation}
Let $v_j({\bf y}) =\Gamma({\bf x}-{\bf y})$ for a fixed ${\bf x} \in \Omega$.
Since $\Delta v_j({\bf y}) = \delta({\bf x}-{\bf y})$, by applying the divergence theorem over $\Omega \setminus B_\epsilon({\bf x})$ for sufficiently small $\epsilon$ ($B_\epsilon({\bf x})$ is the ball of radius $\epsilon$ centered at ${\bf x}$) we see from \eqnref{weakform11} that
\begin{equation}\label{repre}
w_j ({\bf x}) = \frac{1}{\beta_j} x_j + \alpha \int_{D} \nabla w_j({\bf y}) \cdot \nabla_{\bf y} \Gamma({\bf x}-{\bf y})d{\bf y} + \pd{}{x_j} N_\Omega({\bf x}) , \quad {\bf x} \in \Omega, \quad j=1,2,3,
\end{equation}
where $N_\Omega$ is the Newtonian potential on a domain $\Omega$, {\it i.e.},
\begin{equation}
N_\Omega({\bf x}) := \int_\Omega \Gamma({\bf x}-{\bf y}) d{\bf y}, \quad {\bf x} \in \mathbb{R}^3.
\end{equation}
Let
$$
f_j({\bf x}) := \int_{D} \nabla w_j({\bf y}) \cdot \nabla_{\bf y} \Gamma({\bf x}-{\bf y}) d{\bf y}, \quad j=1,2,3,
$$
and let ${\bf f}= (f_1, f_2, f_3)^T$. Note that $f_j$ is harmonic in $\mathbb{R}^3 \setminus \overline{D}$, and \eqnref{repre} can be rewritten as
\begin{equation}\label{ujy}
{\bf w}({\bf x}) = \alpha {\bf f}({\bf x}) + \nabla \left( \frac{1}{2} {\bf x}\cdot {\bf B}{\bf x} + N_\Omega({\bf x}) \right) , \quad {\bf x} \in \Omega, \quad j=1,2,3.
\end{equation}
For any fixed ${\bf x} \in \mathbb{R}^3 \setminus \overline{\Omega}$, let
$$
v_j({\bf y})= \pd{}{x_j} \Gamma({\bf x}-{\bf y}), \quad j=1,2,3.
$$
Then $\div {\bf v} ({\bf y})= -\Delta_{\bf y} \Gamma({\bf x}-{\bf y})=0$ and $\Delta {\bf v}({\bf y})=0$ for ${\bf y} \in \Omega$. So we see from \eqnref{weakform2} that
$$
\int_D \nabla {\bf w} : \nabla {\bf v} =0,
$$
and hence
\begin{equation}\label{sumfj}
\div {\bf f}({\bf x}) = \int_{D} \sum_j \nabla w_j({\bf y}) \cdot \nabla \pd{}{x_j} \Gamma({\bf x}-{\bf y}) d{\bf y}= \int_D \nabla {\bf w} : \nabla {\bf v} =0, \quad {\bf x} \in \mathbb{R}^3 \setminus \overline{\Omega}.
\end{equation}
Since $f_j$ is harmonic in $\mathbb{R}^3 \setminus \overline{D}$, \eqnref{sumfj} holds for all ${\bf x} \in \mathbb{R}^3 \setminus \overline{D}$.
Again fix ${\bf x} \in \mathbb{R}^3 \setminus \overline{\Omega}$. Let $\{i,j,k \}$ be a permutation of $\{ 1,2,3 \}$ and let
$$
v_i({\bf y})= \pd{}{x_j} \Gamma({\bf x}-{\bf y}), \quad v_j({\bf y})= -\pd{}{x_i} \Gamma({\bf x}-{\bf y}), \quad v_k=0, \quad {\bf y} \in \Omega.
$$
Then, $\Delta {\bf v}=0$ and $\div {\bf v}=0$ in $\Omega$. So we have from \eqnref{weakform2}
$$
\int_{D} \nabla w_i({\bf y}) \cdot \nabla \pd{}{x_j} \Gamma({\bf x}-{\bf y}) d{\bf y} - \int_{D} \nabla w_j({\bf y}) \cdot \nabla \pd{}{x_i} \Gamma({\bf x}-{\bf y}) d{\bf y}=0,
$$
which implies that
\begin{equation}
\partial_i f_j({\bf x}) = \partial_j f_i({\bf x})
\end{equation}
for all ${\bf x} \in \mathbb{R}^3 \setminus \overline{\Omega}$ and hence for all ${\bf x} \in \mathbb{R}^3 \setminus \overline{D}$. Moreover, since $\mathbb{R}^3 \setminus \overline{D}$ is simply connected, by the Stokes theorem
there is $\varphi$ such that
\begin{equation}\label{fjy}
{\bf f}({\bf x}) = \nabla \varphi({\bf x}), \quad {\bf x} \in \mathbb{R}^3 \setminus \overline{D}.
\end{equation}
Because of \eqnref{sumfj}, we have
\begin{equation}\label{Deltag}
\Delta \varphi({\bf x})=0 , \quad {\bf x} \in \mathbb{R}^3 \setminus \overline{D}.
\end{equation}
Let
\begin{equation}
\psi({\bf x}) = \alpha \varphi({\bf x}) + \frac{1}{2} {\bf x} \cdot {\bf B}{\bf x} + N_\Omega({\bf x}) , \quad {\bf x} \in \Omega \setminus \overline{D}.
\end{equation}
Then, we have from \eqnref{ujy} and \eqnref{fjy}
\begin{equation}\label{ujy2}
{\bf w} ({\bf x}) = \nabla \psi({\bf x}) , \quad {\bf x} \in \Omega\setminus \overline{D}, \quad j=1,2,3.
\end{equation}
Since $\Delta N_\Omega({\bf x})=1$ for ${\bf x} \in \Omega$, we have from \eqnref{Deltag} that
\begin{equation}
\Delta \psi({\bf x})= \mbox{Tr}\, {\bf B} + 1, \quad {\bf x} \in \Omega\setminus \overline{D}.
\end{equation}
So far we have shown that $\nabla {\bf w}$ is symmetric, $\div {\bf w}$ is constant, and \eqnref{bwgrad} holds when $\sigma_c$ is finite.
We now assume that $\sigma_c=0$. In this case the problem \eqnref{overdetermine} becomes
\begin{equation}\label{overzero}
\left\{
\begin{array}{ll}
\Delta u_j = 0 \quad &\mbox{in } \Omega \setminus \overline{D}, \\
\noalign{\smallskip}
\displaystyle \pd{u_j}{\nu} =0 \quad &\mbox{on } \partial D, \\
\noalign{\smallskip}
u_j = x_j, \ \ \displaystyle \pd{u_j}{\nu} = \frac{\sigma_{m,j}}{\sigma_s} n_j \quad &\mbox{on } \partial\Omega.
\end{array}
\right.
\end{equation}
So, we see in a way similar to \eqnref{weakform} that
\begin{equation}\label{weakzero}
\int_{\Omega} y_j \Delta v_j - \int_{\Omega \setminus D} u_j \Delta v_j - \int_{\partial D} u_j \pd{v_j}{\nu} - \beta_j \int_{\Omega} \pd{v_j}{y_j}=0
\end{equation}
for all $v_j \in C^2(\overline{\Omega})$. So we obtain a representation of the solution similar to \eqnref{repre}:
\begin{equation}\label{reprezero}
w_j ({\bf x}) = \frac{1}{\beta_j} x_j - \int_{\partial D} w_j({\bf y}) \pd{}{\nu}\Gamma({\bf x}-{\bf y})d\sigma({\bf y}) + \pd{}{x_j} N_\Omega({\bf x}) , \quad {\bf x} \in \Omega \setminus D.
\end{equation}
So, we infer in the exactly same way as in the previous sections that $\nabla{\bf w}$ is symmetric and $\div {\bf w}$ is constant, and there is a function $\psi$ such that \eqnref{bwgrad} holds.
Suppose that $\sigma_c=\infty$. In this case the problem \eqnref{overdetermine} becomes
\begin{equation}\label{overinfty}
\left\{
\begin{array}{ll}
\Delta u_j = 0 \quad &\mbox{in } \Omega \setminus \overline{D}, \\
u_j = \gamma_j \ (\mbox{constant}) \quad &\mbox{on } \partial D, \\
\noalign{\smallskip}
u_j = x_j, \ \ \displaystyle \pd{u_j}{\nu} = \frac{\sigma_{m,j}}{\sigma_s} n_j \quad &\mbox{on } \partial\Omega.
\end{array}
\right.
\end{equation}
The constant $\gamma_j$ is determined by the condition
$$
\int_{\partial D} \pd{u_j}{\nu} \Big|_+ =0.
$$
We then obtain similarly to \eqnref{weakform}
\begin{equation}\label{weakinfty}
\int_{\Omega} y_j \Delta v_j - \int_{\Omega \setminus D} u_j \Delta v_j + \int_{\partial D} \pd{u_j}{\nu} v_j - \gamma_j \int_{D} \Delta v_j - \beta_j \int_{\Omega} \pd{v_j}{y_j}=0
\end{equation}
for all $v_j \in C^2(\overline{\Omega})$. We then obtain a representation of the solution similar to \eqnref{repre}:
\begin{equation}\label{repreinfty}
w_j ({\bf x}) = \frac{1}{\beta_j} x_j + \int_{\partial D} \pd{w_j}{\nu}({\bf y}) \Gamma({\bf x}-{\bf y})d\sigma({\bf y}) + \pd{}{x_j} N_\Omega({\bf x}) , \quad {\bf x} \in \Omega \setminus D.
\end{equation}
So, we infer that $\nabla {\bf w}$ is symmetric, $\div {\bf w}$ is constant, and there is a function $\psi$ such that \eqnref{bwgrad} holds.
\subsection{Proof of (ii)}\label{sub22}
The transmission conditions \eqnref{trans2} on $\partial D$ can be rephrased as
\begin{equation}\label{transvec}
{\bf w}|_+= {\bf w}|_-, \quad \sigma_s \nabla {\bf w}|_+ \nu = \sigma_c \nabla {\bf w}|_- \nu.
\end{equation}
Let ${\bf t}_1$ and ${\bf t}_2$ be two orthonormal tangent vector fields to $\partial D$.
Then, we have
$$
(\div {\bf w})_- = \langle (\nabla {\bf w})_- \nu, \nu \rangle + \langle (\nabla {\bf w})_- {\bf t}_1, {\bf t}_1 \rangle + \langle (\nabla {\bf w})_- {\bf t}_2, {\bf t}_2 \rangle,
$$
and
$$
(\div {\bf w})_+ = \langle (\nabla {\bf w})_+ \nu, \nu \rangle + \langle (\nabla {\bf w})_+ {\bf t}_1, {\bf t}_1 \rangle + \langle (\nabla {\bf w})_+ {\bf t}_2, {\bf t}_2 \rangle.
$$
Here $(\div {\bf w})_-$ denotes the limit of $\div {\bf w}$ to $\partial D$ from inside $D$, and $(\div {\bf w})_+$ denotes that from outside $D$. Since
$$
\langle (\nabla {\bf w})_- {\bf t}_j, {\bf t}_j \rangle = \langle (\nabla {\bf w})_+ {\bf t}_j, {\bf t}_j \rangle, \quad j=1,2,
$$
we have
$$
(\div {\bf w})_- - (\div {\bf w})_+ = \langle (\nabla {\bf w})_- \nu, \nu \rangle - \langle (\nabla {\bf w})_+ \nu, \nu \rangle.
$$
It then follows from the second identity in \eqnref{transvec} that
\begin{equation}\label{transvec2}
\left\langle \big( (\nabla {\bf w})_-^T - \frac{\sigma_c}{\sigma_s} (\nabla {\bf w})_- \big) \nu, \nu \right\rangle = (\div {\bf w})_- - (\div {\bf w})_+ .
\end{equation}
On the other hand, since $(\nabla {\bf w})_+$ is symmetric, we obtain
\begin{align}
\left\langle \big( (\nabla {\bf w})_-^T - \frac{\sigma_c}{\sigma_s} (\nabla {\bf w})_- \big) \nu, {\bf t}_j \right\rangle
& = \langle \nu, (\nabla {\bf w})_- {\bf t}_j \rangle - \frac{\sigma_c}{\sigma_s} \langle (\nabla {\bf w})_- \nu, {\bf t}_j \rangle \nonumber \\
& = \langle \nu, (\nabla {\bf w})_+ {\bf t}_j \rangle - \langle (\nabla {\bf w})_+ \nu, {\bf t}_j \rangle =0. \label{transvec3}
\end{align}
We then infer from \eqnref{transvec2} and \eqnref{transvec3} that
\begin{equation}\label{transvec4}
\left( (\nabla {\bf w})_-^T - \frac{\sigma_c}{\sigma_s} (\nabla {\bf w})_- \right) \nu = (\div {\bf w})_- \nu - (\div {\bf w})_+ \nu.
\end{equation}
Recall that $\div {\bf w}$ is constant in $\Omega \setminus D$. Let
\begin{equation}
{\bf v}({\bf x})= {\bf w}({\bf x}) - \frac{(\div {\bf w})_+}{2+ \frac{\sigma_c}{\sigma_s}} {\bf x}, \quad {\bf x} \in D.
\end{equation}
Then one can see from \eqnref{transvec4} that
\begin{equation}\label{zerobdry}
\left( (\nabla {\bf v})^T - \frac{\sigma_c}{\sigma_s} (\nabla {\bf v}) \right) \nu - (\div {\bf v}) \nu =0 \quad \mbox{on } \partial D.
\end{equation}
Let ${\bf g}$ be a smooth vector field on $\overline{D}$.
It follows from \eqnref{zerobdry} and the divergence theorem that
\begin{align*}
0 &= \int_{\partial D} \nu \cdot \left( (\nabla {\bf v}){\bf g} - \frac{\sigma_c}{\sigma_s} (\nabla {\bf v})^T {\bf g} - (\div {\bf v}) {\bf g} \right) d\sigma \\
&= \int_{D} \div \left( (\nabla {\bf v}){\bf g} - \frac{\sigma_c}{\sigma_s} (\nabla {\bf v})^T {\bf g} - (\div {\bf v}) {\bf g} \right) d{\bf x}.
\end{align*}
One can easily show that
$$
\div \left( (\nabla {\bf v}){\bf g} - \frac{\sigma_c}{\sigma_s} (\nabla {\bf v})^T {\bf g} - (\div {\bf v}) {\bf g} \right) =
\nabla {\bf v}^T : \nabla {\bf g} - \frac{\sigma_c}{\sigma_s} \nabla {\bf v}: \nabla {\bf g} - (\div {\bf v})(\div {\bf g}),
$$
and so we obtain
\begin{equation}
\int_D \nabla {\bf v}^T : \nabla {\bf g} - \frac{\sigma_c}{\sigma_s} \nabla {\bf v}: \nabla {\bf g} - (\div {\bf v})(\div {\bf g})=0.
\end{equation}
Using notation
$$
\widehat{\nabla} {\bf v} :=\frac{1}{2}(\nabla {\bf v} +\nabla {\bf v}^{T}) \quad\mbox{and}\quad \breve{\nabla} {\bf v} :=\frac{1}{2}(\nabla {\bf v} -\nabla {\bf v}^{T}),
$$
it can be rewritten as
\begin{equation}
(1- \frac{\sigma_c}{\sigma_s}) \int_D \widehat{\nabla} {\bf v}: \widehat{\nabla}{\bf g} - (1+\frac{\sigma_c}{\sigma_s}) \int_D \breve{\nabla} {\bf v}: \breve{\nabla}{\bf g} - \int_D (\div {\bf v})(\div {\bf g})=0.
\end{equation}
If $\sigma_c > \sigma_s$, then we take ${\bf g}={\bf v}$ so that
\begin{equation}\label{239}
(1- \frac{\sigma_c}{\sigma_s}) \int_D |\widehat{\nabla} {\bf v}|^2 - (1+\frac{\sigma_c}{\sigma_s}) \int_D |\breve{\nabla} {\bf v}|^2 - \int_D (\div {\bf v})^2=0.
\end{equation}
Thus, we infer that ${\bf v}$ is constant in $D$ and hence
\begin{equation}
{\bf w}({\bf x}) = \frac{(\div {\bf w})_+}{2+ \frac{\sigma_c}{\sigma_s}} {\bf x} + \mbox{a constant vector}, \quad {\bf x} \in D.
\end{equation}
If $\sigma_c = \infty$, then ${\bf u}$ is constant on $\partial D$, and hence $(\nabla {\bf w}){\bf t}=0$ on $\partial D$ for any tangential vector ${\bf t}$ to $\partial D$. Since $\nabla {\bf w}$ is symmetric and $\div {\bf w}$ is constant, it implies that
$$
(\nabla {\bf w})\nu = c_0\nu \quad \mbox{on } \partial D
$$
for some constant $c_0$. So, we can see that (ii) holds.
\section{Existence of solutions on confocal ellipsoids}\label{sec:ellipsoid}
We first mention that the solution $w$ to \eqnref{free} is unique in the sense that if $w_1$ and $w_2$ are two solutions (with different $k,\, {\bf A}$'s, and ${\bf d}$'s), then $w_1=C w_2 +E$ for some constants $C$ and $E$. In fact, if $w_j$ is a solution to \eqnref{free} with $k=k_j \neq 0$, ${\bf A}={\bf A}_j$ and ${\bf d}={\bf d}_{j}$ ($j=1,2$), then $w=w_1-\frac{k_1}{k_2} w_2$ satisfies $\Delta w= 0$ in $\Omega \setminus \overline{D}$ and $\nabla w = 0$ on $\partial\Omega$, so we have that $w$ must be a constant.
We now construct a solution to \eqnref{free} when $D$ and $\Omega$ are confocal ellipsoids. To do so, assume that $\partial D$ is given by
\begin{equation} \label{defb3}
\frac{x^2_1}{c^2_1}+\frac{x^2_2}{c^2_2}+\frac{x^2_3}{c^2_3}=1.
\end{equation}
We then use the confocal ellipsoidal coordinates
$\rho, \ \mu,\ \xi$ such that
\begin{align*}
\frac{x^2_1}{c^2_1+\rho}+\frac{x^2_2}{c^2_2+\rho}+\frac{x^2_3}{c^2_3+\rho}=1,\\
\noalign{\smallskip}
\frac{x^2_1}{c^2_1+\mu}+\frac{x^2_2}{c^2_2+\mu}+\frac{x^2_3}{c^2_3+\mu}=1,\\
\frac{x^2_1}{c^2_1+\xi}+\frac{x^2_2}{c^2_2+\xi}+\frac{x^2_3}{c^2_3+\xi}=1,
\end{align*}
subject to the conditions $-c^2_3 < \xi < -c^2_2 < \mu < -c^2_1 < \rho$. Then the confocal ellipsoid $\partial\Omega$ is given by $\rho=\rho_0$ for some $\rho_0>0$.
Let
\begin{equation}\label{grho}
g(\rho) = (c_1^2+\rho)(c_2^2+\rho)(c_3^2+\rho),
\end{equation}
and define
\begin{equation}\label{vpform}
\varphi_j (\rho)= \int_\rho^\infty \frac{1}{(c_j^2+s)\sqrt{g(s)}} ds, \quad j=1,2,3.
\end{equation}
Then the function $w$ defined by
\begin{equation}
w({\bf x}) = \frac{1}{2} \int_\rho^\infty \frac{1}{\sqrt{g(s)}} ds - \frac{1}{2} \sum_{j=1}^3 \varphi_j(\rho) x_j^2 + \frac{1}{2} \sum_{j=1}^3 \varphi_j(\rho_0) x_j^2
\end{equation}
is a solution of \eqnref{free}. In fact, we can see that
$$
\pd{}{x_i} \left[\frac{1}{2} \int_\rho^\infty \frac{1}{\sqrt{g(s)}} ds - \frac{1}{2} \sum_{j=1}^3 \varphi_j(\rho) x_j^2 \right] = \left( - \frac{1}{\sqrt{g(\rho)}} - \sum_{j=1}^3 \varphi_j'(\rho) x_j^2 \right) \pd{\rho}{x_i} - \varphi_i(\rho) x_i.
$$
Since
$$
\sum_{j=1}^3 \varphi_j'(\rho) x_j^2 = - \sum_{j=1}^3 \frac{x_j^2}{(c_j^2+\rho)\sqrt{g(\rho)}} = -\frac{1}{\sqrt{g(\rho)}},
$$
we have
$$
\pd{}{x_i} \left[\frac{1}{2} \int_\rho^\infty \frac{1}{\sqrt{g(s)}} ds - \frac{1}{2} \sum_{j=1}^3 \varphi_j(\rho) x_j^2 \right] = - \varphi_i(\rho) x_i,
$$
from which we see that
\begin{equation}
\nabla w({\bf x})= - (\varphi_1(\rho)x_1, \varphi_2(\rho)x_2, \varphi_3(\rho)x_3) + (\varphi_1(\rho_0)x_1, \varphi_2(\rho_0)x_2, \varphi_3(\rho_0)x_3).
\end{equation}
Using the relation
\begin{equation}\label{derivative}
\pd{\rho}{x_i}=\frac{2x_i}{c_i^2+\rho}\left[ \frac{x_1^2}{(c_1^2+\rho)^2}+\frac{x_2^2}{(c_2^2+\rho)^2}+\frac{x_3^2}{(c_3^2+\rho)^2}\right]^{-1},
\end{equation}
we see that $\Delta w$ is constant. Note that $\nabla w=0$ on $\partial\Omega$ ($\rho=\rho_0$) and $\nabla w={\bf A} {\bf x}$ on $\partial D$ where
\begin{equation}
{\bf A}= \mbox{diag} [ \varphi_1(\rho_0) - \varphi_1(0), \varphi_2(\rho_0) - \varphi_2(0), \varphi_3(\rho_0)-\varphi_3(0) ].
\end{equation}
We emphasize that ${\bf A}$ is negative-definite.
\section{Proof of Theorem \ref{conBall}}\label{sec:proof}
Let $w$ be the solution to \eqnref{free} with ${\bf A}=c{\bf I}$. We notice that $c \not=0$. Indeed, if $c = 0$, then we have
$$
0 \not=k|\Omega\setminus\overline{D}|= \int_{\Omega\setminus\overline{D}}\Delta w\ dx = \int_{\partial\Omega} \frac {\partial w}{\partial\nu}\ d\sigma - \int_{\partial D} \frac {\partial w}{\partial\nu}\ d\sigma = 0 -\int_{\partial D} \nu\cdot {\bf d}\ d\sigma = 0,
$$
which is a contradiction. Since $c\not=0$, by introducing new variables
$$
{\bf y} = {\bf x} + \frac 1c {\bf d},
$$
we may assume that ${\bf d} =\bold 0$. Set
\begin{equation}
A_{ij} = x_j \pd{}{x_i} - x_i \pd{}{x_j}, \quad i \neq j.
\end{equation}
It is worth mentioning that $A_{ij}$ is the angular derivative. Observe that $A_{ij}$ commutes with $\Delta$, namely, $A_{ij} \Delta = \Delta A_{ij}$. So, we have $\Delta A_{ij} w=0$ in $\Omega \setminus \overline{D}$. Note that $A_{ij} w=0$ on $\partial\Omega$. Since $\nabla w({\bf x})= c{\bf x}$ on $\partial D$, we see that $A_{ij} w=0$ on $\partial D$. Then the maximum principle yields that
\begin{equation}
A_{ij} w=0 \quad\mbox{in } \Omega \setminus \overline{D}.
\end{equation}
Since $\Delta w=k$ in $\Omega \setminus \overline{D}$, $w$ satisfies the ordinary differential equation
\begin{equation}\label{ode1}
\frac{\partial^2 w}{\partial r^2} + \frac{2}{r} \frac{\partial w}{\partial r} =k \quad\mbox{in } \Omega \setminus \overline{D}
\end{equation}
for $r=|{\bf x}|$. Choose a ball $B$ with $\overline{B} \subset \Omega \setminus \overline{D}$. By \eqnref{ode1}, $w$ is of the form
\begin{equation}\label{ode2}
w(r)= \frac{k}{6} r^2 + \frac{k_1}{r} + k_2 \quad\mbox{in } \overline{B}
\end{equation}
for some real constants $k_1$ and $k_2$. Since $\Omega \setminus \overline{D}$ is connected and
$$
\Delta\left(w- \frac{k}{6} r^2 - \frac{k_1}{r} - k_2\right)=0 \quad\mbox{in } \Omega \setminus \overline{D},
$$
we have from \eqref{ode2}
\begin{equation}\label{ode3}
w(r)= \frac{k}{6} r^2 + \frac{k_1}{r} + k_2 \quad\mbox{in } \Omega \setminus \overline{D}.
\end{equation}
Since $\pd{w}{r}=0$ on $\partial\Omega$, we must have
$$
\frac{k}{3} r - \frac{k_1}{r^{2}}= 0 \quad\mbox{on } \partial\Omega,
$$
and hence
$$
r^{3} = \frac{3k_{1}}k\quad\mbox{on } \partial\Omega.
$$
This means that $\partial\Omega = \partial B_{R}(\bold 0)$ for some $R >0$. Therefore we have
$$
\nabla w({\bf x})= \frac{k}{3} {\bf x} - \frac{kR^3}{3} \frac{{\bf x}}{r^3}, \quad {\bf x} \in \Omega \setminus \overline{D}.
$$
Since $\nabla w({\bf x})=c{\bf x}$ for all ${\bf x} \in \partial D$, we must have
$$
\frac{k}{3} - \frac{kR^3}{3} \frac{1}{r^3} = c \quad\mbox{on } \partial D,
$$
or $r=\mbox{constant}$ for all ${\bf x} \in \partial D$. It means that $\partial D$ is a sphere centered at $\bold 0$. This completes the proof. \hfill $\Box$
\section{Newtonian potential formulation}\label{sec:newton}
In this section we reformulate the problem \eqnref{free} in terms of the Newtonian potentials and relate the problem with known characterization of ellipsoids using the property of the Newtonian potential.
Suppose that \eqnref{free} admits a solution and let $w$ be the solution. Notice that by the second equation of \eqref{free} $w$ is constant on each connected component of $\partial\Omega$, and by the third equation of \eqref{free} $w({\bf x})= \frac{1}{2} {\bf x} \cdot {\bf A} {\bf x} + {\bf d}\cdot{\bf x} + C$ for ${\bf x} \in \partial D$ for some constant $C$.
Fix ${\bf x} \notin \overline{\Omega} \setminus D$. We obtain from the divergence theorem that
\begin{align*}
k \int_{\Omega \setminus D} \Gamma({\bf x}-{\bf y}) d{\bf y} &= \int_{\Omega \setminus D} \left[ \Delta w({\bf y}) \Gamma({\bf x}-{\bf y}) - w({\bf y}) \Delta_{\bf y} \Gamma({\bf x}-{\bf y}) \right] d{\bf y} \\
= -\int_{\partial D} &\left[ \pd{w}{\nu}({\bf y}) \Gamma({\bf x}-{\bf y}) - w({\bf y}) \pd{}{\nu_{\bf y}} \Gamma({\bf x}-{\bf y}) \right] d\sigma({\bf y}) - \int_{\partial \Omega} w({\bf y}) \pd{}{\nu_{\bf y}} \Gamma({\bf x}-{\bf y}) d\sigma({\bf y})\\
= -\int_{\partial D}& \left[ (\nu\cdot {\bf A} {\bf y} +\nu\cdot{\bf d}) \Gamma({\bf x}-{\bf y}) - (\frac{1}{2} {\bf y} \cdot {\bf A} {\bf y} + {\bf d}\cdot{\bf y} +C) \pd{}{\nu_{\bf y}} \Gamma({\bf x}-{\bf y}) \right] d\sigma({\bf y})\\
&- \int_{\partial \Omega} w({\bf y}) \pd{}{\nu_{\bf y}} \Gamma({\bf x}-{\bf y}) d\sigma({\bf y}).
\end{align*}
If ${\bf x} \in \mathbb{R}^3 \setminus \overline{\Omega}$, then by dealing with the last integral for each component of $\partial\Omega$ we have
$$
\int_{\partial \Omega} w({\bf y}) \pd{}{\nu_{\bf y}} \Gamma({\bf x}-{\bf y}) d\sigma({\bf y}) = 0,
$$
and hence
\begin{align*}
k \int_{\Omega \setminus D} \Gamma({\bf x}-{\bf y}) d{\bf y} = - \mbox{Tr}\, {\bf A} \int_{D} \Gamma({\bf x}-{\bf y}) d{\bf y}.
\end{align*}
We can find a relation between $k$ and $\mbox{Tr}\, {\bf A}$ from this formula. In fact, we have
$$
\lim_{|{\bf x}| \to \infty} \frac{1}{\Gamma({\bf x})} \int_{\Omega \setminus D} \Gamma({\bf x}-{\bf y}) d{\bf y} = |\Omega \setminus D|
$$
and a similar formula for $\int_{D} \Gamma({\bf x}-{\bf y}) d{\bf y}$. So we obtain
\begin{equation}
k |\Omega \setminus D|= -\mbox{Tr}\, {\bf A} |D|.
\end{equation}
We then see that
\begin{equation}
k \int_{\Omega \setminus D} \Gamma({\bf x}-{\bf y}) d{\bf y} + \mbox{Tr}\, {\bf A} \int_{D} \Gamma({\bf x}-{\bf y}) d{\bf y} = k|\Omega| \left[ \widehat N_\Omega({\bf x})- \widehat N_D({\bf x}) \right],
\end{equation}
where $\widehat N_\Omega$ and $\widehat N_D $ are the (averaged) Newtonian potentials on $\Omega$ and $D$, respectively, namely,
\begin{equation}
\widehat N_\Omega({\bf x}):= \frac{1}{|\Omega|} \int_{\Omega} \Gamma({\bf x}-{\bf y}) d{\bf y},
\end{equation}
and similarly for $\widehat N_D$.
If ${\bf x} \in D$, then we have for some constant $C^*$
\begin{align*}
k \int_{\Omega \setminus D} \Gamma({\bf x}-{\bf y}) d{\bf y} = - \mbox{Tr}\, {\bf A} \int_{D} \Gamma({\bf x}-{\bf y}) d{\bf y} + \frac{1}{2} {\bf x} \cdot {\bf A} {\bf x} + {\bf d} \cdot {\bf x} + C^*,
\end{align*}
and hence
\begin{equation}
k|\Omega| \left[ \widehat N_\Omega({\bf x})- \widehat N_D({\bf x}) \right] = \frac{1}{2} {\bf x} \cdot {\bf A} {\bf x} + {\bf d} \cdot {\bf x} + C^*.
\end{equation}
We have shown that if \eqnref{free} admits a solution, then
\begin{equation}\label{dive}
\widehat N_\Omega({\bf x})- \widehat N_D({\bf x}) =
\begin{cases}
0, \quad & {\bf x} \in \mathbb{R}^3 \setminus \Omega, \\
\mbox{a quadratic polynomial}, \quad & {\bf x} \in D.
\end{cases}
\end{equation}
So we may reformulate the question: If \eqnref{dive} holds, then $D$ and $\Omega$ are confocal ellipsoids. This is reminiscent of a question related to the Newton potential problem: If a Newtonian potential of a simply connected domain is a quadratic polynomial in the domain, then the domain must be an ellipsoid. This problem has been solved by Dive \cite{dive} and Nikliborc \cite{nikl} (see also \cite{DF} and \cite{km08}).
|
\section{Introduction}\label{sec:intro}
Given a compact Kähler manifold $(X,g)$ and a complex space $S$, a family of Hermite-Einstein bundles on $X$ parameterized by $S$ is a holomorphic vector bundle $F \to X \times S$ with a hermitian metric $h$ such that the restriction $h_s$ of $h$ to $F_s = F|X\times\{s\}$ is a Hermite-Einstein metric for every $s \in S$. If $p: X \times S \to S$ denotes the projection, the higher direct image sheaf $R^q p_* \mathcal{O}(F)$ for $q \in \mathbb{N}$ is locally free outside some analytic set of $S$ and carries a natural hermitian metric, fiberwise induced by the $L_2$ inner product of harmonic representatives.
If $\rho_k\in {\mathcal A} ^{0,1}(X,\mathrm{End}(F_s))$ denotes the harmonic representatives of Kodaira-Spencer classes, then for all $q$ there exist natural mappings
\begin{eqnarray*}
\rho_k \cup \textvisiblespace &:& {\mathcal A} ^{0,q}(X, F_s) \to {\mathcal A} ^{0,q+1}(X, F_s)\\
\rho^*_k \cap \textvisiblespace &:& {\mathcal A} ^{0,q}(X, F_s) \to {\mathcal A} ^{0,q-1}(X, F_s),
\end{eqnarray*}
where the second mappings are adjoint to the first.
Denote by $\xi_\sigma$ harmonic sections of ${\mathcal A} ^{0,q}(X, F_s)$. Then we denote the pointwise inner product of such sections (on a fiber $X\times \{s\}$) that is induced by $h_s$ and $g$ by $(\xi_\rho,\xi_\sigma)$ and the above $L_2$ inner product by
$$
\langle\xi_\rho,\xi_\sigma\rangle= \int_{X \times \{s\}} (\xi_\rho,\xi_\sigma)\frac{\omega^n}{n!},
$$
where $n$ is the dimension of $X$ and $\omega$ the Kähler form of $g$.
We assume that $S$ is smooth and compute the curvature of the natural metric on $R^q p_* \mathcal{O}(F)$. Under the above assumptions we have:
\begin{theorem}\label{thm:maintheorem}
Let $(F,h)$ be a holomorphic, hermitian vector bundle on $X \times S$ such that all vector bundles $F_s=F|X\times\{s\}$, $s\in S$ are simple, and all metrics $h_s$ are Hermite-Einstein. Then the curvature tensor of the induced metric on $R^qp_*{\mathcal O} (F)$ is given by:
\begin{align*}
R_{\rho\ol\sigma k \overline{l}}
&=
\langle
G
(
\rho_l^*
\cap
\xi_\rho
)
,
\rho_k^*
\cap
\xi_\sigma
\rangle \\
&\hphantom{=}+
\langle
G
\left(
\sqrt{-1}
\Lambda_g
\left[
\rho_k,
\rho_l^*
\right]
\right)
\xi_\rho
,
\xi_\sigma
\rangle \\
&\hphantom{=}-
\langle
G
(
\rho_k
\cup
\xi_\rho
),
\rho_l
\cup
\xi_\sigma
\rangle \\
&\hphantom{=}+
\langle
H(\rho_{k\overline{l}})\,\xi_\rho,
\xi_\sigma
\rangle .
\end{align*}
For $q=0$ the first, and for $q=\dim X$ the third summand vanishes.
\end{theorem}
The meaning of the function $H(\rho_{k \ol l})$ that depends only on the base parameter $s$ will be explained in Section~\ref{sec:curvature}.
The obvious key point about this formula is to express the curvature tensor in terms of intrinsically given data like harmonic Kodaira-Spencer classes. Technically this means the elimination of second order derivatives of the metric tensor when computing the curvature.
This result contains the case of direct image sheaves ($q=0$) which was solved in the work of To and Weng \cite{TW98}. Our main motivation is the study of the curvature of the Weil-Petersson metric on the moduli space of stable bundles in \cite{ST92} and also Berndtsson's positivity results for higher direct image sheaves \cite{Be09}, which was continued in the work of Berndtsson-P\u{a}un \cite{BP08} and Mourougane-Takayama \cite{MT08,MT09}.
At the end of section \ref{sec:curvature} we obtain the following results that are related to the Theorem~\ref{thm:maintheorem}. Namely, for families with fixed determinant we can rescale the Hermite-Einstein metrics locally and see that
$H(\rho_{k\overline{l}}|_{s_0})= 0$ so that the fourth summand in the curvature formula vanishes for such families. Furthermore, this fourth summand vanishes for induced families of endomorphism bundles without further assumptions.
This last case is especially interesting for applications to the moduli spaces of stable bundles. Let us consider the Kodaira-Spencer map
\[
\rho_s: T_s S \longrightarrow H^1(X, \mathcal{O}(\mathrm{End}(F_s)))
\]
for $s \in S$, associated to a family $(F,h) \to X \times S$ of hermitian bundles on $X$. For any complex tangent vector $v$ of $S$ at $s$ the Kodaira-Spencer map in terms of Dolbeault cohomology is given by
\begin{equation}\label{intro:overhaus}
\rho_s(v|_s) =
\left[ -( v \,\lrcorner\, \Omega^h)|_s \right]
\end{equation}
where $\Omega^h$ denotes the curvature form of $h$, and $|_s$ always denotes a restriction to $X\times \{s\}$. We use this notation unless obvious. Furthermore, the Hermite-Einstein condition implies that
$-( v \,\lrcorner\, \Omega^h)|_s
$
is {\em harmonic} \cite{Ov92,ST92}. In particular, one can read off this formula that the harmonic Kodaira-Spencer tensors depend in a differentiable way on the parameter. Altogether, we have a close relationship between the variation of the holomorphic structure on a complex vector bundle and the metric structure that is induced by $h$ on $F$ over $X\times S$.
For complex tangent vectors $v$ and $w$ on $S$ at $s$ the induced natural inner product reads
\[
\langle
v, w
\rangle _{\mathrm{WP}}(s)
=
\langle
\rho_s(v) , \rho_s(w)
\rangle(s)
=
\langle
\left.
(v \,\lrcorner\, \Omega^h)
\right|_s,
\left.
(w \,\lrcorner\, \Omega^h)
\right|_s
\rangle
\]
for $s \in S$. For effective families we get a hermitian metric on $S$ which is known to be Kähler. Moreover, this construction is functorial, i.e.\ compatible with base change and, as a consequence, descends to the corresponding moduli space of stable bundles. Due to its analogy with the Weil-Petersson metric on the Teichmüller space of Riemann surfaces, it is also called {\em Weil-Petersson metric}. In \cite{ST92} the curvature tensor of this metric was computed. The result follows as a special case of our main theorem. Under our assumptions the fibers of $R^1 p_* \mathcal{O}(\mathrm{End}(F))$ are $H^1(X, \mathcal{O}(\mathrm{End}(F_s)))$. So it is sufficient to apply the Theorem to the induced family of endomorphism bundles in the case $q=1$ and use the simplification for families of endomorphism bundles mentioned above.
{\bf Acknowledgement.} {\em This article is dedicated to Professor Yujiro Kawamata on the occasion of his birthday with respect and admiration.}
\section{The natural $L_2$ metric}\label{sec:l2metric}
By a theorem of Grauert, for any family of holomorphic vector bundles $(F,h) \to X \times S$ parameterized by a reduced space $S$, the dimension $h^q(X, \mathcal O(F_s))$ is constant on the complement of a certain analytic subset of $S$. Furthermore, the sheaf $R^q p_* \mathcal{O}(F)$ is locally free over this complement and the natural morphism
\begin{equation}\label{l2metric:fiberofimagesheave}
(R^q p_* \mathcal{O}(F))_s
\otimes_{\mathcal{O}_{S,s}}
\mathbb{C}(s)
\stackrel{\sim}{\longrightarrow}
H^q(X, \mathcal{O}(F_s))
\end{equation}
is an isomorphism by the base-change theorem (cf.\ \cite{BS76}). Given the Kähler structure on $X$ and the hermitian metric $h$ on $F$ over $X\times S$ we identify the cohomology groups with the spaces $\mathcal{H}^{0,q}(X, F_s)$ of forms that are harmonic with respect to $g$ and $h_s$. On these we have a natural inner product, given by
\[
\langle \mu, \eta \rangle
=
\int_X
(\mu, \eta)
\,\frac{\omega^n}{n!}.
\]
Since our arguments will be local with respect to the parameter space $S$, we may assume that $S$ is Stein and identify sections of direct image sheaves with cohomology classes in $H^q(X\times S, \mathcal{O}(F))$. On one hand, these are represented by {\em $\ol{\partial}$-closed forms} over $X\times S$, on the other hand, by {\em families of harmonic forms} in $\mathcal H^{0,q}(X,F_s)$. The following lemma (cf.\ \cite[Lemma 2]{Sch12}) shows that both properties can be achieved simultaneously, a fact that is necessary for later computations.
\begin{lemma}\label{lemma:correctionforsections}
Let $(F,h) \to X \times S$ be a family of hermitian bundles on a compact Kähler manifold $(X,g)$ parameterized by a complex manifold $S$ and assume that $R^q p_*\mathcal O(F)$ is locally free. If
$
\phi\in\mathcal{A}^{0,q}(X\times S, F)
$
is $\ol{\partial}$-closed and $s_0 \in S$ any point, then there exists a form $\chi\in \mathcal{A}^{0,q-1}(X\times V,F) $ on some open neighborhood $V \subset S$ of $s_0$ such that
\[
(\phi + \ol{\partial} \chi)|_s = H(\phi|_s)
\]
for every $s \in V$, where $H$ is the harmonic projection for the fiber $F_s$. In particular, any class in $H^q(X\times S, \mathcal O(F))$ can be represented by a $\ol{\partial}$-closed form, whose restrictions to all fibers of $s\in V$ are harmonic.
\end{lemma}
\begin{proof}
First, there exist relative $(0,q)$ forms
$
\psi_G^\prime
$
and
$
\psi_H^\prime
$
along the fibers $F_s \to X$ with $\psi_G^\prime|_s = G(\phi|_s)$ as well as
$\psi_H^\prime|_s = H(\phi|_s)$ for every $s \in S$, where $G$ denotes the Green operator on $F_s$-valued forms. In fact, this is a consequence of our assumption and follows from the fundamental theorem in \cite{KS58}, proved in \cite[Theorem 7]{KS60}. On some open neighborhood $V$ of $s_0$ the relative form
$
\psi_H^\prime
$
as well as
$
\ol{\partial}^*_{\mathrm{rel}}
\psi_G^\prime
$,
where
$
\ol{\partial}^*_{\mathrm{rel}}
$
is the relative $\ol{\partial}^*$ operator, are induced by forms
$
\psi_H
\in
\mathcal{A}^{0,q}(X \times V, F)
$
and
$
\psi_G
\in
\mathcal{A}^{0,q-1}(X \times V, F)
$,
respectively. Initially the extensions of the relative forms can be taken locally, and then glued together by a partition of unity. We get $\psi_G|_s = \ol{\partial}^* G(\phi|_s)$ and also $\psi_H|_s = H(\phi|_s)$ for every $s \in V$. Using $\ol{\partial} \phi = 0$ we obtain
\[
\phi|_s
=
H(\phi|_s)+\ol{\partial}\ol{\partial}^*G(\phi|_s)
\]
and, as a consequence, by choosing $\chi := - \psi_G$ we finally have
\[
(\phi + \ol{\partial}\chi)|_s
=
\phi|_s -
\ol{\partial}(\psi_G|_s) = H(\phi|_s).
\]
\end{proof}
\section{Computation of the curvature}\label{sec:curvature}
Again we will always assume that all $R^qp_*\mathcal O(F)$ are locally free. Although some statements are possible for reduced base spaces, we will also assume that $S$ is smooth. Since curvature computations are local with respect to the base, we can assume without loss of generality that $S$ is Stein with local coordinates $(s^1,\ldots,s^m)$ and replace the space $S$ by a neighborhood of a given point $s_0\in S$, if necessary.
We will use holomorphic coordinates $(z^1,\ldots,z^n)$ for the Kähler manifold $X$ and denote the Kähler form by
$$
\omega=\frac{\sqrt{-1}}{2}g_{\alpha\ol \beta} dz^\alpha\wedge dz^{\ol{\beta}}.
$$
Greek indices refer to the $z$-coordinates on $X$ (or to sections of the given vector bundle), whereas Latin indices are reserved for $s$-coordinates.
Let
$
\Xi_1, \ldots, \Xi_R
\in R^q p_* \mathcal{O}(F)(S)
$
define a holomorphic frame so that the inner product is given by $H_{\tau\ol\sigma}$. Then the curvature form $\Omega \in \mathcal A^{1,1}(X \times S,{\rm End}(R^qp_*\mathcal O(F)))$ is equal to
$$
\Omega = R^{\;\tau}_{\rho\; k \ol l} ds^k\wedge ds^{\ol l},
$$
in $s$-coordinates, and we have
$$
R_{\rho\ol\sigma k \ol l} = H_{\tau\ol\sigma}R^{\;\tau}_{\rho\; k \ol l}.
$$
Furthermore, we will work in normal coordinates at a given point $s_0$. After replacing $S$ by a neighborhood of $s_0$, if necessary, we can assume that
$$
H_{\tau \ol \sigma}(s_0)= \delta^\sigma_\tau \text{ and } \frac{{\partial }}{{\partial } s^k}H_{\rho\ol \sigma}(s_0)=0 \text{ for all } k,\sigma, \rho.
$$
We apply Lemma~\ref{lemma:correctionforsections} which provides representatives
$$
\xi_1, \ldots, \xi_R
\in
\mathcal{A}^{0,q}(X \times S,F)
$$
of the Dolbeault cohomology classes of $\Xi_\rho$, such that the restrictions $\xi_\rho|_s$ are {\em harmonic}. We note
\begin{equation}\label{curvature:hrhosigma}
H_{\rho\overline{\sigma}}(s)
=
\langle
\Xi_\rho, \Xi_\sigma
\rangle (s)
=
\langle
\xi_\rho, \xi_\sigma
\rangle (s).
\end{equation}
With respect to the given normal coordinates, we get
\begin{equation}\label{curvature:curvature}
R_{\rho\ol \sigma k \ol l}(s_0)= - {\partial }_{\ol l}{\partial }_k H_{\rho \ol \sigma}(s_0) = - {\partial }_{\ol l}{\partial }_k \langle \xi_\rho, \xi_\sigma\rangle(s_0)
\end{equation}
for the curvature on the direct image sheaves.
For computational reasons we equip $S$ with the flat metric with respect to the coordinates $s$ and $X \times S$ with the product metric. This convention is not essential, but we are in a position to use covariant derivatives for tensors with values in $F$ over the total space $X\times S$.
We will use the following notation for differentiable sections $\chi$ of $F$ over $X\times S$.
The connection form of $h$ is
$$
\theta^h = {\partial } h \cdot h^{-1}
$$
so that
$$
\nabla_\alpha \chi = {\partial }_\alpha \chi + \theta^h_\alpha\circ \chi, \text{ and } \nabla_k \chi = {\partial }_k \chi + \theta^h_k\circ \chi \text{ resp.}
$$
(and $\nabla_\ol\beta={\partial }_\ol \beta$, $\nabla_\ol l = {\partial }_\ol l$).
Now
\begin{equation}\label{eq:curvh}
\nabla_\ol{\beta}\nabla_\alpha \chi = \nabla_\alpha\nabla_\ol{\beta}\chi - R^h_{\alpha \ol{\beta}} \circ \chi
\end{equation}
where $R^h_{\alpha \ol{\beta}}= -\nabla_{\ol{\beta}}\theta^h_\alpha$ etc.\ denote the components of the curvature tensor of $h$ with values in ${\rm End}(F)$ over $X \times S$. For the components in base direction the analogous equations hold.
For later use we note that for the metric induced by $h$ on ${\rm End}(F)$ and differentiable sections $\zeta$ the above formula reads
\begin{equation}\label{eq:curvhend}
\nabla_\ol{\beta}\nabla_\alpha \zeta = \nabla_\alpha\nabla_\ol{\beta}\zeta - [R^h_{\alpha \ol{\beta}} , \zeta].
\end{equation}
In this section $\eta$ and $\mu$ will denote $F$-valued $(0,q)$-forms. The following construction will be essential:
\begin{lemma}\label{lemma:covarianttoexterior}
Let $\mu\in {\mathcal A} ^{0,q}(X\times S,F)$, $q > 0$ with $\ol{\partial} \mu = 0$. Then for every $1 \leq l \leq m$ there exists a form
$
F_{\overline{l}}(\mu)
\in
\mathcal{A}^{0,q-1}(X \times S,F)
$
satisfying the following equation for every $s \in S$:
\[
\ol{\partial}(F_{\overline{l}}(\mu)|_s)
=(\nabla_{\overline{l}}\mu)|_s.
\]
In other words, the derivative in a (conjugate) parameter direction of a $\ol{\partial }$-closed form, restricted to a fiber is $\ol{\partial }$-exact.
\end{lemma}
\begin{proof}
We write
\begin{align*
\mu
&=
\frac{1}{q!}
\sum_{\beta} \mu_{\overline{\beta}_1,\ldots,\overline{\beta}_q} dz^{\overline{\beta}_1} \wedge \ldots \wedge dz^{\overline{\beta}_q} \\\nonumber
&\hphantom{=}+
\frac{1}{(q-1)!}
\sum_{\beta,k} \mu_{\overline{\beta}_1,\ldots,\overline{\beta}_{q-1},\overline{k}} dz^{\overline{\beta}_1} \wedge \ldots \wedge dz^{\overline{\beta}_{q-1}} \wedge ds^{\overline{k}} \\\nonumber
&\hphantom{=}+
\mbox{summands with more than one } \overline{d s} \mbox{-factor}.
\end{align*}
Because $\ol{\partial} \mu = 0$, for the fixed $l$ the factors
$
dz^{\overline\beta_1} \wedge \ldots \wedge dz^{\overline\beta_q} \wedge ds^{\overline{\ell}}
$
yield:
\begin{align}\label{lemma:covarianttoexterior:p1}
&\frac{1}{q!}
\sum_\beta
\partial_{\overline{l}}
\mu_{\overline{\beta}_1,\ldots,\overline{\beta}_q}
dz^{\overline{\beta}_1} \wedge \ldots \wedge dz^{\overline{\beta}_q}\\\nonumber
&\hspace{1cm}=
\frac{(-1)^{q+1}}{q!}
\sum_\beta
\sum_\nu
(-1)^{\nu+1}
\partial_{\overline{\beta_\nu}}
\mu_{\overline{\beta}_1,\ldots,\widehat{\overline{\beta}_\nu},\ldots,\overline{\beta}_q,\overline{l}}
dz^{\overline{\beta}_1} \wedge \ldots \wedge dz^{\overline{\beta}_q}.
\end{align}
After restricting to $X\times \{s\}$ the left hand side of \eqref{lemma:covarianttoexterior:p1} equals
$
\left.
\left(
\nabla_{\overline{l}} \mu
\right)
\right|_s
$.
So
\begin{equation}\label{lemma:covarianttoexterior:p2}
F_{\ol l}(\mu)
:=
\frac{(-1)^{q+1}}{(q-1)!}
\sum_\beta
\mu_{\overline{\beta}_1,\ldots,\overline{\beta}_{q-1},\overline{l}}
dz^{\overline{\beta}_1} \wedge \ldots \wedge dz^{\overline{\beta}_{q-1}}
\end{equation}
has the desired properties.
\end{proof}
We will use the notion $\ol{\partial }^*$ only for differential forms {\em on the fibers} $X\times\{s\}$ with respect to $g$ and $h_s$ (with variable parameter). In this sense we have a first application:
\begin{corollary}\label{corollary:vanishingofproduct}
If $\ol{\partial}\eta = 0$ and
$
\ol{\partial}^*(\mu|_s) = 0
$
for some $s \in S$, then also
$
\langle
\mu, \nabla_{\overline{k}} \eta
\rangle (s)
=
0
$.
\end{corollary}
\begin{proof}
In the case $q=0$ the result is clear because $\eta$ is holomorphic. In the case $q >0$ we can apply Lemma \ref{lemma:covarianttoexterior} to get on all fibers $X\times \{s\}$:
\[\langle \mu,\nabla_{\overline{k}} \eta \rangle(s) = \langle \mu,\ol{\partial}(F_{\overline{k}}\eta)\rangle(s)
=\langle \ol{\partial}^*\! \mu, F_{\overline{k}}(\eta)\rangle (s) = 0.
\]
\end{proof}
We also obtain a formula for the second order derivatives:
\begin{corollary}\label{corollary:secondderivatives}
If we have $\ol{\partial}\eta = 0$ and
$
\ol{\partial}^*(\mu|_s) = 0
$
on all fibers we get:
\[
\partial_{\overline{l}} \partial_k \langle \mu, \eta\rangle = \langle \nabla_k \mu, \nabla_l \eta\rangle + \langle \nabla_{\overline{l}}\nabla_k \mu, \eta \rangle .
\]
\end{corollary}
The values of
$
\ol{\partial}^*\!(\nabla_k \mu|_s)
$
as well as
$
\ol{\partial}(\nabla_k \mu|_s)
$
are of particular importance for the computation of the curvature. For the first type we find:
\begin{lemma}\label{lemma:dbarstarofnablamu}
If
$
\ol{\partial}^*(\mu|_s) = 0
$
for every $s \in S$ then also
\[
\ol{\partial}^*(\nabla_k \mu|_s) = 0
\]
holds for every $s \in S$.
\end{lemma}
\begin{proof}
Since the connection and curvature forms of $h$ in the direction of $X$ restricted to $X\times \{s\}$ equal the connection and curvature forms resp.\ of $h_s$, we have
\begin{gather*}
g^{\ol{\delta \gamma}}\nabla_\gamma(\nabla_k \mu_{\ol{\beta}_1,\ldots,\ol{\beta}_{q-1}|\ol{ \delta}}|_s )= (g^{\ol{\delta \gamma}}\nabla_k\nabla_\gamma \mu_{\ol{\beta}_1,\ldots,\ol{\beta}_{q-1}|\ol{ \delta}})|_s\\ \hspace{5cm} = (\nabla_k (g^{\ol{\delta} \gamma} \nabla_\gamma \mu_{\ol{\beta}_1,\ldots,\ol{\beta}_{q-1}|\ol{ \delta}}) )|_s
\end{gather*}
which implies the claim.
\end{proof}
In order to compute expressions of the second type $\ol{\partial}(\nabla_k \mu|_s)$ we have to introduce some more notation. If $A$ is a $(p,q)$ form with values in some endomorphism bundle $\mathrm{End}(E)$, we agree to denote with $A \cup$ the operator on $(r,s)$ forms with values in $E$ which consists of the application of the endomorphism part and an exterior multiplication of the form parts. We will use $A^* \cap$ to denote the formal adjoint of $A \cup$.
In local coordinates we need the following case: Let $A=\sum A_\ol \delta dz^{\ol \delta}$ be an
${\rm End}(E)$-valued $(0,1)$-form, and $\mu=(1/q!) \sum \mu_{\ol\beta_1,\ldots,\ol \beta_q}dz^{\ol \beta_1}\wedge\ldots\wedge dz^{\ol \beta_q}$ an $E$-valued $(0,q)$-form. Then
\begin{equation}\label{eq:cup}
A\cup \mu = \frac{1}{(q+1)!} \sum A_{\ol\beta_0}(\mu_{\ol\beta_1,\ldots,\ol \beta_q})dz^{\ol \beta_0}\wedge\ldots\wedge dz^{\ol \beta_q},
\end{equation}
and for an $E$-valued $(0,q+1)$-form $\sigma$ (whose coefficients are already assumed to be skew-symmetric) we have
\begin{equation}\label{eq:cap}
A^* \cap \sigma =\sum g^{\ol \delta \gamma} A^*_{\gamma}( \sigma_{\ol \delta, \ol \beta_1,\ldots, \ol\beta_q}) dz^{\ol \beta_1}\wedge\ldots dz^{\ol \beta_q}.
\end{equation}
Furthermore, we will use the abbreviations
$$
\rho_k
:=
- \partial_k \,\lrcorner\, \Omega^h
\text{\quad and \quad}
\rho_{k\overline{l}}
:=
\partial_{\overline{l}} \,\lrcorner\, (\partial_k \,\lrcorner\, \Omega^h)
$$
as well as some obvious variations on these, where again $\Omega^h$ denotes the curvature form of $h$. Note that, in particular, we can use these to describe the harmonic representatives of Kodaira-Spencer classes, because \eqref{intro:overhaus} implies that in our notation the harmonic representative $\rho_s(\partial_k|_s)$ equals $\rho_k|_s$.
After this remark we obtain:
\begin{lemma}\label{lemma:dbarofnablamu}
Let $\mu\in{\mathcal A} ^{0,q}(X\times S,F)$ with $\ol{\partial}(\mu|_s)=0$ for every $s \in S$. Then
\[
\ol{\partial}(\nabla_k \mu|_s) = \rho_k|_s \cup \mu|_s
\]
for every $s \in S$.
\end{lemma}
\begin{proof}
We observe that
$$
\nabla_{\ol{\beta}_{0}}\nabla_k\, \mu_{\ol \beta_1, \ldots, \ol \beta_q} = \nabla_k\nabla_{\ol \beta_{0}} \mu_{\ol \beta_1, \ldots, \ol \beta_q} - R^h_{k\ol \beta_{0}}( \mu_{\ol \beta_1,\ldots,\ol\beta_q})
$$
and replace
\begin{equation}\label{eq:harmrep}
R^h_{k \ol{\beta}} dz^\ol\beta = {\partial }_k\, \lrcorner\, \Omega^h= - \rho_k.
\end{equation}
\end{proof}
We continue the calculation of the curvature. First we know from the construction of the good representatives $\xi_\rho$ of the Dolbeault classes that the restrictions
$
\left.
\xi_\rho
\right|_s
$
are harmonic and the above identities are applicable. In particular, Corollary \ref{corollary:secondderivatives} together with \eqref{curvature:curvature} implies
\[
R_{\rho\sigma k \overline{l}}(s_0)
=
-
\big(
\langle
\nabla_k \xi_\rho, \nabla_l \xi_\sigma
\rangle
(s_0)
+
\langle
\nabla_{\overline{l}} \nabla_k \xi_\rho, \xi_\sigma
\rangle
(s_0)
\big)
=:
-(S_1 + S_2).
\]
Using Corollary~\ref{corollary:vanishingofproduct} and the fact that we chose normal coordinates at $s_0$ we find
\[
\langle \nabla_k \xi_\rho|_{s_0}, \xi_\sigma|_{s_0}\rangle =\partial_k\langle\xi_\rho, \xi_\sigma\rangle(s_0) = \partial_k H_{\rho\overline{\sigma}}(s_0) = 0
\]
i.e.\ the derivatives $\nabla_k\xi_\rho|_{s_0}$ are perpendicular to the space of harmonic forms $\mathcal H^{0,q}(X,F_{s_0})$ that is spanned by all forms $\xi_\sigma|_{s_0}$.
Hence, the harmonic projections $H(\nabla_k \xi_\rho|_{s_0})$ vanish so that by Lemma~\ref{lemma:dbarstarofnablamu}
\[ \nabla_k \xi_\rho|_{s_0} =\ol{\partial} G \ol{\partial}^*(\nabla_k \xi_\rho|_{s_0}) + \ol{\partial}^* G\ol{\partial}(\nabla_k \xi_\rho|_{s_0}) = \ol{\partial}^* G \ol{\partial}( \nabla_k \xi_\rho|_{s_0}),
\]
where $G$ denotes the respective Green operator. As a consequence we establish for the first summand $S_1$
\begin{align*}
S_1
&= \langle \nabla_k \xi_\rho, \nabla_l \xi_\sigma\rangle(s_0) = \big\langle \ol{\partial}^* G \ol{\partial}(\nabla_k \xi_\rho),\nabla_l \xi_\sigma \big\rangle (s_0) \\
&= \big\langle G \ol{\partial}( \nabla_k \xi_\rho),\ol{\partial} \nabla_l \xi_\sigma\big\rangle(s_0) \\
&= \big\langle G(\rho_k\cup\xi_\rho), \rho_l\cup\xi_\sigma \big\rangle(s_0)
\end{align*}
where we used Lemma \ref{lemma:dbarofnablamu} in the last step.
In order to calculate the second summand $S_2$ we first proceed with collecting further relations. For this purpose we note the following local formula
\begin{equation}\label{curvature:liebracketcovariantderiv}
\nabla_k \nabla_{\overline{l}} \,\mu - \nabla_{\overline{l}} \nabla_k\, \mu
=
\rho_{k\overline{l}} \cup \mu
\end{equation}
which follows from \eqref{eq:curvh} and the definition of $\rho_{k\overline{l}}$. (It only depends on $h$ because of the choice of the Kähler metric on $X\times S$). We denote by $\Box$ the Laplacian (with non-negative eigenvalues) for $F_s$-valued $(0,q)$-forms.
\begin{lemma}\label{lemma:laplacerhokl}
In our situation on all fibers $X\times \{s\}$ the equation
\[\Box( \rho_{k\overline{l}}) =\sqrt{-1} \Lambda_g\left[\rho_k, \rho_{\overline{l}}
\right]
\]
holds.
\end{lemma}
\begin{proof}
Let $R^h$ be the ${\rm End}(F)$-valued curvature tensor of $h$ over $X\times S$ as above.
The quantity $R^h_{k\ol l}$\/, when restricted to a fiber, has to be treated as a differentiable section of the endomorphism bundle of $F_s$. We compute
\begin{align*}
\Box(R^h_{k\ol l}) &= \ol{\partial }^* \ol{\partial } R^h_{k\ol l} = - g^{\ol \delta \gamma} \nabla_\gamma\nabla_\ol\delta R^h_{k \ol l}= - g^{\ol \delta \gamma} \nabla_\gamma\nabla_\ol l R^h_{k \ol \delta}\\
&=- g^{\ol \delta \gamma} [R^h_{\gamma \ol l}, R^h_{k \ol \delta}] - g^{\ol \delta \gamma} \nabla_\ol l\nabla_k R^h_{\gamma \ol \delta}.
\end{align*}
The very last term vanishes because of the Hermite-Einstein condition, since the degree of the bundles is constant. So the claim follows as above from \eqref{eq:harmrep}.
\end{proof}
We need two properties of the forms $F_{\overline{l}}(\mu)$ from Lemma \ref{lemma:covarianttoexterior}:
\begin{lemma}\label{lemma:property2ofCTEconstruction}
If $\ol{\partial} \mu = 0$ and $\ol{\partial}^*\!\mu|_s = 0$
for every $s \in S$, we obtain fiberwise for all $s$
\[
\ol{\partial}^* \ol{\partial}(F_{\overline{l}}(\mu)) =\rho^*_{l} \cap \mu.
\]
\end{lemma}
\begin{proof}
We have
\begin{align*}
\ol{\partial }^*\!\ol{\partial }(F_\ol l(\mu))&=\ol{\partial }^*\! (\nabla_\ol l \mu) \\
& =(-1)^q \sum g^{\ol \delta \gamma}\nabla_\gamma\nabla_\ol l \mu_{\ol \beta_1, \ldots, \ol \beta_{q-1},\ol \delta}\, dz^{\ol \beta_1}\wedge \ldots \wedge dz^{\ol \beta_{q-1}}\\
&=(-1)^{q+1}\sum g^{\ol \delta \gamma} R^h_{\gamma\ol l}(\mu_{\ol\beta_1,\ldots,\ol \beta_{q-1},\ol \delta})\, dz^{\ol \beta_1}\wedge \ldots \wedge dz^{\ol \beta_{q-1}}\\
&= - \sum g^{\ol \delta \gamma}R^h_{\gamma\ol l} (\mu_{\ol \delta, \ol \beta_1,\ldots, \ol \beta_{q-1}}) dz^{\ol \beta_1}\wedge\ldots\wedge dz^{\ol \beta_{q-1}}
\end{align*}
so that the claim follows with \eqref{eq:cap}.
\end{proof}
\begin{lemma}\label{lemma:property1ofCTEconstruction}
Let $\mu$ be a $\ol{\partial }$-closed $(0,q)$-form with values in $F$. Then on all fibers $X\times \{s\}$ the following identity holds.
\[(\nabla_k \nabla_{\overline{l}}\,\mu) =\ol{\partial}(\nabla_k F_{\overline{l}}(\mu)) -(\rho_k \cup F_{\overline{l}}(\mu)).
\]
\end{lemma}
\begin{proof}
We pick up the statement of Lemma \ref{lemma:covarianttoexterior} and compute
$\nabla_k \ol{\partial } F_\ol l(\mu)$. The claim follows from \eqref{eq:curvhend}.
\end{proof}
Now we will finish the proof of Theorem~\ref{thm:maintheorem} by treating the summand $S_2$. The first step is to use \eqref{curvature:liebracketcovariantderiv} and split $S_2$ into two parts. For the inner product taken at $s_0$ we have
\begin{align*}
\nonumber
S_2
&=
\langle
\nabla_{\overline{l}} \nabla_k \xi_\rho, \xi_\sigma
\rangle
=
\langle
\nabla_k \nabla_{\overline{l}} \xi_\rho, \xi_\sigma
\rangle
-
\langle
\rho_{k\overline{l}} \cup \xi_\rho, \xi_\sigma
\rangle \\
&=:
S_{2a} + S_{2b}\, .
\end{align*}
For $S_{2a}$ we get from Lemma \ref{lemma:property1ofCTEconstruction} that
\begin{align*}
S_{2a}
&=
\langle
\ol{\partial}( \nabla_k F_{\overline{l}}(\xi_\rho)), \xi_\sigma\rangle -\langle \rho_k\cup F_{\overline{l}}(\xi_\rho), \xi_\sigma \rangle \\
&=
\big\langle (\nabla_k F_{\overline{l}}(\xi_\rho)), \ol{\partial}^* \xi_\sigma \big\rangle - \langle F_{\overline{l}}(\xi_\rho), \rho^*_k \cap \xi_\sigma \rangle
\end{align*}
holds on $X\times \{s_0\}$.
By assumption the forms $\xi_\rho$ are fiberwise harmonic: $\ol{\partial}^*\xi_\sigma = 0$. We apply Lemma~\ref{lemma:property2ofCTEconstruction}, and again at $s_0$ we obtain
\begin{align*}
S_{2a}&= - \langle F_{\overline{l}}(\xi_\rho),\ol{\partial}^* \ol{\partial}(F_{\overline{k}}(\xi_\sigma))\rangle =
- \langle \ol{\partial} F_{\overline{l}}(\xi_\rho), \ol{\partial}(F_{\overline{k}}(\xi_\sigma))\rangle .
\end{align*}
We use the fiberwise equation
\[
\ol{\partial} F_{\overline{l}}(\xi_\rho) = \ol{\partial} G\ol{\partial}^* \ol{\partial} F_{\overline{l}}(\xi_\rho)
\]
and, consequently, from Lemma~\ref{lemma:property2ofCTEconstruction} we get:
\begin{align*}
S_{2a} &=-\big\langle G \ol{\partial}^*\ol{\partial} F_{\overline{l}}(\xi_\rho),\ol{\partial}^*\ol{\partial}(F_{\overline{k}}(\xi_\sigma))\big\rangle =-\big\langle G(\rho_{l}^* \cap \xi_\rho),\rho_{k}^* \cap\xi_\sigma\big\rangle.
\end{align*}
For the last summand $S_{2b}$ we apply Lemma \ref{lemma:laplacerhokl} to obtain fiberwise
\begin{align*}
\rho_{k\overline{l}}&=
H(\rho_{k\overline{l}}) + G (\sqrt{-1} \Lambda_g \left[\rho_k,\rho_{\overline{l}}\right]).
\end{align*}
Hence, again fiberwise
\begin{align*}
S_{2b}
&=
-
\langle
\rho_{k\overline{l}}
\cup
\xi_\rho,
\xi_\sigma
\rangle \\
&=
-
\langle
G
\left(
\sqrt{-1}
\Lambda_g
\left[
\rho_k,
\rho_{\overline{l}}
\right]
\right)
\cdot
\xi_\rho,
\xi_\sigma
\rangle
-
\langle
H
\left(
\rho_{k\overline{l}}
\right)
\cdot
\xi_\rho,
\xi_\sigma
\rangle.
\end{align*}
This concludes the proof of Theorem \ref{thm:maintheorem}. \qed
In the sequel we discuss the above term $H(\rho_{k\overline{l}})$. Note that $H(\rho_{k\overline{l}})(s)$ is a harmonic section of ${\rm End}(F_s)$ i.e.\ a constant multiple of the identity, since the holomorphic vector bundles $F_s$ are simple by assumption. So $H(\rho_{k\overline{l}})$ can be identified with a differentiable function on $S$. This argument yields the following lemma.
\begin{lemma}\label{le:Hkl}
Let $r= {\rm rk}(F)$. Then the ${\rm End}(F_s)$-valued harmonic sections $H(\rho_{k\overline{l}})(s)$ are of the form
$\Phi_{k\ol l}(s) \cdot {\rm id}_{F_s}$ satisfying the equation
\begin{equation}\label{eq:Hkl}
\Phi_{k\ol l}(s) = \frac{1}{r} \int_{X \times\{s\}} {\rm tr}(R^h_{k \ol l}) \, \omega^n\Big/\int_X \omega^n.
\end{equation}
\end{lemma}
Since Hermite-Einstein metrics are only unique up to a constant positive factor, the given metrics $h$ on $F$ over $X\times S$ may be rescaled, i.e.\ modified by a factor ${\rm exp}(\varphi(s))$, where $\varphi$ denotes a differentiable function:
\begin{proposition}
Let $(F,h) \to X \times S$ be a holomorphic, hermitian vector bundle such that the restrictions $(F_s,h_s)$ are simple Hermite-Einstein vector bundles over a compact Kähler manifold $(X,g)$ such that all ${\rm det}(F_s)$ are isomorphic to a fixed line bundle $L$ on $X$.
Then locally with respect to $S$ the metric $h$ can be rescaled such that all harmonic projections $H(\rho_{k \ol l})(s)$ vanish.
\end{proposition}
\begin{proof}
Let $q:X\times S \to X$ be the canonical projection. We may assume that $\det F \otimes q^*L^{-1}$ is trivial. We equip $L$ with an auxiliary hermitian metric $h_L$ and consider $\det h \cdot q^*(h_L^{-1})$, which is of the form ${\rm exp}(\chi)$. Now
$$
{\rm tr}(R^h_{k\ol l})= -\frac{{\partial }^2 \log \det h}{{\partial } s^k{\partial } s^{\ol l}} = -\frac{{\partial }^2 \chi}{{\partial } s^k{\partial } s^{\ol l}},
$$
since the extra additive term involving $\log h_L$ does not depend upon $s$. Now the components $\Phi_{k\ol l}(s)$ in the sense of \eqref{eq:Hkl} are of the form
$$
\Phi_{k\ol l}(s)= -\frac{{\partial }^2 \varphi(s)}{{\partial } s^k {\partial } s^\ol l},
$$
where
$$
\varphi(s)= \frac{1}{r}\int_{X\times \{s\}}\chi\, \omega^n \Big/ \int_X \omega^n.
$$
We replace $h$ by ${\rm exp }(-\varphi)h$, which yields the claim.
\end{proof}
If $S$ is an arbitrary base space, the bundle $\det F \otimes q^*L^{-1}$ is of the form $p^*M$, where $M$ is a holomorphic line bundle on $S$. With the same methods one can see immediately the following somewhat more general fact.
\begin{proposition}
For any hermitian metric $h_M$ on $M$, there is a hermitian metric $h$ on $F$ that restricts to a family of Hermite-Einstein metrics $h_s$ such that
$$
\sqrt{-1}H(\rho_{k \ol l})ds^k \wedge ds^\ol l = - \sqrt{-1} {\partial } \ol{\partial } \log h_M
$$
on $S$.
\end{proposition}
In a moduli theoretic situation of a family with $\det(F_s)\simeq L$ for all $s$, after replacing $S$ by a finite unbranched covering $\pi:\widetilde S \to S$, there exists a line bundle $\widetilde M$ on $\widetilde S$ such that $\widetilde M^{\otimes r} = \pi^* M$. Now the bundle $F$ can be replaced by $\widetilde F= \pi^* F \otimes \widetilde q^*\widetilde M^{-1}$ where $\widetilde q:X\times \widetilde S \to \widetilde S$ is the canonical projection. The isomorphism classes of the fibers are unchanged.
\begin{corollary}
In the above situation, the bundle $\widetilde F$ possesses a family of Hermite-Einstein metrics such that everywhere all $H(\rho_{k \ol l})(s)$ vanish.
\end{corollary}
Let a family $(F,h)$ of (simple) holomorphic Hermite-Einstein bundles be given. We have the induced family of Hermite-Einstein metrics $\widetilde h_s$ on ${\rm End}(F_s)$, and the connection and curvature forms on ${\rm End}(F_s)$ are given by $[\theta^h_s, \,\,]$ and $[\Omega^h_s, \,\,]$ resp.\ so that ${\rm tr}(R^{\widetilde h}_{k \ol l})=0$. This fact implies that the induced term for the curvature formula for the direct images vanishes. Hence the following proposition holds.
\begin{proposition}
Given a holomorphic family of simple, holomorphic Hermite-Einstein bundles, equip the endomorphism bundles with the induced structure. Then the fiberwise harmonic projections $H(\rho_{k \ol l})$ for the sheaves $R^qp_*({\rm End}(F))$ vanish identically.
\end{proposition}
|
\subsection{Taxonomy of proposed extensions of general relativity}
To frame external tests in terms of hypothesis testing, one would like
to have one or more valid alternatives to GR. What constitutes a
``valid alternative'' is, of course, a matter of taste. From our
perspective (i.e., in terms of tests of strong-field gravity) the
alternative should be a cosmologically viable fundamental theory
passing intermediate energy tests, with a well-posed initial value
formulation, and field equations that follow from an action
principle. Furthermore, the theory should be simple enough to make
definite, calculable prediction in the strong-field regime: ideally,
it should allow us to predict the structure and dynamics of compact
objects and the gravitational radiation that they emit, whether
isolated or in binary systems.
This is a very stringent set of requirements. There are countless
attempts to modify
GR~\cite{Sotiriou:2008rp,DeFelice:2010aj,Nojiri:2010wj,Capozziello:2011et,Clifton:2011jh,Hinterbichler:2011tt,deRham:2014zqa},
but (for the reasons listed above) in several cases the modifications
introduce some screening mechanism in order to be viable at
intermediate energies. Screening mechanisms include chameleons,
symmetrons, dilatons,
MOND-like dynamics,
the Vainshtein mechanism, etcetera,
depending on whether the screening is set by the local value of the
field or by its derivatives~\cite{Joyce:2014kja}.
Chapter~\ref{sec:alt-th} reviews various theories that have been
explored in some detail as phenomenological alternatives to GR in the
strong-field regime.
The chapter begins with a discussion of Lovelock's theorem, a
``uniqueness theorem'' for the field equations of GR. Uniqueness is
based on a small set of definite assumptions. The interest of
Lovelock's theorem from a pragmatic point of view is that it can be
``turned around,'' and used to classify extensions of GR based on
which of the underlying assumptions of Lovelock's theorem they
violate. Within this classification framework, we list and discuss
several theories that have been seriously considered as plausible
alternatives to GR in the context of strong-field tests. This
selection is necessarily incomplete, and the authors of this review
have different opinions on the intrinsic merits, viability and
aesthetic appeal of these theories. The main criterion we used to
choose these particular theories is that they are simple enough to
make definite (and sometimes ``orthogonal'') predictions for the
strong-field dynamics of compact objects. The theories we discuss
include:
\begin{itemize}
\item[1)] scalar-tensor theories and their generalizations (including
tensor-multiscalar and Horndeski theories);
\item[2)] $f(R)$ theories;
\item[3)] theories whose action contains terms quadratic in the
curvature, including in particular Einstein-dilaton-Gauss-Bonnet
(EdGB) and dynamical Chern-Simons (dCS) theories;
\item[4)] Lorentz-violating theories, including Einstein-\AE ther,
Ho\v rava and $n$-Dirac-Born-Infeld ($n$-DBI) gravity;
\item[5)] massive gravity theories;
\item[6)] theories involving nondynamical fields, including the
Palatini formulation of $f(R)$ gravity and Eddington-inspired
Born-Infeld (EiBI) gravity.
\end{itemize}
This broad classification will be a leitmotif of the
review. Table~\ref{tab:theories} lists some key references to the
literature on the
various theories listed above, plus others that are not considered in
depth here. The table is an incomplete (but hopefully useful) ``bird's
eye'' reference guide for further study. Similar tables following the
same classification scheme will support our discussion of the
structure and stability of compact objects.
Since we do not have a full theory of quantum gravity, an effective
field theory (EFT) approach is often invoked when constructing
phenomenological alternatives to
GR~\cite{Burgess:2003jk,Burgess:2007pt}. For example, not all theories
of gravity with action quadratic in the curvature (item 3 in the list)
are acceptable: the equations are of second order in the
strong-coupling limit (a very desirable feature, given that
higher-order derivatives are vulnerable to the so-called Ostrogradskii
instability~\cite{Woodard:2006nt}) only if the quadratic invariants
appear in the special ``Gauss-Bonnet'' combination. To avoid
higher-order derivatives in the equations of motion one must generally
assume that couplings are small, and work in an EFT framework. A more
detailed discussion of EFTs and further references can be found in
Section~\ref{sec:EFT}.
\subsection{Compact objects in modified theories of gravity}
Investigations of compact objects, binary pulsars, cosmology and
gravitational radiation vary in depth and scope for the various
classes of theories listed above. The best studied examples include
scalar-tensor theories and some forms of quadratic gravity.
Chapters~\ref{sec:BHs},~\ref{sec:NSs} and~\ref{sec:CB} are devoted to
a discussion of isolated BHs, isolated NSs and compact binary systems
in various theories.
\paragraph{Isolated black holes.}
In Chapter~\ref{sec:BHs} we discuss BHs, one of the most striking
predictions of GR. There is a consensus in the astronomy community
that the massive compact objects in galactic centers, as well as the
compact objects with mass larger than about $3M_\odot$ found in some
low-mass X-ray binaries, are well described by the Kerr solution in
GR. However, this ``BH paradigm'' rests on somewhat shaky foundations.
From a theorist's point of view, one of the most convincing arguments
in favor of the BH paradigm is that the alternatives are either
unstable (as in the case of dense star clusters, fermion stars or
naked singularities), unnatural (e.g.~``exotic'' matter violating some
of the energy conditions), contrived (such as gravastars), more
implausible than BHs as the end-point of gravitational collapse (boson
stars) or nearly indistinguishable from Kerr.
The experimental evidence that astronomical BH candidates possess
event horizons (more correctly, apparent horizons) rather than solid
surfaces usually rests on plausibility arguments based on accretion
physics~\cite{Narayan:2005ie,Narayan:2013gca}. All of these arguments
are model-dependent, and they leave room for some skepticism~(see
e.g.~\cite{Abramowicz:2002vt}).
Strictly speaking, any tests that probe the Kerr {\em metric} alone
(such as tests based on matter accretion or ray tracing of photon
trajectories) are of little value as internal tests of GR. The reason
is that a large number of extensions of GR admit the Kerr metric as a
solution, and the theories that do not (e.g.~EdGB, dCS and some
Lorentz-violating gravity theories) predict BH solutions that differ
from GR by amounts that may not be astrophysically measurable.
Despite this somewhat pessimistic caveat, many ``quasi-Kerr metrics''
have been proposed to perform GR tests, and we will review these
proposals in Chapter~\ref{sec:BHs}. Most deformations of the Kerr
metric should be viewed as unnatural strawmen: they often have serious
pathologies, and they are therefore unacceptable even for the limited
scope of parametrizing deviations from the Kerr metric~\cite{Johannsen:2013rqa}.
The prospects for testing GR with BHs look brighter when we recall
that all extensions of GR predict different {\em dynamics} and
different GW signatures when compact objects are perturbed away from
equilibrium and/or when they merge. These arguments suggest that the
most promising way to verify whether the compact objects in galactic
centers or low-mass X-ray binaries are actually Kerr BHs is via direct
observation of gravitational radiation, especially in the strong-field
merger/ringdown phase.
Last but not least, astrophysical BHs can be used to constrain
modifications of GR in a different way. Many proposed modifications of
Einstein's theory and extension of the Standard Model of particle
physics predict the existence of light bosonic degrees of
freedom. Light bosons can trigger a superradiant instability, that
extracts angular momentum from rotating BHs. By setting the
superradiant instability timescale equal to the typical timescale for
accretion to spin up the hole (say, the Salpeter time) one can get
very stringent constraints on the allowed masses of light bosons
(e.g.~axions, Proca fields or massive gravitons).
Table~\ref{tab:BHsummary} is a quick reference guide to BH solutions
and stability in various modified theories of gravity, organized in
the same way as Table~\ref{tab:theories}.
\paragraph{Isolated neutron stars.}
In Chapter~\ref{sec:NSs} we discuss NS solutions and their stability
in various extensions of GR. Among other topics, we review the
possibility that NSs in scalar-tensor theory may significantly deviate
from their GR counterparts in the presence of ``spontaneous
scalarization'' (a phase transition akin to spontaneous
magnetization~\cite{Damour:1993hw}), we discuss controversial claims on the existence of
NSs in $f(R)$ theories, and we review the somewhat surprising
``no-hair'' properties of NSs in quadratic gravity.
A major problem in carrying out strong-gravity tests with NSs is the
degeneracy between our ignorance of the equation of state (EOS) of
high-density matter and strong-gravity effects. A possibility to lift
the degeneracy consists of using universal relations between the
moment of inertia, Love number (a measure of tidal deformability) and
quadrupole moment of a NS -- the so-called ``I-Love-Q'' relations~\cite{Yagi:2013bca} -- as well as EOS-independent relations between
the lowest three multipole moments and those of higher order~\cite{Pappas:2013naa,Yagi:2014bxa}. Section~\ref{subsec:ILQ} overviews
the promises and challenges of this approach.
A property of isolated NSs that plays an important role in many
extensions of GR is their ``sensitivity.'' The sensitivity is a
measure of how the gravitational mass of the NS (or any
self-gravitating object) varies as it moves within the nonhomogeneous
extra field(s) mediating the gravitational interactions -- or in other
words, a measure of the violation of the strong equivalence principle
in the theory in question. Chapter~\ref{sec:sensitivities} is a review
of sensitivity calculations, that play an important role in binary
dynamics.
In Table~\ref{tab:NSsummary} we give a quick reference guide to NS
solutions and their stability in various modified theories of gravity.
\paragraph{Compact binaries.}
In preparation for binary pulsar tests (covered in
Chapter~\ref{sec:BP}) and GW tests (Chapter~\ref{sec:GW}), in
Chapter~\ref{sec:CB} we review calculations of compact binary dynamics
in some extensions of GR. The equations of motion and GW fluxes have
been derived using the post-Newtonian (PN) expansion -- an expansion
in powers of $v/c$, where $v$ is the orbital velocity of the binary --
in scalar-tensor theory, $f(R)$ gravity, specific forms of quadratic
gravity (including EdGB and dCS) and Lorentz-violating theories. In
comparison, numerical work is much less developed: at the moment of
writing this review, simulations of compact binary mergers were
carried out only for some of the simplest scalar-tensor theories.
\subsection{Present and future tests of strong gravity}
Chapters~\ref{sec:BP} and~\ref{sec:GW} capitalize on the material
covered in previous chapters. Chapter~\ref{sec:BP} reviews {\em
present} astrophysical tests of GR, more specifically those coming
from binary pulsar and cosmological observations. Chapter~\ref{sec:GW}
focuses on the potential payoff of {\em future} GW observations, and
on how astrophysical modeling will affect our ability to perform tests
of strong-field gravity in this context.
The first part of Chapter~\ref{sec:BP} is an overview of the
spectacular progress of GR tests from binary pulsars. These
extraordinary natural laboratories can be utilized to probe with high
precision various nonradiative strong-field effects, as well as
radiative aspects of gravity~\cite{1992Natur.355..132T}. For instance,
pulsars are now able to test Einstein’s quadrupole formula for GW
emission to an accuracy of less than $0.1\%$. They provide stringent
bounds on dipolar radiation and on violations of the strong
equivalence principle by strongly self-gravitating bodies (the best
tests coming from pulsar-white dwarf systems), and they tightly
constrain hypothetical violations of local Lorentz invariance of
gravity. The near future in this field is particularly
bright. Facilities such as the Five-hundred meter Aperture Spherical
radio Telescope (FAST) and the Square Kilometer Array (SKA) are
expected to come online soon. They should provide drastic improvements
in the precision of current tests, qualitatively new tests with
already known systems, and the discovery of many new ``pulsar
laboratories'' (possibly including the first pulsar-BH system).
The second part of Chapter~\ref{sec:BP} reviews cosmological tests of
GR. In the last few decades, a remarkable wealth of astronomical data
has constrained the expansion rate of the Universe and provided
accurate maps of large-scale structure and the cosmic microwave
background, placing ever-tightening constraints on cosmological
parameters.
In particular, anisotropies in the cosmic microwave background encode
information on the geometry of the Universe, its material constituents
and the initial conditions for structure formation. If GR is assumed
to be correct, 96\% of the material content of the Universe must
consist of dark matter and dark energy. Since the evidence for these
dark constituents of the Universe is purely gravitational, there have
been countless attempts at finding theories in which dark matter and
dark energy arise from modifications of gravity. These modifications
affect the expansion rate of the Universe, but they should also affect
gravitational clustering in a way that might be distinguishable from
GR. The proliferation of alternative theories of gravity has led to
the development of model-independent cosmological tests of modified
gravity somewhat similar to the PPN framework, which are now one of
the primary drivers for future surveys of large scale structure. In
the linear regime, these model-independent tests can be grouped in
three classes, corresponding to three manifestations of a gravity
theory: the action, the field equations derived from that action, and
the combinations of those field equations which influence observable
quantities. Chapter~\ref{sec:BP} reviews these tests as well as recent
progress in the nonlinear regime, where screening effects are
important and numerical simulations are necessary.
Last but not least, in Chapter~\ref{sec:GW} we turn our attention to
the future of strong-gravity tests, focusing on the promise of GW
observations by Earth- and space-based detectors. The main target for
both classes of detectors is the inspiral and merger of compact
binaries. A technique called matched filtering, based on a careful
monitoring of the GW phase to extract the (generally weak) signal from
the detector's noise, is used to observe these systems and to measure
their parameters. GR makes very specific and testable predictions on
the GW phasing of compact binaries as they inspiral, and on the
oscillation frequencies of the compact objects that they produce as a
result of the merger. If observed, any deviations from these
predictions may identify problems in Einstein's theory, and even point
us to specific ways in which it could be modified.
There are several comprehensive reviews on GW-based tests of GR. In
particular, the recent {\em Living Reviews in Relativity} article by
Yunes and Siemens~\cite{Yunes:2013dva} provides an excellent
introduction to the literature on GR tests with Earth-based detectors
(such as Advanced LIGO/Virgo, LIGO-India and KAGRA) and Pulsar Timing
Arrays, and the review by Gair et al.~\cite{Gair:2012nm} expounds the
great potential of future space-based detectors such as eLISA.
We find it unnecessary to reproduce that material here, and therefore
we focus on aspects that are not covered in detail in those reviews,
namely: (1) the data analysis implementation of GR tests in advanced
Earth-based detectors (the TIGER framework), arguably our best hope to
constrain modified gravity using GW observations in the near future;
and (2) an analysis of how astrophysical effects can limit (or
sometimes enhance) our ability to test strong-field gravity with GW
observations.
As a rule, in this paper we use geometrical units where the
gravitational constant and the speed of light are set to unity:
$G_N=c=1$. Factors of $G_N$ and $c$ are occasionally reinstated for
clarity, and in isolated cases (e.g. in Section~\ref{sec:EFT}) we
switch to units such that $\hbar=c=1$. We adopt the mostly positive
signature for the metric, and the same conventions as in Misner,
Thorne, and Wheeler~\cite{MTW} for the Riemann tensor.
\section{Introduction}
\label{sec:intro}
\input{intro.tex}
\input{summarytables.tex}
\clearpage
\section{Extensions of general relativity: motivation and overview}
\label{sec:alt-th}
\input{sec-alt-th/alt-th.tex}
\clearpage
\section{Black holes}
\label{sec:BHs}
\input{sec-compobj/BHs.tex}
\clearpage
\section{Neutron stars}
\label{sec:NSs}
\input{sec-compobj/NSs.tex}
\clearpage
\section{Compact binaries}
\label{sec:CB}
\input{sec-binaries/stbinaries.tex}
\input{sec-binaries/other_binaries.tex}
\clearpage
\section{Binary pulsar and cosmological tests of general relativity}
\label{sec:BP}
\input{sec-gr-tests/sec-gr-tests.tex}
\clearpage
\section{Gravitational wave tests}
\label{sec:GW}
\input{sec2_astro}
\clearpage
\section{Discussion and conclusions}
\input{conclusions}
\clearpage
\section*{Acknowledgments}
This review was conceived during a workshop funded by the
FP7-PEOPLE-2011-IRSES Grant No.~295189 ``NRHEP'' \cite{TestGR14}.
We thank Marco Cavagli\`a, Neil Cornish, Lu\'is Crispino and Nicol\'as
Yunes for attending the workshop and for useful discussions.
We are also grateful to T\'erence Delsate and Claudia de Rham for
comments, and to Alessandro Nagar, Thibault Damour, Loic Villain,
Michael Kramer and Fabian Schmidt for allowing us to use their
figures.
This work has been supported by the H2020-MSCA-RISE-2015 Grant
No.~690904 `StronGrHEP', the European Union's FP7 ERC Starting Grant
``The dynamics of black holes: testing the limits of Einstein's theory'' grant
agreement no.~DyBHo--256667, H2020 ERC Consolidator Grant ``Matter and
strong-field gravity: New frontiers in Einstein's theory'' grant agreement
no.~MaGRaTh--646597,
the FP7-PEOPLE-2011-CIG Grant No.~293412 ``CBHEO,''
the FP7-PEOPLE-2011-CIG Grant PCIG11-GA-2012-321608 ``GALFORMBHS,''
the NSF Grants No.~PHY-1055103, PHY-1260995, PHY-1306069 and PHY-1300903,
the NASA Grant NNX13AH44G,
the ERC-2011-StG Grant No.~279363--HiDGR, the FP7/2007-2013 ERC Grant
No.~306425 ``Challenging General Relativity,'' the DFG Research Training Group
1620 ``Models of Gravity'' FP7-PEOPLE-2011-IRSES Grant No.~606096,
the STFC GR Consolidator Grant No.~ST/L000636/1,
the FCT-Portugal projects PTDC/FIS/116625/2010,
CERN/FP/116341/2010, CERN/FP/123593/2011, IF/00293/2013,
IF/00797/2014/CP1214/CT0012, and CIDMA strategic funding UID/MAT/04106/2013,
the Marie Curie IEF contracts aStronGR-2011-298297 and AstroGRAphy-2013-623439,
the COST Action MP1304 ``NewCompStar,''
a UIUC Fortner Fellowship,
the S\~ao Paulo Research Foundation (FAPESP) under grants 2011/11973-4 and 2013/14754-7,
the NSF XSEDE Grant No.~PHY-090003,
the Cosmos system, part of DiRAC, funded by STFC and BIS under
Grant Nos.~ST/K00333X/1, ST/H008586/1, ST/J001341/1 and ST/J005673/1,
and
the CESGA-ICTS Grant No.~249.
Computations have been performed on
the ``Baltasar Sete-Sois'' cluster at IST,
the ``venus'' cluster at YITP,
the COSMOS supercomputer,
the Trestles cluster at SDSC,
the Kraken cluster at NICS,
and
Finis Terrae at CESGA.
T.~Baker is supported by All Souls College, Oxford.
D.~Doneva would like to thank the Alexander von Humboldt Foundation
for support.
P.~G.~Ferreira acknowledges support from STFC, BIPAC and Oxford Martin
School.
D.~Gerosa is supported by the UK STFC and the Isaac Newton Studentship
of the University of Cambridge.
J.~Kunz acknowledges support from DFG Research Training Group 1620 “Models of
Gravity” and FP7-PEOPLE-2011-IRSES Grant No.~606096.
A.~Matas would like to thank Claudia de Rham and Andrew Tolley for many
useful conversations and support. A.~Matas is supported by an NSF-GRFP
fellowship.
B.~S.~Sathyaprakash acknowledges the support of the LIGO Visitor Program
through the National Science Foundation award PHY-0757058 and STFC grant
ST/J000345/1.
L.~C.~Stein acknowledges that support for this work was provided by
the NASA through Einstein Postdoctoral Fellowship Award Number
PF2-130101 issued by the Chandra X-ray Observatory Center, which is
operated by the Smithsonian Astrophysical Observatory for and on
behalf of the National Aeronautics Space Administration under contract
NAS8-03060.
This research was supported in part by Perimeter Institute for
Theoretical Physics. Research at Perimeter Institute is supported by
the Government of Canada through Industry Canada and by the Province
of Ontario through the Ministry of Economic Development \& Innovation.
\clearpage
\subsection{Science opportunities}\label{sec:TestConcepts}
\input{sec-astro/intro.tex}
\subsection{Parameter estimation and model selection}\label{sec:paramestimation}
\input{sec-astro/SS_detector}
\subsection{Direct versus parametrized tests of gravity}\label{sec:paramtests}
\input{sec-astro/ppE}
\subsubsection{Implementation of direct tests: the TIGER pipeline}
\label{sec:tiger}
\input{sec-astro/SS_source}
\subsection{Waveform and astrophysical systematics}
\label{sec:systematics}
\input{sec-astro/waveforms.tex}
\subsubsection{Supermassive black holes}
\label{sec:sub:supermassive}
\input{sec-astro/environment}
\subsection{A compass to navigate the modified-gravity atlas}
\label{subsec:review-modgrav}
\input{sec-alt-th/mg-review}
\subsection{Scalar-tensor gravity}
\label{subsec:ST}
\input{sec-alt-th/scalar-tensor}
\subsection{Metric $f(R)$ theories}
\label{subsec:f(R)}
\input{sec-alt-th/fofR}
\subsection{Quadratic gravity}
\label{subsec:quadratic}
\input{sec-alt-th/quadratic}
\subsection{Lorentz-violating theories}
\label{subsec:lorentz-viol}
\input{sec-alt-th/lorentz-viol}
\subsection{Massive gravity and Galileons}
\label{subsec:massgrav}
\input{sec-alt-th/massive-and-galileons}
\subsection{Gravity with auxiliary fields}
\label{subsec:auxiliary}
\input{sec-alt-th/auxiliary}
\subsection{General relativity and quantum mechanics: an effective field theory approach}
\label{sec:EFT}
\input{sec-alt-th/EFT_app.tex}
\subsection{Open problems}
\label{op:theories}
\input{sec-alt-th/openproblems}
\subsubsection{The Bergmann-Wagoner formulation.}
The most general action of scalar-tensor gravity with one scalar field
which is at most quadratic in derivatives of the fields was studied by
Bergmann and Wagoner~\cite{Bergmann:1968ve,Wagoner:1970vr}, and can be
written (after an appropriate field redefinition) as:
\begin{equation}
S=\frac{1}{16\pi}\int d^4x\sqrt{-g}\left[\phi R-\frac{\omega(\phi)}{\phi}g^{\mu\nu}
\left(\partial_\mu\phi\right)\left(\partial_\nu\phi\right)
-U(\phi)\right]+S_M[\Psi,g_{\mu\nu}]\,,\label{STactionJ}
\end{equation}
where $\omega$ and $U$ are arbitrary functions of the scalar field
$\phi$, and $S_M$ is the action of the matter fields $\Psi$. When
$\omega(\phi)=\omega_{BD}$ is constant and $U(\phi)=0$, the theory
reduces to (Jordan-Fierz-)Brans-Dicke
gravity~\cite{Jordan:1959eg,Fierz:1956zz,Brans:1961sx}, an extension
of GR which was proposed in the mid-$20^{th}$ century
(see~\cite{Goenner:2012cq,Brans:2008zz,Brans:2005ra} for a historical
account).
The Bergmann-Wagoner theory \eqref{STactionJ} can be expressed in a
different form through a scalar field redefinition
$\varphi=\varphi(\phi)$ and a conformal transformation of the metric
$g_{\mu\nu}\rightarrow g^\star_{\mu\nu}=A^{-2}(\varphi)
g_{\mu\nu}$. In particular, fixing $A(\varphi)=\phi^{-1/2}$, the
action~(\ref{STactionJ}) -- generally referred to as the {\it Jordan-frame}
action -- transforms into the {\it Einstein-frame} action
\begin{equation}
S=\frac{1}{16\pi}\int d^4x\sqrt{-g^\star}\left[R^\star-2g^{\star\mu\nu}
\left(\partial_\mu\varphi\right)\left(\partial_\nu\varphi\right)
-V(\varphi)\right]+S_M[\Psi,A^2(\varphi)g^\star_{\mu\nu}]\,,\label{STactionE}
\end{equation}
where $g^\star$ and $R^\star$ are the determinant and Ricci scalar of
$g_{\mu\nu}^{\star}$, respectively, and the potential
$V(\varphi) \equiv A^4(\varphi)U(\phi(\varphi))$.
The price paid for the minimal
coupling of the scalar field in the gravitational sector is the
nonminimal coupling in the matter sector of the action: particle
masses and fundamental constants depend on the scalar field.
We remark that the actions~\eqref{STactionJ} and \eqref{STactionE} are
just different representations of the same theory: the outcome of an
experiment will not depend on the chosen representation, as long as
one takes into account that the units of physical quantities do scale
with powers of the conformal factor
$A$~\cite{Flanagan:2004bz,Sotiriou:2007zu}. It is then legitimate,
when
modeling a physical process, to choose the conformal frame in which
calculations are simpler: for instance, in vacuum the Einstein-frame
action \eqref{STactionE} formally reduces to the GR action minimally
coupled with a scalar field. It may then be necessary to change the
conformal frame when extracting physically meaningful statements
(since the scalar field is minimally coupled to matter in the Jordan
frame, test particles follow geodesics of the {\em Jordan-frame}
metric, not of the Einstein-frame metric).
The relation between Jordan-frame and Einstein-frame quantites is
simply $\phi=A^{-2}(\varphi)$, $3+2\omega(\phi)=\alpha(\varphi)^{-2}$,
where $\alpha(\varphi)\equiv d(\ln
A(\varphi))/d\varphi$~\cite{Will:2014xja}. Note that the theory is
fixed once the function $\omega(\phi)$ -- or, equivalently,
$\alpha(\varphi)$ -- is fixed, and the form of the scalar potential is
chosen. Moreover, many phenomenological studies neglect the scalar
potential. This approximation corresponds to neglecting the
cosmological term, the mass of the scalar field and any possible
scalar self-interaction. In an asymptotically flat spacetime the
scalar field tends to a constant $\phi_0$ at spatial infinity,
corresponding to a minimum of the potential. Taylor expanding
$U(\phi)$ around $\phi_0$ yields a cosmological constant and a mass
term for the scalar field to the lowest
orders~\cite{Wagoner:1970vr,Alsing:2011er}.
Scalar-tensor theory with a vanishing scalar potential is
characterized by a single function $\alpha(\varphi)$. The expansion of
this function around the asymptotic value $\varphi_0$ can be written
in the form
\begin{equation}
\alpha(\varphi)=\alpha_0+\beta_0(\varphi-\varphi_0)+\dots\label{DEalphabeta}
\end{equation}
As mentioned above, the choice $\alpha(\varphi)=\alpha_0=$constant
(i.e., $\omega(\phi)=$constant) corresponds to Brans-Dicke theory. A
more general formulation, proposed by Damour and Esposito-Far\`ese, is
parametrized by $\alpha_0$ and $\beta_0$~\cite{Damour:1993hw,Damour:1996ke}. Another simple variant is massive
Brans-Dicke theory, in which $\alpha(\varphi)$ is constant, but the
potential is nonvanishing and has the form
$U(\phi)=\frac{1}{2}U''(\phi_0)(\phi-\phi_0)^2$, so that the scalar field
has a mass $m_s^2\sim U''(\phi_0)$. Note that since the scalar field
$\varphi$ in the action (\ref{STactionE}) is dimensionless, the
function $\alpha(\varphi)$ and the constants $\alpha_0$, $\beta_0$ are
dimensionless as well.
The field equations of scalar-tensor theory in the Jordan frame are
(see e.g.~\cite{Eardley:1975,will1993theory})
\begin{subequations}
\label{ST:Jordan}
\begin{align}
G_{\mu \nu} &= \frac{8\pi}{\phi}T_{\mu \nu}+\frac{\omega(\phi)}{\phi^2}\left(\partial_\mu\phi\partial_\nu\phi
-\frac{1}{2}g_{\mu\nu}\partial_\lambda\phi\partial^\lambda\phi\right)+\frac{1}{\phi}(\nabla_\mu\nabla_\nu\phi-g_{\mu\nu}\Box_g \phi)
-\frac{U(\phi)}{2\phi}g_{\mu\nu}\, ,
\label{eq:tensoreqnJ}\\
\Box_g \phi &= \frac{1}{3+2\omega(\phi)}\left(8\pi T - 16\pi\phi\frac{\partial T}{\partial \phi}
-\frac{d\omega}{d\phi}\partial_\lambda\phi\partial^\lambda\phi+\phi\frac{dU}{d\phi}-2U(\phi)\right) \, ,
\label{eq:scalareqnJ}
\end{align}
\end{subequations}
where $T^{\mu\nu}=-2(-g)^{-1/2}\delta S_M(\Psi,g_{\mu\nu})/\delta
g_{\mu\nu}$ is the Jordan-frame stress-energy tensor of matter
fields, and $T=g^{\mu\nu}T_{\mu\nu}$.
In the Einstein frame, the field equations are
\begin{subequations}
\label{ST:Einstein}
\begin{align}
G^\star_{\mu \nu} &=2\left(\partial_\mu\varphi\partial_\nu\varphi-
\frac{1}{2}g^\star_{\mu\nu}\partial_\sigma\varphi\partial^\sigma\varphi\right)-
\frac{1}{2}g^\star_{\mu\nu}V(\varphi)+8\pi T^\star_{\mu\nu}\,,\label{eq:tensoreqnE} \\
\Box_{g^\star} \varphi &=-4\pi\alpha(\varphi)T^\star+\frac{1}{4}\frac{dV}{d\varphi}\,,\label{eq:scalareqnE}
\end{align}
\end{subequations}
where $T^{\star\,\mu\nu}=-2(-g)^{-1/2} \delta
S_M(\Psi,A^2g^\star_{\mu\nu})/\delta g^\star_{\mu\nu}$ is the Einstein-frame
stress-energy tensor of matter fields and
$T^\star=g^{\star\,\mu\nu}T^\star_{\mu\nu}$ (see
e.g.~\cite{Damour:1992we}). Eq.~\eqref{eq:scalareqnE} shows that
$\alpha(\varphi)$ couples the scalar fields to
matter~\cite{Damour:1995kt}, as does $(3+2\omega(\phi))^{-1}$ in the
Jordan frame: cf.~Eq.~\eqref{eq:scalareqnJ}].
Astrophysical observations set bounds on the parameter space of
scalar-tensor theories. In the case of Brans-Dicke theory, the best
observational bound ($\alpha_0<3.5\times 10^{-3}$) comes from the
Cassini measurement of the Shapiro time delay. In the more general
case with $\beta_0\neq0$, current constraints on $(\alpha_0,\beta_0)$
have been obtained by observations of NS-NS and NS-WD binary
systems~\cite{Freire:2012mg}, and will be discussed in
Section~\ref{sec:BP} (cf.~Figure~\ref{fig:stg}). Observations of compact
binary systems also constrain massive Brans-Dicke theory, leading to
exclusion regions in the $(\alpha_0,m_s)$ plane~\cite{Alsing:2011er}.
An interesting feature of scalar-tensor gravity is the prediction of
certain characteristic physical phenomena which do not occur at all in
GR. Even though we know from observations that $\alpha_0\ll1$ and that
GR deviations are generally small, these phenomena may lead to
observable consequences. There are at least three possible smoking
guns of scalar-tensor gravity. The first is the emission of dipolar
gravitational radiation from compact binary
systems~\cite{Eardley:1975,Will:1989sk}, which will be discussed in
Section~\ref{sec:binaries/scal-tens-theor}. Dipolar gravitational
radiation is ``pre-Newtonian,'' i.e.~it occurs at lower PN order than
quadrupole radiation, and it does not exist in GR. The second is the
existence of nonperturbative NS solutions in which the scalar field
amplitude is finite even for $\alpha_0\ll1$. This {\it spontaneous
scalarization} phenomenon~\cite{Damour:1993hw,Damour:1996ke} will be
discussed in detail in Section~\ref{sec:NS_ST}. Here we only remark
that spontaneous scalarization would significantly affect the mass and
radius of a NS, and therefore the orbital motion of a compact binary
system, even far from coalescence. The third example is also
nonperturbative, and it involves massive fields. The coupling of
massive scalar fields to matter in orbit around rotating BHs leads to
a surprising effect: because of superradiance, matter can hover into
``floating orbits'' for which the net gravitational energy loss at
infinity is entirely provided by the BH's rotational
energy~\cite{Cardoso:2011xi}.
The phenomenology of scalar-tensor theory in vacuum spacetimes, such
as BH spacetimes, is less interesting. When the matter action $S_M$
can be neglected, the Einstein-frame formulation of the theory is
equivalent to GR minimally coupled to a scalar field. BHs in
Bergmann-Wagoner theories satisfy the same {\it no-hair theorem} as in
GR, and thus the stationary BH solutions in the two theories
coincide~\cite{Heusler:1995qj,Sotiriou:2011dz}. Moreover, dynamical
(vacuum) BH spacetimes satisfy a similar {\it generalized no-hair
theorem}: the dynamics of a BH binary system in Bergmann-Wagoner
theory with vanishing potential are the same as in
GR~\cite{Damour:1992we}, up to at least $2.5$ PN order for generic
mass ratios~\cite{Mirshekari:2013vb} and at any PN order in the
extreme mass-ratio limit~\cite{Yunes:2011aa} (see
Section~\ref{sec:pn_st}). These no-hair theorems will be discussed in
Section~\ref{sec:ST_BHs}.
\subsubsection{Scalar-tensor theories with multiple scalar fields.} \label{sec:multiscalars}
When gravity is coupled with more than one scalar field, the
action~\eqref{STactionJ} has the more general form~\cite{Damour:1992we}
\begin{align}
S\,=\,\,& \frac1{16\pi}\int d^4x\sqrt{-g}\left(
F(\phi)R-\gamma_{ab}(\phi)g^{\mu\nu}\partial_\mu\phi^a
\partial_\nu\phi^b-V(\phi)\right)+S_M[\Psi,\,g_{\mu\nu}]\,,
\label{STactionJ_multi}
\end{align}
where $F,V$ are functions of the $N$ scalar fields $\phi^a$
($a=1\ldots N$). The scalar fields live on a manifold (the {\it target
space}) with metric $\gamma_{ab}(\phi)$. The
action~\eqref{STactionJ_multi} is invariant not only under space-time
diffeomorphisms, but also under target-space diffeomorphisms,
i.e.~scalar field redefinitions. These theories have a richer
structure
than those with a single scalar field, since the geometry of the
target space can affect the dynamics. For instance, the theories with
a complex scalar field discussed in Section~\ref{sec:complscal}, in
which the no-hair theorems can be circumvented, can also be seen as
multiscalar-tensor theories with $N=2$.
\subsubsection{Horndeski gravity.}\label{subsec:horndeski}
The most general scalar-tensor theory with second-order field
equations (and one scalar field) is Horndeski gravity~\cite{Horndeski:1974wa}. The action of Horndeski gravity can be
written in terms of Galileon interactions (see~\cite{Deffayet:2011gz}
and Section~\ref{subsec:massgrav}) as
\begin{equation}
\begin{aligned}
S={}&\int d^4x\sqrt{-g}\Big\{K(\phi,X)-G_3(\phi,X)\Box\phi\\
&{}+G_4(\phi,X)R+G_{4,X}(\phi,X)\left[(\Box\phi)^2
-(\nabla_\mu\nabla_\nu\phi)(\nabla^\mu\nabla^\nu\phi)\right]\\
&{}+G_5(\phi,X)G_{\mu\nu}\nabla^\mu\nabla^\nu\phi
-\frac{G_{5,X}(\phi,X)}{6}\big[
(\Box\phi)^3-3\Box\phi(\nabla_\mu\nabla_\nu\phi)(\nabla^\mu\nabla^\nu\phi)
\\
&{}+2(\nabla_\mu\nabla_\nu\phi)(\nabla^\mu\nabla_\sigma\phi)
(\nabla^\nu\nabla^\sigma\phi)\big]\Big\}\,,
\end{aligned}
\label{action_horndeski}
\end{equation}
where $K$ and the $G_i$'s ($i=1\dots 5$) are functions of the scalar
field $\phi$ and of its kinetic term
$X=-1/2\partial^\mu\phi\partial_\mu\phi$, and $G_{i,X}$ are
derivatives of $G_i$ with respect to the kinetic term $X$. For a
particular choice of these functions, this theory coincides with
Gauss-Bonnet gravity (see Section~\ref{subsec:quadratic}).
As we shall discuss in Section~\ref{sec:ST_BHs}, in Horndeski theory
the no-hair theorem can be circumvented, and thus stationary BH
solutions can be different from GR.
\subsubsection{Einstein-dilaton-Gauss-Bonnet gravity}\label{subsec:EdGB}
When $f_2=-4f_1=-4f_{3}$ and $f_4=0$, the
theory~\eqref{eq:action_quadratic} reduces to EdGB
gravity~\cite{Kanti:1995vq}, with action
\begin{equation}
S=\frac{1}{16\pi}\int\sqrt{-g} d^4x \left[R-2\nabla_\mu\phi\nabla^\mu\phi-
V(\phi)+f_1(\phi) R_{\rm GB}^2\right]\,, \label{EdGBaction}
\end{equation}
where $f_1(\phi)$ is a generic coupling function and the Gauss-Bonnet invariant $R_{\rm GB}^2$
has been defined below Eq.~(\ref{curvatureinvariants}). This is the only
quadratic theory of gravity whose field equations are of second
differential order for {\em any} coupling, and not just in the
weak-coupling limit. Indeed, when $f_1(\phi)=\alpha_{\rm GB}
e^{-2\phi}$, the theory reduces to the bosonic sector of heterotic
string theory~\cite{Gross:1986mw}.
Gauss-Bonnet gravity can also be seen as a particular case of
Horndeski gravity~\cite{Kobayashi:2011nu}, as mentioned in
Section~\ref{subsec:horndeski}. For instance, in the case
$f_1(\phi)=\alpha\phi$, the action~\eqref{EdGBaction} can be shown to
be equivalent to the action~\eqref{action_horndeski} with $K=X/2$,
$G_3=0$, $G_4=1/2$, $G_5=-2\alpha\ln|X|$~\cite{Sotiriou:2013qea}.
As in all of these theories, the coupling parameter is
\emph{dimensionful} and, specifically, it has dimensions of an inverse
curvature.
It is thus natural to expect that the strongest constraints
on the theory should come from physical systems involving high
curvature: BHs, NSs and the early Universe.
We postpone a discussion
of BHs and NSs to Sections~\ref{sec:BHs} and~\ref{sec:NSs},
respectively. Here we anticipate the observational bounds that have
been derived.
Most bounds have been derived in the weak-coupling approximation, where one expects
\begin{equation}
\sqrt{|\alpha_{\rm GB}|}\lesssim {\cal O}(L)\,,
\end{equation}
where $L$ is the typical curvature radius in the system under
consideration. Thus, Solar System constraints---such as those derived
by measuring the Shapiro time delay of the Cassini
probe~\cite{Bertotti:2003rm}---give a mild bound $\sqrt{|\alpha_{\rm
GB}|}\lesssim10^{13}{\rm cm}$, which is in fact of the order of an
astronomical unit. On the other hand, as we shall discuss in
Section~\ref{sec:BHs}, BHs in this theory carry a scalar charge, and
observations of BH low-mass X-ray binaries give a constraint which is
six orders of magnitude stronger~\cite{Yagi:2012gp}:
\begin{equation}
\sqrt{|\alpha_{\rm GB}|}\lesssim 5\times 10^6{\rm cm} \label{constraint_EdGB}
\end{equation}
(in the units of Eq.~\eqref{EdGBaction}). As expected, this
constraint is comparable to the typical size of a stellar-mass BH. On
the other hand, the only bound on EdGB gravity as an \emph{exact}
theory is of theoretical nature, because the existence of BH solutions
implies that $\sqrt{|\alpha_{\rm GB}|}$ be smaller than the BH horizon
size~\cite{Kanti:1995vq}; this bound implies
$\alpha_{\rm GB}/M^2\lesssim0.691$~\cite{Pani:2009wy}. Thus, the
observational constraint~\eqref{constraint_EdGB} is likely to be a
good estimate also for the exact EdGB gravity.
As previously mentioned, the bounds listed above are clearly satisfied
if one assumes that quadratic curvature corrections become relevant
only at the Planck scale. Nonetheless, they represent the best
constraints on quadratic gravity to date, and they were obtained
without any a priori assumptions on the regime in which deviations
from GR should be relevant.
\subsubsection{Chern-Simons gravity}
\label{sec:ChernSimons}
While the terms proportional to $f_1$, $f_2$ and $f_3$ in the
action~\eqref{eq:action_quadratic} are all associated with
qualitatively similar corrections to GR, the term proportional to
$f_4$ is peculiarly different, to the extent that the special case
$f_1=f_2=f_3=0$ describes a specific theory (Chern-Simons gravity)
which has been widely scrutinized in recent years
(see~\cite{Alexander:2009tp} for a review). At variance with EdGB
gravity, to avoid higher-order derivatives in the field equations
Chern-Simons theory must be considered as an EFT. Almost all work so
far has focused on the special case $f_4=\alpha_{\rm CS}\phi$, working
perturbatively in the coupling constant $\alpha_{\rm CS}$. Then the
action reads
\begin{equation}
S=\frac{1}{16\pi}\int\sqrt{-g} d^4x \left[R-2\nabla_\mu\phi\nabla^\mu\phi-V(\phi) +\alpha_{\rm CS}\, \phi\, {}^{*}\!RR\right]\,, \label{CSaction}
\end{equation}
and most of the literature considered the case of a vanishing scalar
potential: $V(\phi)=0$. Like the Gauss-Bonnet term, the
Chern-Simons term ${{}^{*}\!RR}$ is also a topological invariant, so that
if $f_4={\rm const}$ the theory is equivalent to GR.
For historical reasons, Chern-Simons gravity comes in two flavors: (i)
a nondynamical version in which the scalar kinetic term
in~\eqref{CSaction} is absent, and (ii) a theory where the scalar is a
true dynamical degree of freedom, that goes under the name of
dynamical Chern-Simons (dCS) gravity. These two theories are actually
very different from each other. Despite some confusion in the
literature, only the nondynamical theory is parity breaking, whereas
dCS gravity simply has different solutions than GR for spacetimes
which are not reflection-invariant, as in the case of spinning
objects. Furthermore, the nondynamical version introduces a
constraint, ${{}^{*}\!RR}=0$, arising from the variation of the CS
action with respect to the nondynamical scalar
field~\cite{Grumiller:2007rv}. This constraint limits the space of
solutions of the modified gravitational equations and introduces other
problems~\cite{Alexander:2009tp}. For these reasons, the dynamical
version of the theory has received much more attention in recent
years.
It can be shown that any spherically symmetric solution of GR is also
a solution of dCS gravity~\cite{Yunes:2007ss}, and this makes it
challenging to distinguish between the two theories. On the other
hand, dCS gravity is almost unique as an extension of GR, as it
predicts corrections only in the presence of a parity-odd source such
as rotation. Among the most studied predictions of the theory are an
amplitude birefringence in GW propagation~\cite{Alexander:2009tp} and
modified spinning solutions, including corrections to Kerr BHs and
rotating NSs, to be discussed in detail in Sections~\ref{sec:BHs}
and~\ref{sec:NSs}.
To lowest order in the rotation rate, the CS modification to GR only
affects the gravitomagnetic sector of the metric. Tests of the theory
might therefore rely on frame-dragging effects.
Using the results of Gravity Probe B~\cite{Everitt:2011hp},
Ref.~\cite{AliHaimoud:2011fw} derived the bound
\begin{equation}
\sqrt{|\alpha_{\rm CS}|}<{\cal O}(10^{13}){\rm cm}\,.\label{csobsb}
\end{equation}
As mentioned above, dCS gravity should be interpreted as an EFT, and
to have perturbative control requires $\alpha_{\rm CS}/M^2 \ll
1$. This requirement is stronger than the bound~\eqref{csobsb} for
BHs with masses $M\lesssim 10^8\,M_\odot$.
Similar bounds come from the Lense-Thirring effect as measured by the
LAGEOS satellites, which have also been used to constrain the
nondynamical version of the theory~\cite{Alexander:2009tp}. Note that
these bounds are of the order of an astronomical unit, as expected
from the previous dimensional analysis. The detection of GWs from an
extreme mass-ratio inspiral (EMRI) can potentially yield constraints
of the order $\sqrt{|\alpha_{\rm CS}|}<{\cal O}(10^{10}){\rm cm}$ or
even determine the Chern-Simons parameter with fractional errors below
5\%~\cite{Canizares:2012ji}.
Since large-curvature environments are expected to put
stronger bounds on the theory, the optimal systems to constrain
quadratic gravity are compact binaries. Indeed,
Ref.~\cite{Yagi:2012vf} derived projected bounds that are six orders
of magnitude more stringent than the one above by considering future
GW detection of the late inspiral of BH binaries. Similar bounds were
also recently estimated in Ref.~\cite{Stein:2014xba} by analyzing CS
corrections of rapidly-spinning Kerr BHs. Such corrections could be
constrained from electromagnetic observations of accreting stellar
mass BHs such as those in low-mass X-ray binaries, e.g.~GRO J1655$-$40
(cf.~e.g.~\cite{Shafee:2005ef}).
\subsubsection{Einstein-\AE ther}
To break boost invariance in the most generic way, one can describe
the gravitational degrees of freedom by means of a metric and a
timelike vector field, $\boldsymbol{u}$, usually referred to as the
``\ae ther.'' Up to total divergences, the most general action composed of
the metric and two or fewer derivatives of the \ae ther, and that
couples the \ae ther minimally to matter (so as to enforce the weak
equivalence principle and experimental evidence against the existence
of ``fifth forces'') is given by the Einstein-\AE ther
action~\cite{Jacobson:2000xp,Eling:2004dk,Jacobson:2008aj}
\begin{equation}
\label{eq:S-AE}
S_{\ifmmode\text{\AE}\else\AE\fi} = \frac{1}{16\pi G_{\ifmmode\text{\AE}\else\AE\fi}}\int \sqrt{-g}~ (R -M^{\alpha\beta}{}_{\mu\nu} \nabla_\alpha u^\mu \nabla_\beta u^\nu)~d^{4}x
+S_{{\rm mat}}[\Psi,\,g_{\mu\nu}]
\,,
\end{equation}
where
\begin{equation}
M^{\alpha\beta}{}_{\mu\nu} = c_1 g^{\alpha\beta}g_{\mu\nu}+c_2\delta^{\alpha}_{\mu}\delta^{\beta}_{\nu}
+c_3 \delta^{\alpha}_{\nu}\delta^{\beta}_{\mu}-c_4 u^\alpha u^\beta g_{\mu\nu}\,,
\end{equation}
$c_i$ ($i=1,\dots,4$) are dimensionless couplings, and $\Psi$ denotes
the matter degrees of freedom. In this Section we do not assume
$G_N=1$; the ``bare'' $G_{\ifmmode\text{\AE}\else\AE\fi}$ is related to the ``Newtonian''
gravitational constant $G_N$ measured locally by Cavendish-type
experiments via~\cite{Carroll:2004ai}
\begin{equation}
\label{eq:GN}
G_N=\frac{2 G_{\ifmmode\text{\AE}\else\AE\fi}}{2-(c_1+c_4)}\,.
\end{equation}
To enforce the timelike character of the \ae ther, one has to impose
\begin{equation}
\label{unit}
g_{\mu\nu}u^\mu u^\nu=-1\,,
\end{equation}
either implicitly or by adding a Lagrange multiplier $\ell(g_{\mu\nu}u^\mu u^\nu+1)$
in the variation of the action
above.
The field equations for Einstein-\AE ther theory can be derived by
varying the action \eqref{eq:S-AE} with respect to $g^{\alpha\beta}$ and
$u^\mu$, while imposing the constraint \eqref{unit}.
This results
in the following modified Einstein equations:
\begin{equation}\label{E_def}
E_{\alpha\beta}\equiv G_{\alpha\beta} - T^{\ifmmode\text{\AE}\else\AE\fi}_{\alpha\beta}-8\pi G_{\ifmmode\text{\AE}\else\AE\fi} T^{\rm mat}_{\alpha\beta}=0\,,
\end{equation}
where $ T^{\rm mat}_{\alpha\beta}$ is the matter stress-energy tensor,
\begin{align}\label{Tae}
T^{\ifmmode\text{\AE}\else\AE\fi}_{\alpha\beta}&=-\nabla_\mu\left(J^{\phantom{(\alpha}\mu}_{(\alpha}u_{\beta)}-J^\mu_{\phantom{\mu}(\alpha}u_{\beta)}-J_{(\alpha\beta)}u^\mu\right)\nonumber
-c_1\,\left[
(\nabla_\mu u_\alpha)(\nabla^\mu u_\beta)-(\nabla_\alpha u_\mu)(\nabla_\beta u^\mu)
\right]\nonumber\\
&+\left[ u_\nu(\nabla_\mu J^{\mu\nu}){+}c_4 \dot{u}^2
\right] u_\alpha u_\beta +c_4 \dot{u}_\alpha \dot{u}_\beta
-\frac{1}{2} M^{\sigma\rho}{}_{\mu\nu} \nabla_\sigma u^\mu \nabla_\rho u^\nu
g_{\alpha\beta}\,,
\end{align}
$J^\alpha{}_{\mu}=M^{\alpha\beta}{}_{\mu\nu}\nabla_\beta u^\nu$,
and $\dot{u}^{\alpha}\equiv u^{\beta}\nabla_{\beta}u^{\alpha}$.
These are completed by the \ae ther equations
\begin{equation}\label{AE_def}
\ifmmode\text{\AE}\else\AE\fi_\mu\equiv\left(\nabla_\alpha J^{\alpha\nu}+c_4\dot{u}_\alpha\nabla^\nu u^\alpha\right) \left(g_{\mu\nu}+u_\mu u_\nu\right)=0\,.
\end{equation}
Strong constraints on the coupling constants $c_i$ come from Solar
System tests. This is because Einstein-\AE ther theory predicts that
the (dimensionless) preferred-frame parameters $\alpha_1$ and
$\alpha_2$ of the PPN expansion will in general be nonzero functions
of the $c_i$'s~\cite{Foster:2005dk,Jacobson:2008aj}. Because Solar
System experiments constrain $|\alpha_1|\lesssim 10^{-4}$ and
$|\alpha_2|\lesssim 10^{-7}$~\cite{Will:2014xja}, one can expand the
theory in $\alpha_1$ and $\alpha_2$, and reduce the parameter space to
just two independent couplings $c_\pm\equiv c_1\pm c_3$. The
remaining couplings are given by $c_2=(-2 c_1^2-c_1 c_3+c_3^2)/(3
c_1)+{\cal O}(\alpha_1,\alpha_2)$, $c_4=-c_3^2/c_1+{\cal
O}(\alpha_1,\alpha_2)$~\cite{Foster:2005dk,Jacobson:2008aj}.
Further constraints on the two independent couplings $c_\pm$ come from
requiring that the theory should have positive energy (i.e.~no ghosts)
and that Minkowski space should be linearly stable (i.e.~no gradient
instabilities)~\cite{Will:2014xja}.
Einstein-\AE ther theory predicts the existence of not only spin-2
gravitational perturbations (like in GR), but also spin-1 and spin-0
gravitational perturbations. All these propagating modes have speeds
that are functions of the $c_i$, and which differ in general from the
speed of light~\cite{Jacobson:2004ts}. However, if these modes were
propagating at speeds lower than the speed of light, photons (or
relativistic particles) could Cherenkov radiate into the gravitational
field and lose energy to these modes, and this would lead to
(unobserved) experimental consequences~\cite{Elliott:2005va}.
Therefore, one has to impose that the speed of the spin-2, spin-1 and
spin-0 gravitons is larger than (or equal to) the speed of light.
Taking into account these constraints, one obtains the viable region
plotted in cyan in Figure~\ref{fig:LVconstraints} (left panel) for the
two independent couplings $c_\pm$. As previewed in that figure and
discussed in Section~\ref{sec:pulsar_LV}, more stringent constraints on
$c_\pm$ come from binary pulsar data~\cite{Yagi:2013qpa,Yagi:2013ava}.
\subsubsection{Khronometric theory}
If we impose that the \ae ther is always hypersurface-orthogonal, one can express it as
\begin{equation}
\label{ho}
u_\alpha=-\frac{\partial_\alpha T}{\sqrt{-g^{\mu\nu}\partial_\mu T \partial_\nu T}}\,,
\end{equation}
where $T$ is the hypersurface-defining scalar and the constraint
\eqref{unit} has already been enforced.
By assumption, surfaces of constant $T$ foliate the spacetime,
and one can re-express the action \eqref{eq:S-AE}
adapted to this ``preferred time'' $T$. This yields a different theory, described by the ``khronometric
theory'' action~\cite{Horava:2009uw,Blas:2009qj,Jacobson:2010mx}
\begin{multline}\label{action-K}
S_{K}=\frac{1-\beta}{16\pi G_{\ifmmode\text{\AE}\else\AE\fi}}\!\int dT d^3x \, N\sqrt{h} \, \left(K_{ij}K^{ij} - \frac{1+\lambda}{1-\beta} K^2\right.
\\+\left. \frac{1}{1-\beta}{}^{(3)}\!R + \frac{\alpha}{1-\beta}\, a_ia^i\right)+S_{{\rm mat}}[\Psi,\,g_{\mu\nu}]\,,
\end{multline}
where $N=(-g^{TT})^{-1/2}$ is the lapse function, $K^{ij}$ is the extrinsic curvature of $T=$ constant hypersurfaces,
$h_{ij}$ is the induced spatial metric on those hypersurfaces,
${}^{(3)}\!R$ their three-dimensional Ricci curvature,
$a_i=\partial_i\ln{N}$, and the \ae ther is now related to the lapse via
$u_\alpha=-N\delta_{\alpha}^{T} \,.$
We have also redefined the theory's parameters via
\begin{equation}\label{eq:EAtoKH}
\lambda\equiv c_2, \quad \beta\equiv c_3+c_1, \quad \alpha\equiv c_4+c_1.
\end{equation}
It should be stressed at this stage that the action \eqref{action-K} only depends on three couplings, as opposed to four
in the action~\eqref{eq:S-AE}. This is because the hypersurface-orthogonality constraint~\eqref{ho} makes it possible to
re-express one of those four couplings in terms of the remaining three without loss of generality.
The field equations of khronometric theory are obtained by replacing the hypersurface orthogonality constraint~\eref{ho}
in the action~\eqref{eq:S-AE}, and then varying the action with respect to $g_{\alpha\beta}$ and $T$; they are
\begin{align}
\label{hl1}
E_{\alpha\beta}+2\ifmmode\text{\AE}\else\AE\fi_{(\alpha}u_{\beta)}&=0\,,\\
\label{hleq}
\nabla_\mu \left(\frac{\ifmmode\text{\AE}\else\AE\fi^\mu}{\sqrt{-\nabla^\alpha T \nabla_\alpha T}} \right)&=0\,.
\end{align}
Note that Eq.~\eref{hleq} actually follows from Eq.~\eref{hl1} and from the conservation of the matter stress-energy
tensor, i.e.~the only independent equations are the modified Einstein equations and the equations of motion of
matter~\cite{Jacobson:2010mx}. By comparing this set of equations with
the Einstein-\ifmmode\text{\AE}\else\AE\fi ther equations~\eqref{E_def} and~\eqref{AE_def}, it
is easy to see that the hypersurface-orthogonal solutions of
Einstein-\AE ther theory will also be solutions of khronometric
theory. The converse is true in spherical
symmetry~\cite{Jacobson:2010mx,Blas:2011ni,Blas:2010hb,Barausse:2012ny},
but not in more general situations. For instance, slowly rotating BH
solutions are different in the two
theories~\cite{Barausse:2012ny,Barausse:2012qh,Barausse:2013nwa}.
As in the Einstein-\AE ther case, in Khronometric theory the PPN
preferred-frame parameters $\alpha_1$ and $\alpha_2$ are nonzero and
functions of the couplings. In light of the bounds $|\alpha_1|\lesssim
10^{-4}$ and $|\alpha_2|\lesssim 10^{-7}$~\cite{Will:2014xja}, one can
expand khronometric theory in $\alpha_1$ and $\alpha_2$. As a result,
one is left with two independent parameters (say $\beta$ and
$\lambda$), while the third parameter $\alpha$ is related to the first
two by $\alpha=2\beta+{\cal O}(\alpha_1,\alpha_2)$.\footnote{%
Though it may seem that the conditions $\alpha_1=\alpha_2=0$ would
reduce the dimensionality of the parameter space to a
one-dimensional subspace, both $\alpha_1$ and $\alpha_2$ happen to
vanish for $\alpha=2\beta$ in
Khronometric theory. Thus, the conditions $\alpha_1=\alpha_2=0$ still select
a two-dimensional
subspace.
However, this only holds at the origin in $\alpha$-space,
and so saturating the bounds to $|\alpha_1|\approx10^{-4}$ and
$|\alpha_2|\approx10^{-7}$ reduces to a one-dimensional subspace. We
refer the reader to~\cite{Yagi:2013ava} for a detailed discussion.}
From the hypersurface orthogonality constraint~\eqref{ho}, there are
no propagating spin-1 gravitational modes. Requiring positive
energies, linear stability of Minkowski space, and the absence of
gravitational Cherenkov radiation for the remaining spin-0 and spin-2
degrees of freedom still selects a sizeable region of the parameter
space
$(\lambda,\beta)$~\cite{Blas:2010hb,Barausse:2011pu,Elliott:2005va},
shown in cyan in Figure~\ref{fig:LVconstraints} (right panel). Further
constraints come from requiring that the theoretically predicted Big
Bang nucleosyntesis elemental abundances agree with
observations~\cite{Carroll:2004ai,Yagi:2013qpa,Yagi:2013ava}; these
constraints are much stronger for Khronometric theory than for
Einstein-\AE{}ther~\cite{Carroll:2004ai,Jacobson:2008aj}, and are represented by
the orange region in Figure~\ref{fig:LVconstraints} (right panel). As
reviewed in that Figure and discussed
later in Section~\ref{sec:pulsar_LV}, even more stringent constraints
on $\lambda$ and $\beta$ come from binary pulsar
observations~\cite{Yagi:2013qpa,Yagi:2013ava}.
\subsubsection{Ho\v rava gravity}
The khronometric theory action \eqref{action-K} is particularly interesting because it is the low-energy (or infrared)
limit of Ho\v rava gravity~\cite{Horava:2009uw}, a renormalizable quantum field theory which has only spatial diffeomorphism invariance. The complete action of Ho\v rava gravity is~\cite{Blas:2009qj}
\begin{equation}
\label{SBPSHfull}
S_{\rm H}= \frac{1}{16\pi G_{\rm H}}\int dT d^3x \, N\sqrt{h}
\left(L_2+\frac{\hbar^2}{M_\star^2}L_4+\frac{\hbar^4}{M_\star^4}L_6\right)\,,
\end{equation}
where
\begin{equation}\label{L2}
L_2=K_{ij}K^{ij} - \frac{1+\lambda}{1-\beta} K^2
+ \frac{1}{1-\beta}{}^{(3)}\!R + \frac{\alpha}{1-\beta}\, a_ia^i
\end{equation}
is the Lagrangian density of Khronometric theory [c.f. Eq.~\eqref{action-K}],
$M_\star$ is a mass scale, and $L_4$ and $L_6$ are terms of fourth- and
sixth-order in the spatial derivatives, but contain no
derivatives with respect to the preferred time $T$.
Complete constraints on $M_\star$ are somewhat elusive to obtain, and are probably one of the most important open
questions in Ho\v rava gravity~\cite{Liberati:2012jf}. The reason is that one would expect Lorentz violations to
percolate from gravity into the matter sector, where Lorentz symmetry has been verified to high precision by
particle physics and cosmic-ray
experiments~\cite{Kostelecky:2003fs,Kostelecky:2008ts,Mattingly:2005re,Jacobson:2005bg,Liberati:2013xla}. However,
several mechanisms have been put forward to suppress this percolation. For instance, it has been suggested that the
operators that violate Lorentz symmetry in the matter sector might be finely tuned to much smaller values than those in
the gravity sector. Also, Lorentz invariance in matter might be an emergent property at low
energies~\cite{Froggatt:1991ft}, as an accidental symmetry~\cite{GrootNibbelink:2004za} or due to renormalization group
phenomena~\cite{Chadha:1982qq,Bednik:2013nxa} . Finally, it has been shown that two sectors with different Lorentz
violation degrees can easily coexist if their interaction is suppressed by a high mass scale~\cite{Pospelov:2010mp}, and
this could be the case for the gravity and matter sector. Therefore,
taking into account only the gravitational bounds
(i.e.~assuming that percolation of Lorentz violation into the matter sector is efficiently suppressed), one obtains
$M_{\star}\gtrsim 10^{-2}$ eV from sub-millimeter gravitational experiments. Also, perhaps surprisingly, $M_\star$
has an upper bound ($M_{\star}\lesssim 10^{16}$ GeV) from the requirement that the theory remains perturbative at all
scales~\cite{Papazoglou:2009fj,Kimpton:2010xi,Blas:2009ck}, so that the power-counting renormalizability arguments
proposed in~\cite{Horava:2009uw} apply.
Three things are worth stressing about the higher-order derivative terms $L_4$ and $L_6$ in the action. First, the
presence of sixth-order spatial derivatives is essential for power-counting renormalizability~\cite{Horava:2009uw}.
Second, the fourth and sixth order terms in the spatial derivatives generally lead to nonlinear dispersion relations
for the gravitational degrees of freedom of the theory, i.e.~the spin-2 and spin-0 gravitons (the latter present in
the theory because of the foliation-defining scalar $T$) satisfy
\begin{equation}
\label{mdsr}
\omega^2\propto k^2+\alpha_4 \left(\frac{\hbar}{M_\star}\right)^2 k^4+\alpha_6 \left(\frac{\hbar}{M_\star}\right)^4 k^6+\ldots\, ,
\end{equation}
where $\omega$ and $k$ are respectively the frequency and the wave-number, while $\alpha_4$ and $\alpha_6$
are dimensionless constants. Because such a dispersion relation allows for infinite propagation speeds in the
ultraviolet limit, the notion of a BH may appear problematic in these theories. However, we will return to this problem
in Section~\ref{sec:BH_LV} and show that the presence of a dynamical foliation-defining scalar $T$ actually allows for BHs
to be defined in this theory as well~\cite{Barausse:2011pu,Blas:2011ni,Barausse:2013nwa}. Third, aside from
instantaneous propagation at very high energies, the higher-order terms $L_4$ and $L_6$ are typically negligible in
astrophysical settings~\cite{Barausse:2013nwa}. These terms induce corrections on the spacetime geometry around
astrophysical objects that are of order ${\cal{O}}(G_{N}^{-2} M^{-2} M_\star^{-2}) \sim {\cal{O}}(M_{\rm Pl}^4/(M
M_{\star})^2)$ for an object of mass $M$, which translates into an error $\lesssim 10^{-16} (M_{\odot}/M)^2$.
\subsubsection{$n$-DBI gravity \label{subsec:n-dbi}}
Inspired by the approximate scale invariance of the Universe at early and late
times, when it is believed to be approximately
de Sitter, Herdeiro et al.~\cite{Herdeiro:2011im,Herdeiro:2011km} proposed a modification of GR that automatically results in
inflation at early times, without the need for additional scalar fields. This model, dubbed $n$-DBI gravity, was
designed so that it yields the Dirac-Born-Infeld type conformal scalar theory when the Universe is conformally
flat and resembles Einstein's gravity
in weakly curved space-times.
Interestingly, not only does it result in inflation, but it can also accommodate a smooth transition to radiation- and
matter-dominated epochs, followed by late time acceleration. The two distinct accelerating periods, with two distinct
effective cosmological constants, are a manifestation that the cosmological constant can vary in this theory. Moreover,
a large hierarchy between these two cosmological constants can be naturally achieved if the naive cosmological constant
appearing in a weak-curvature expansion of the theory is associated to
the TeV scale, which also suggests a new mechanism to
address the cosmological constant problem. The action for $n$-DBI
gravity is~\cite{Herdeiro:2011im}
\begin{equation}
S_{n{\rm DBI}}=-\frac{3\lambda}{4\pi G_N^2}\int d^4x\,\sqrt{-g}\,
\left\{\sqrt{1+\frac{G_N}{6\lambda}(R+\mathcal{K})}-q\right\}\,,
\qquad \mathcal{K}=-2\nabla_\mu(n^\mu\nabla_\nu n^\nu)\,.\label{action-nDBI}
\end{equation}
It contains two dimensionless parameters $\lambda$ and $q$ and an everywhere time-like vector field $\boldsymbol{n}$
coupled to the gravitational sector which breaks Lorentz invariance and makes the theory invariant under \emph{foliation
preserving diffeomorphisms}, in a way similar to Ho\v{r}ava-Lifschitz gravity. Concretely, if we perform an
Arnowitt-Deser-Misner (ADM) decomposition~\cite{Arnowitt:1960es}, then
$\boldsymbol{n}$ determines the lapse function $N$
through $n_\mu=-N\,dt$. This gives rise to a scalar degree of freedom, in addition to the two tensor polarizations of
GR~\cite{Coelho:2012xi}. Remarkably, the term $\mathcal{K}$ in (\ref{action-nDBI}) allows for the equations of motion to remain at most
second order in time derivatives, despite an infinite power series in the Ricci curvature.
Any solution of Einstein's gravity with cosmological
constant plus matter, admitting a foliation with constant
$R+\mathcal{K}$, is also a solution of $n$-DBI gravity. Moreover,
any Einstein space admitting a foliation with constant
$^{(3)}R-N^{-1}\Delta N$ (where $^{(3)}R$ is
the Ricci scalar of the 3-dimensional hypersurfaces) is a solution of $n$-DBI gravity~\cite{Herdeiro:2011im}. By
requiring spherical symmetry, one can explicitly obtain the Schwarzschild, Reissner-Nordstr\"om and (anti-)de Sitter BH
solutions, albeit in an unusual set of coordinates. Unlike GR, however, the cosmological constant is not determined at
the level of the action, but appears instead as an integration constant. The foliation condition of constant
$^{(3)}R-N^{-1}\Delta N$ can be interpreted as the {\it maximal slicing} gauge
condition common in numerical
relativity~\cite{Alcubierre:2008}, and it is then straightforward to show that
the Kerr metric in Boyer-Lindquist
coordinates is also a solution of $n$-DBI gravity~\cite{Coelho:2013zq}.
Since $n$-DBI gravity has a preferred foliation, one might expect the
PPN preferred-frame parameters $\alpha_1$ and $\alpha_2$ to be
nonvanishing.
Then experimental bounds on the PPN parameters should provide a lower bound for $\lambda$ which, together with
the estimate coming from inflation, would in principle define a finite
interval of viability $\lambda_{\rm PPN}<\lambda<\lambda_{\rm inf}$.
However, in a perturbative expansion about Minkowski, solutions of
$n$-DBI gravity coincide with those of GR and exist for all values of
$\lambda,q$~\cite{Coelho:2013dya}. Thus, at least to first PN order,
we have $\alpha_{1}=\alpha_{2}=0$, and $n$-DBI is indistinguishable
from GR in the Solar System.
\subsubsection{Power-counting and the semiclassical approximation.}
What does any of this have to do with the classical
approximation? The connection to EFTs arises for two reasons. First,
$S_{\rm eff}(l)$ enters into expressions in precisely the same way as
would a classical action; the influence of the heavy fields makes the
system behave at low energies {\em as if} its classical action were
$S_{\rm eff}(l)$. Second, much of the nitty gritty of EFT techniques
aims to identify how successive heavy-field corrections to $S_{\rm
eff}$ propagate through to contribute to observables, to make their
calculation as efficient as possible. The point is that these same
techniques can be used to track which combinations of parameters arise
order-by-order in the loop expansion, and so whose small size
ultimately justifies this expansion. Since classical physics is just
the leading (nonloop) contribution, such arguments also justify when
it suffices to stop with a classical result.
It is worth illustrating this with a specific example. A particularly
simple class of observables for a low-energy gravity theory consist of
the scattering of weakly coupled gravitons moving through a weakly
curved classical geometry. If $c_4 = 0$ we can take the background
geometry to be flat space,\footnote{The assumption of flatness here is
purely for convenience, and the conclusions below apply equally well
to curved spaces, since they rely essentially on dimensional
arguments.} allowing us to expand the metric around the Minkowski
background: $g_{\mu\nu} = \eta_{\mu\nu} + h_{\mu\nu}$. We then ask how
each of the terms in the gravitational action, Eq.~\eqref{gravaction},
contribute. In particular, with a view to asking how large quantum
corrections can be, we can ask about the relative size of different
contributions to the amplitude, ${\cal A}(E)$, for 2 gravitons to
scatter into another 2 with energy $E$.
As shown in~\cite{Burgess:2003jk,Goldberger:2007hy,Donoghue:2012zc} in
some detail, the contribution to this amplitude of an $L$-loop Feynman
graph built using $V_{i,r}$ vertices built from a term in $S_{\rm eff}$
involving $r$ powers of the curvature tensor, involving the emission
or absorption of $i$ gravitons, is of order
\begin{equation}
\label{GRcount}
{\cal A}(E) \sim \left( {E \over M_{\rm Pl}} \right)^2 \left(
{E \over 4 \pi M_{\rm Pl}} \right)^{2 L} {\prod_{i} \prod_{r\ge 2}}
\left[{E^2 \over M_{\rm Pl}^2} \left( {E \over M} \right)^{(2r-4)}
\right]^{V_{i,r}} \,.
\end{equation}
Here $M$ is the mass that sets the dimensions of the coefficients
$c_{(d,k)} \propto M^{d}$ for $d < 0$, which is assumed for simplicity
to be the same order of magnitude for all negative
$d$. Eq.~(\ref{GRcount}) has several useful consequences.
First, because $r \ge 1$ for all terms in Eq.~(\ref{gravaction}) the
contribution to ${\cal A}$ contains no negative powers of $E$. This
illustrates how $S_{\rm eff}$ encapsulates how observables
simplify in the hierarchical low-energy limit, where $E \ll M$,
$E\ll M_{\rm Pl}$. In particular, this expression quantifies why the weakness of the
graviton's coupling follows purely from the low-energy approximation,
$E \ll M_{\rm Pl}$ and $E \ll M$.
Second, these expressions identify precisely which kinds of
interactions dominate scattering amplitudes at low energies. The
minimum suppression by powers of $E$ comes when $L = 0$ and we choose
$V_{i,r} = 0$ unless $r = 1$, and so is given by arbitrary tree graphs
constructed purely from the Einstein-Hilbert action. This tells us
what we would be inclined to believe in any case: it is $L = 0$
(no-loop) graphs built only from the Einstein-Hilbert action --
i.e.~classical GR -- which govern the low-energy dynamics of
GWs, giving a result of order $(E/M_{\rm Pl})^2$.
But we may also identify the next-to-leading contributions. These
are proportional to $(E/M_{\rm Pl})^4$ and can appear in one of two ways:
\begin{enumerate}
\item either: $L = 1$ and $V_{i,r} = 0$ for any $r \ne 1$,
\item or: $L = 0$ with $\sum_i V_{i,2} = 1$ and $V_{i,1}$ arbitrary, and
$V_{i,r} = 0$ for all $r \ge 3$.
\end{enumerate}
That is, the next-to-leading contribution is obtained by computing the
one-loop corrections using only Einstein gravity, or by working to
tree level and including precisely one curvature-squared interaction
in addition to any number of interactions from the Einstein-Hilbert
term. Both are suppressed compared to the leading term by a factor of
$(E/M_{\rm Pl})^2$. At this order the ultraviolet divergences that famously
plague gravitational loops in option (i) above are absorbed into
renormalizations of the coefficients of the curvature-squared
contributions that appear in option (ii), and so on further down the
$E/M_{\rm Pl}$ expansion.
These conclusions are borne out by explicit calculations. At tree
level the only nonzero amplitudes are related by crossing symmetry to
the amplitude for which all graviton helicities have the same sign,
and this is given by~\cite{DeWitt:1967yk,DeWitt:1967ub,DeWitt:1967uc}:
\begin{equation}
-i {\cal A}_{(++,++)}^{\rm tree} = 8 \pi G \,\left(\frac{s^3}{tu} \right)\,,
\end{equation}
where $s$, $t$ and $u$ are the usual Mandelstam invariants built from
inner products of the graviton four-momenta, all of which are
proportional to the square of the center-of-mass energy, $E_{\rm
cm}$. This shows that it is the frame-independent center-of-mass
energy that appears in the $E/M_{\rm Pl}$ expansion of ${\cal A}$. The
one-loop corrections are also computed~\cite{Dunbar:1994bn}, and are
infrared divergent. These infrared divergences cancel in the usual way
with tree-level Bremsstrahlung diagrams~\cite{Weinberg:1965nx}, leading to a
finite result~\cite{Donoghue:1999qh}, which is suppressed as expected
relative to the tree contribution by terms of order $(E/M_{\rm Pl})^2$, up to
logarithmic corrections.
It is expressions like the amplitude scaling~\eqref{GRcount} that make
the explicit connection between EFTs and the domain of validity of the
semi-classical (or loop) approximation. This expression reveals that
the loop expansion for gravity is secretly a low-energy
approximation. This turns out to be generic for any nonrenormalizable
field theory~\cite{Weinberg:1978kz}. For such theories the only known
way to extract sensible quantum corrections is within a low-energy
approximation, for which the classical action should be regarded as a
general derivative expansion along the lines of
Eq.~\eqref{gravaction}. All terms in this action consistent with
symmetries and field content are compulsory, since their presence is
required to renormalize the ultraviolet divergences that are generated
by loops involving terms arising at lower orders in the derivative
expansion.
\subsubsection{Modified gravity seen through EFT glasses.}
We can now return to our road map of modified gravity theories to see
what it leads us to expect. Following~\cite{Burgess:2009ri}, we argue
that EFT can provide useful guidelines.
\paragraph{New particles and/or dimensions.}
The most conservative modifications simply involve
the addition of new light particles or the addition of more dimensions
(or both), with the new additions resembling those about which we
already know. It is certainly true that such modifications can be
sensible in principle, and explicit examples exist (such as string
theory) for higher-energy physics that can produce such
modifications. It makes sense to constrain such possibilities
observationally.
There are also issues that can be expected to arise in such theories
if the new particles are light enough to be relevant over
astrophysical or cosmological distances. This is true in particular
for proposals meant to describe present-epoch dark
energy~\cite{Weinberg:1988cp,Polchinski:2006gy,Burgess:2013ara}. Such
particles are so close to massless that many of the constraints on
massless particles in practice are likely to apply. In particular, it
can be expected that in the Lorentz-invariant framework of Special
Relativity the new particles must be spin zero, half or must be gauge
particles with spin one or $\frac32$ or smaller.
Another problem potentially can also arise, associated with the size
of quantum corrections to the mass, particularly for spinless
particles represented by a scalar field, $\phi$. Then the low-energy
EFT contains a mass term of the form ${\cal L}_{\rm eff} = -
c_{(2,2)} \, \sqrt{-g} \; \phi^2$ whose coefficient $c_{(2,2)}
\propto M^2$ should be large, for the same reasons (given above) that
lead one to expect $c_{(2,1)}$ and $c_4$ are large. This is a
``hierarchy'' problem, similar to the cosmological constant problem;
very light spinless particles very rarely arise as the low-energy
limit of something more fundamental because their masses are sensitive
to quantum contributions from every heavy state at higher energies
that is integrated out.
A similar problem does not occur for spin-half particles, because
for these the particle mass can be forbidden by a chiral symmetry,
under which the fermion's left- and right-handed components rotate
differently: $\psi \to i \gamma_5 \psi$. Because of this it can only
receive quantum corrections from particles that also break this
symmetry. Only very few symmetries (supersymmetry and scale
invariance) are known that can forbid a scalar mass, making this
mechanism more challenging to use at low energies for spinless
particles.
\paragraph{Modifications to the Einstein equations.}
Short-distance modifications to the left-hand side
of Einstein's equations are also very plausible, since these can
easily be generated by integrating out various kinds of heavy
fields. In well understood situations these usually lead to modified
actions along the lines of Eq.~(\ref{gravaction}) that are
local polynomials of the metric and its derivatives, and involve all
possible kinds of interactions allowed by the assumed symmetries. In
particular, it should be noted that generic higher-derivative
interactions are allowed, and explicit
calculations~\cite{Burgess:2014lwa} show these need not take the
specific Horndeski
or Lovelock forms that are sometimes advocated as being required to
avoid the presence of ghosts. What is hard to achieve in this way are
modifications like $f(R)$ gravity where $f(R)$ is an unusual function,
such as $1/R$. Proposals such as this one step away from the
underlying EFT understanding of the validity of the semi-classical
approximation, and so the onus is on proponents to justify the regime
of validity of any classical approximation. This is particularly so in
situations like dark energy proposals, where one of the basic problems
(the cosmological constant problem) cannot be seen until quantum
effects are examined.
\paragraph{Breaking diffeomorphism invariance.}
As described above, it is ultimately the consistency
of Lorentz invariance and quantum mechanics that drives many of the
consistency conditions for massless (and very light) particles,
including the requirements for the gauge invariance of their
interactions. However the constraints are no longer quite as exacting
once the particles are not exactly massless. In this case the general
consistency issues can be expected to persist if the particle is light
enough compared with the higher scales of the theory, $m \ll M$, but
can be evaded if this hierarchy is not too large.
This observation has prompted some to put aside until later
understanding the embedding into higher-energy physics, and instead to
explore the implications of relaxing the assumptions of gauge
invariance (and so usually also Lorentz invariance) at low
energies. The hope is to find a consistent low-energy effective
description that applies only up to relatively low energy scales, and
hope that once this is done a consistent ultraviolet completion can be
found. In most cases of this type, no candidate ultraviolet extension
is yet known.
Modifications to gravity provide a rich theoretical laboratory as to
how quantum field theories work, that display their tight consistency
issues in new and instructive situations. Sensible modifications --
i.e.~those that can be embedded into well-understood ultraviolet
completions -- are the goal, but are also not that easy to come
by. Together with their success in describing astrophysical and
cosmological observations, theoretical soundness should be regarded as
one part of the evidence to be used when assessing the likelihood of
such theories describing Nature.
\subsubsection{Stellar mass objects}
\label{sec:sub:stellarmass}
For stellar-mass compact binaries, the assumptions listed above
neglect the delicate but nontrivial impact of composition and initial
conditions, modifying the expected signal at both high and low
frequencies.
Just as these effects mimic or mask our ability to measure parameters
(see e.g.~\cite{Favata:2013rwa}), they can also mimic or mask
modifications to GR, weakening our ability to test it with real
astrophysical systems.
Conversely, astrophysical processes like precessing spins and
eccentricity~\cite{2014ApJ...784...71S,O'Leary:2007qa,2011MNRAS.416..133D,Morscher:2012se,2014ApJ...781...45A,2014MNRAS.439.1079A}
introduce more structure in the gravitational waveforms, potentially
enabling stronger and new constraints on both astrophysics and
modified gravity.
As outlined above, existing proposals to test modified theories of
gravity with second-generation GW detectors such as Advanced LIGO and
Advanced Virgo rely on carefully monitoring the phase evolution of the
leading-order GW harmonic (see
e.g.~\cite{DelPozzo:2013ala,2014PhRvD..89f4037S}). NS-NS binaries are
the preferred laboratory for testing GR because (a) they have been
observed in the electromagnetic spectrum, and therefore (unlike
stellar-mass BH binaries) their rates are constrained by
observations~\cite{Abadie:2010cf,Dominik:2014yma}; (b) most of the
SNR is accumulated during the inspiral, which is
both analytically tractable and less sensitive to systematic effects,
due to uncertainties in the modeling of the late inspiral and merger;
and (c) for most formation scenarios, astrophysics suggests that the
gravitational waveforms will be relatively simple (e.g., the
eccentricity of the orbit and the spin of NSs is small).
Individual events with large SNR could prove
definitive, but some techniques to test GR build statistical
confidence by ``stacking'' many individual events, searching for a
common signature~\cite{Damour:2012yf,DelPozzo:2013ala,Wade:2014vqa}.
Astrophysical effects not included in these models, if sufficiently
common, might mimic or mask the effects of modified gravity.
Conversely, the near-universal ``I-Love-Q'' relationships discussed in
Section~\ref{subsec:ILQ} suggest that, to leading order, EOS-dependent
effects can be encapsulated in a handful of
parameters~\cite{Yagi:2013bca,Yagi:2013awa,Yagi:2014bxa}, potentially
enabling direct tests of strong-field gravity even in the presence of
nontrivial and poorly constrained matter physics.
According to the above discussion, a test of strong-field gravity
would seem to require a detailed understanding of all possible
astrophysical influences, including the nuclear EOS and tidal
deformability of NSs. In fact, as discussed in \S~\ref{subsec:ILQ},
strong relationships exist between matter-sourced ``tidal'' multipoles
of different orders, both in GR and in a broad class of
modified-gravity theories. Specifically, the Q-Love relation may help
us to break the degeneracy between the NS spin and quadrupole moment
in GW observations.
To see this, let us consider GWs from a NS-NS binary. The quadrupole
moment first enters at 2PN order in the waveform
phase~\cite{Poisson:1997ha,Mikoczi:2005dn}, together with the NS
spin-spin coupling
term~\cite{Kidder:1992fr,Mikoczi:2005dn,Arun:2008kb}, which leads to a
strong degeneracy between the NS quadrupole moment and spin. On the
other hand, the NS finite-size effect enters first at 5PN
order~\cite{Flanagan:2007ix} through the tidal Love number, which can
be measured with second-generation ground-based interferometers such
as Advanced
LIGO~\cite{Flanagan:2007ix,Read:2009yp,Hinderer:2009ca,Lackey:2011vz,Damour:2012yf,Lackey:2013axa,Read:2013zra,
Favata:2013rwa,Yagi:2013baa,Yagi:2013sva,Wade:2014vqa}. The Q-Love
relation allows us to express the quadrupole moment in terms of the
Love number, leading to degeneracy breaking.
This observation suggests that strong-field gravity could be coarsely
tested even \emph{without knowing the nuclear EOS}.
Since (with notable exceptions, such as scalar-tensor theories and
EiBI theory) the functional form of the I-Love-Q relations depends on
the underlying gravitational theory, if one can measure {\em any two}
of the I-Love-Q quantities independently, one can in principle perform
a model-independent consistency test of GR or test a specific
alternative theory. These tests could exploit combined observations of
binary pulsars in the electromagnetic spectrum and binary inspirals in
the GW spectrum. For example, future observations of the double binary
pulsar~\cite{Burgay:2003jj,Lyne:2004cj} may measure the NS moment of
inertia to an accuracy $\sim
10\%$~\cite{Lattimer:2004nj,Kramer:2009zza}, and the tidal Love number
may be measured to an accuracy $\sim 60\%$ with GW
observations~\cite{Yagi:2013awa}. These measurements would identifiy a
measurement point with an error box in the I-Love plane (see the left
panel of Figure~\ref{fig:I-Love-error}). If the I-Love relation is
modified from GR in a specific alternative theory, such a theory can
only be valid if the ``modified I-Love relation'' is consistent with
the error box.
A problem with this idea is that (in the example considered here) the
double binary pulsar and the hypothetical GW observations would target
different NSs, while the universal relations are valid for any single
star. However, since the parameter that is fixed for different models
satisfying the universal relation is the NS mass (or compactness), the
relations would still hold if the NSs in the two systems have the same
mass. This is not too unlikely, given that NSs have a rather peaked
mass distribution~\cite{Lattimer:2010zz,Steiner:2012xt}.
Moreover, even if the two NSs have different masses, it turns out that
the EOS-universality is preserved to good accuracy for NS models with
fixed mass ratio~\cite{Yagi:2013awa}.
To exploit these opportunities, then, requires care in identifying
possible confusing astrophysical degrees of freedom, their expected
magnitude, and methods to robustly disentangle them from the
signatures of modified gravity. A useful way to illustrate how
different astrophysical degrees of freedom come into play in different
frequency bands is shown in Figure~\ref{fig:DamourNagarVillain}.
\begin{figure}[tb]
\capstart
\begin{center}
\includegraphics[width=8cm]{DamourNagarVillain_fig3-eps-converted-to.pdf}
\caption{Integrands, per frequency octave, of the integrals determining the SNR
$\rho$ and the measurement accuracy of the chirp mass ${{\cal M}}$,
the symmetric mass ratio (denoted by $\nu$ in the figure legend, and
by $\eta$ elsewhere in this review), and the tidal parameter
$\lambda_T\propto \kappa_2^T$. These integrands are plotted as a
function of the rescaled frequency $\hat{f}=f/(56.56 {\rm Hz})$ for a
typical $1.4M_\odot + 1.4M_\odot$ binary of two NSs with equal
compactness $M/R=0.1645$ ($M$ and $R$ being the mass and the
radius of each star). While most of the SNR is gathered around
frequencies $\hat{f}\sim 1$, the measurability of ${\cal M}$ and
$\eta$ is concentrated towards lower frequencies ($\hat{f}<1$), and
that of the tidal parameter $\lambda_T$ gets its largest contribution
from the late inspiral up to the merger. The rightmost vertical line
indicates the merger frequency, while the leftmost vertical line
corresponds to a GW frequency of 450~Hz. [From~\cite{Damour:2012yf}.]}
\label{fig:DamourNagarVillain}
\end{center}
\end{figure}
Using a Fisher matrix approximation to parameter estimation accuracy,
this figure illustrates that different frequency bands of the GW
signal encode information about different astrophysical parameters,
namely the chirp mass, the symmetric mass ratio and the tidal (Love)
parameter.
Extending this argument to
eccentricity~\cite{Yunes:2009yz,1995PhRvD..52.2089K,2011PhRvD..84l4007G,Huerta:2014eca},
precession~\cite{2014PhRvD..89d4021L} and merger, one finds the
following hierarchy: information on small residual eccentricity and
chirp mass is encoded at low frequencies (being tied to the overall
number of cycles); information on mass ratio and spin is encoded at
intermediate frequencies; finally, information on tidal interactions
and strong-field effects comes mostly from the highest frequencies.
Similarly -- and roughly speaking -- different modifications to
gravity also predominantly occur at different frequencies, i.e., at
different PN orders (cf.~Section~\ref{sec:paramtests}).
Qualitatively, each frequency scale couples strongly to itself and
neighboring scales; for example, mass ratio and (aligned) spins are
strongly degenerate~\cite{Berti:2004bd}.
For this physical reason, we expect that confusing effects of
astrophysical phenomena are principally entangled with modifications
to gravity that dominate at the corresponding frequency interval or PN
order (along with all other strongly coupled degrees of freedom).
However, because each process has distinctive radiation content (e.g.,
higher harmonics; precession-induced modulations), astrophysical
calculations within GR suggest that these degeneracies can be broken.
This will require more precise modeling of gravitational waveforms in
various proposed extensions of GR, and further study to quantify
precisely what beyond-GR properties are accessible to experiment in
the presence of all astrophysical parameters that may contaminate the
signal.
\begin{figure}[tb]
\centering
\includegraphics[width=0.7\textwidth]{hist_inj_tf2_tt4_rec_tf2_1sources.pdf}
\caption{Single-source background distributions where the
simulated sources are generated by the \texttt{TaylorF2}
waveform approximant (blue) or by the \texttt{TaylorT4}
approximant (red).}
\label{fig:waveformMismatch}
\end{figure}
In the case of NS-NS coalescence, accurate waveforms have been
available for some time now. As shown in~\cite{Buonanno:2009zt}, in
the NS-NS mass regime the PN waveform approximants agree with each
other and with effective-one-body waveforms to a high degree of
accuracy. Nevertheless, it was checked explicitly that (at least for
NS-NS sources) the remaining small differences between approximants
will not cause one to declare a violation of GR when none is
present. Figure~\ref{fig:waveformMismatch} shows single-source
background distributions where the sources are simulated using
the \texttt{TaylorF2} waveform family (red) and the
\texttt{TaylorT4} waveform family (blue).
We see that the impact of the waveform mismatch on the background
distribution indeed appears to be minimal.
To summarize, binary NSs provide a relatively clean laboratory to
investigate modifications to strong-field GR. Unfortunately, in these
systems, non-GR deviations in the GW signal could be degenerate with
tidal effects, eccentricity and spin.
While tidal effects are strong and EOS-dependent at late times, tidal
interactions seem to be relatively universal, potentially enabling
tests of GR without detailed knowledge of the properties of nuclear
matter.
\paragraph{Higher-mass objects: Future directions.}
Quasicircular NS-NS binaries are not the only target of Earth-based GW
interferometers. Binaries containing BHs have larger total mass than
NS-NS binaries, and their GW signal encodes more nonlinear
strong-field dynamics in the LIGO sensitivity band. Fortunately, even
though massive binaries may be intrinsically rare, GW detectors are
expected to be much more likely to observe
them~\cite{Abadie:2010cf,Dominik:2014yma}.
NS-NS coalescences are relatively simple from the observational point
of view: it is mostly the inspiral part that is visible in the
detectors' sensitive frequency band, and NSs in binaries are expected
to have small spins. By contrast, for NS-BH and BH-BH coalescences
the full process of inspiral, merger, and ringdown will be visible.
Moreover, BH spins will likely be large and strongly precessing. Hence
NS-BH and BH-BH events are dynamically much richer than NS-NS, but
this also makes it more challenging to use them in tests of GR. One
problem has been the unavailability of accurate and faithful
(semi-)analytical waveform models, though recently there has been
great progress in that regard. For instance, Pan et al.~arrived at an
effective-one-body model with full precessing spins that appears to
have great faithfulness with waveforms obtained from large-scale
numerical simulations~\cite{2014PhRvD..89f1501P}, although it is still
rather costly to generate on a computer. Other options could include
(improvements of) the phenomenological time-domain
\texttt{PhenSpin} waveforms of Sturani et al.~\cite{Sturani:2010yv,Sturani:2010ju}, or the frequency-domain (and
hence computationally cheap) waveforms for precessing spins developed
both for inspiral \cite{2014PhRvD..89d4021L} and for
inspiral-merger-ringdown~\cite{Hannam:2013oca,Schmidt:2014iyl}.
The application of these waveforms to parameter estimation and model
selection of alternative theories of gravity is an exciting research
topic for the future.
\subsection{Scalar-tensor theories}
\label{sec:binaries/scal-tens-theor}
Two complementary approaches are used to model compact binaries in GR:
analytical calculations (usually PN expansions) and numerical
relativity simulations. The first approach is valid at large
separations, and for orbital velocities that are small compared to the
speed of light; the second is necessary when the components of the
binary system get close and merge. Both approaches have been extended
to scalar-tensor theories and will be discussed in this section. We
will focus mostly on the best studied case where the gravitational
interaction is mediated by the metric and a single scalar field
(see~\cite{Damour:1992we} for a pioneering study of binary dynamics in
tensor-multiscalar theories).
\subsubsection{Analytical calculations}\label{sec:pn_st}
Comprehensive studies of compact binaries in scalar-tensor theory were
carried out by Eardley~\cite{Eardley:1975}, Will and
Zaglauer~\cite{Will:1977wq,Will:1989sk} and Damour and
Esposito-Far\`{e}se~\cite{Damour:1992we,Damour:1996ke,Damour:1998jk},
among others. At present, the most accurate description of the orbital
motion and GW emission in scalar-tensor theory is due to work by
Mirshekari and Will~\cite{Mirshekari:2013vb} and
Lang~\cite{Lang:2013fna,Lang:2014osa}, who derived the equations of
motion and the radiation flux considering a single
scalar field with vanishing potential (and therefore vanishing
mass). The relevant field equations for this theory are given in
Eq.~\eqref{ST:Jordan}, with $U(\phi) = 0$.
Following Eardley~\cite{Eardley:1975}, Mirshekari, Will and Lang
describe the orbital motion of the binary in terms of the NS
sensitivities (reviewed in Section~\ref{sec:sensitivities}).
The calculation considers only the inspiral phase of the binary's
evolution. It is therefore appropriate to use the PN approximation, an
expansion in powers of $v/c \sim (Gm/rc^2)^{1/2}$. The scalar-tensor
equations are solved by adapting the ``direct integration of the
Einstein equations'' (DIRE) method developed by Will, Wiseman, and
Pati~\cite{Will:1996zj,Pati:2000vt,Pati:2002ux}, which has proven
successful in GR and which has been extended to scalar-tensor theory (see~\cite{PW:2014} for a pedagogical
introduction). To begin with, it is convenient to consider a rescaled
version of the scalar field $\phi$: $\varphi \equiv \phi/\phi_0$, where
$\phi_0$ is the value of $\phi$ at infinity (assumed to be
constant).\footnote{The rescaled (Jordan-frame) field, denoted by
$\varphi$ in this section, should not be confused with the
Einstein-frame field, denoted by the same symbol elsewhere in this
review.} By introducing a new tensorial quantity
\begin{equation}
\tilde{h}^{\mu \nu} \equiv \eta^{\mu \nu} - \sqrt{-\tilde{g}}\tilde{g}^{\mu \nu} \,
\end{equation}
(where $\eta^{\mu \nu}$ is the inverse Minkowski metric,
$\tilde{g}_{\mu \nu} \equiv \varphi g_{\mu \nu}$, and $\tilde{g}$ is
its determinant) and choosing the gauge condition
$\tilde{h}^{\mu \nu}_{\hphantom{\mu \nu},\nu} = 0$, the
field equations reduce to two {\em flat-spacetime} wave
equations,
\begin{align}
\Box_\eta \tilde{h}^{\mu \nu} &= -16\pi \tau^{\mu \nu} \, , \\
\Box_\eta \varphi &= -8\pi \tau_s \, ,
\end{align}
where the sources $\tau^{\mu \nu}$ and $\tau_s$ on the right-hand side
contain terms depending not only on the matter stress-energy tensor $T^{\mu
\nu}$, but also on the fields $\tilde{h}^{\mu \nu}$ and $\varphi$.
The formal solution of these ``relaxed'' Einstein equations can be
written down using the usual flat-spacetime retarded Green's function,
\begin{subequations}
\label{eq:retardedGF}
\begin{align}
\tilde{h}^{\mu \nu}(t,\mathbf{x}) &= 4\int \frac{\tau^{\mu \nu}(t',\mathbf{x}')\delta(t'-t+|\mathbf{x}-\mathbf{x}'|)}{|\mathbf{x}-\mathbf{x}'|}d^4x' \, ,\label{eq:hintegral}\\
\varphi(t,\mathbf{x}) &= 2\int \frac{\tau_s(t',\mathbf{x}' )\delta(t'-t+|\mathbf{x}-\mathbf{x}'|)}{|\mathbf{x}-\mathbf{x}'|}d^4x' \, .
\label{eq:phiintegral}
\end{align}
\end{subequations}
The main qualitative difference with electromagnetism is that the
sources in these integrals do not have compact support. In the DIRE
method (and extensions thereof), spacetime is split into two regions, the ``near
zone'' and the ``radiation zone.'' In the near zone close to the source (at
distances smaller than the typical gravitational wavelength $\lambda$,
$|\mathbf{x}'|<\mathcal{R}$ where $\mathcal{R}\sim \lambda$), the
integral is calculated using a slow-motion approximation via the
usual PN expansion in powers of $v/c$. In the radiation zone far from
the source (at distances $|\mathbf{x}'|>\mathcal{R}$), a change of
variables is used to evaluate the integral. The integration procedure
is different depending on whether the field point $\mathbf{x}$ itself
lies in the near zone or the radiation zone, so there are four different
classes of integrals.
In general, the integrals will produce terms which depend on the arbitrary quantity
$\mathcal{R}$, but it is safe to ignore these terms because any
$\mathcal{R}$-dependent contributions from near-zone integrals must be
completely canceled by contributions from radiation-zone integrals.
The first step to understanding compact binary systems is to find the
equations of motion for the bodies. Mirshekari and Will~\cite{Mirshekari:2013vb} carried out this calculation up to 2.5PN
order, or $\mathcal{O}((v/c)^5)$. Their procedure requires evaluating the
integrals \eqref{eq:retardedGF} at field points in the near zone,
where the bodies are located. The procedure is
iterative: The lowest-order source, comprising only the compact object
stress-energy, is used to find the lowest-order fields. These are
then substituted in to find the next-highest-order source, which can
then be used to find the next-highest fields, and so on. Once the
fields, and thus the metric, have been calculated to the necessary
order, the equations of motion are found from the geodesic equations
(with a slight modification due to the $\phi$-dependence of mass in
the Eardley approach). Schematically, the relative acceleration $\mathbf{a} \equiv \mathbf{a}_1-\mathbf{a}_2$ takes the form
\begin{align}
a^i =& -\frac{G\alpha m}{r^2}\hat{n}^i+\frac{G\alpha m}{r^2}(A_\text{PN}\hat{n}^i+B_\text{PN}\dot{r}v^i)+\frac{8}{5}\eta\frac{(G\alpha m)^2}{r^3}(A_\text{1.5PN}\dot{r}\hat{n}^i-B_\text{1.5PN}v^i) \nonumber\\
&{}+\frac{G\alpha m}{r^2}(A_\text{2PN}\hat{n}^i+B_\text{2PN}\dot{r}v^i) \, ,
\end{align}
where $m \equiv m_1+m_2$, $\eta \equiv m_1m_2/m^2$, $r$ is the orbital separation,
$\mathbf{\hat{n}}$ is a unit vector pointing from body 2 to body 1,
and $\mathbf{v} \equiv \mathbf{v}_1-\mathbf{v}_2$ is the relative
velocity. The (typically time-dependent) coefficients $A_\text{PN}$,
$B_\text{PN}$, $A_\text{1.5PN}$, $B_\text{1.5PN}$, $A_\text{2PN}$, and
$B_\text{2PN}$ are given in~\cite{Mirshekari:2013vb}. We use the
symbol $G$ to represent the combination
$(4+2\omega_0)/[\phi_0(3+2\omega_0)]$ [with $\omega_0 \equiv
\omega(\phi_0)$] because it appears in the metric component $g_{00}$
in the same manner as the gravitational constant $G$ in GR. However,
the coupling in the Newtonian piece of the equations of motion is not
simply $G$ but $G\alpha$, where
\begin{equation}
\alpha \equiv \frac{3+2\omega_0}{4+2\omega_0}+\frac{(1-2s_1)(1-2s_2)}{4+2\omega_0} \,
\end{equation}
and $s_i$ ($i=1\,,2$) are the sensitivities of the two objects,
defined in Eq.~\eqref{sensitivity}. Another important deviation from
GR is the presence of a 1.5PN radiation-reaction contribution to the
equations of motion. In GR, radiation reaction begins at 2.5PN order,
with the lowest-order quadrupole radiation contribution. In
scalar-tensor theory, radiation reaction begins at 1.5PN order, due to
the presence of scalar dipole radiation.
Many other deviations from GR occur within the $A$ and $B$
coefficients: see~\cite{Mirshekari:2013vb} for details. Although the
number of deviations is large, they can all be characterized using a
fairly small number of parameters, all combinations of $\phi_0$, the
Taylor coefficients of $\omega(\phi)$, and the sensitivities $s_A$,
$s_A'$, and $s_A''$. The deviations are considerably simplified if
one object in the system is taken to be a BH (with the other being a
NS). Then the motion of the system is indistinguishable from the
motion in GR up to 1PN order. All deviations beyond 1PN order depend
only on a single parameter, which itself depends on $\omega_0$ and
the sensitivity of the NS. Unfortunately, this parameter does not depend on any
more details of the scalar-tensor theory; if measured, it alone could
not be used to distinguish between Brans-Dicke theory and a more
general scalar-tensor theory.
The equations of motion simplify even more radically for a binary BH
system: They are {\em identical} to the equations of motion in GR,
except for an unobservable mass rescaling.
This result is a generalization to binary systems of
``no-scalar-hair'' theorems that apply to single
BHs~\cite{Hawking:1972qk}. For generic mass ratio, Mirshekari
and Will proved this ``generalized no-hair theorem'' up to 2.5PN
order, but they conjectured that it should hold at all PN orders.
Indeed, Yunes et al.~have shown that the equations of motion are the
same as in GR at any PN order if one considers an extreme mass-ratio
system and works to lowest order in the mass
ratio~\cite{Yunes:2011aa}, and the conjecture is also supported by
numerical relativity studies~\cite{Healy:2011ef,Berti:2013gfa} (see
Section~\ref{sec:nr_st}). This ``generalized no-hair theorem'' for
binary BHs depends on some crucial assumptions: vanishing scalar
potential, asymptotically constant value of the scalar field, and
vanishing matter content. If any one of these assumptions breaks down, the BH binary's behavior will differ from GR.
The next step is the calculation of gravitational radiation. The
tensor part of the radiation, encoded in $\tilde{h}^{ij}$, was
computed up to 2PN order by Lang~\cite{Lang:2013fna}. The procedure
requires evaluating Eq.~\eqref{eq:hintegral} for field points in the
``far-away zone,'' a subset of the radiation zone which is very far
($R \equiv |\mathbf{x}| \gg \mathcal{R}$) from the source. When integrating over source points in the near zone,
the main step is the calculation of certain moments of the source
$\tau^{ij}$, known as ``Epstein-Wagoner moments.'' The first of these,
the quadrupole moment, generates GW contributions at 0PN, 1PN, 1.5PN,
and 2PN orders. (The 1.5PN order contribution does not occur in GR
and is a direct result of scalar dipole radiation in this theory.)
The next moment is the octupole moment, which generates GWs at 0.5PN,
1.5PN, and 2PN orders. In all, Epstein-Wagoner moments with up to 6
indices are required. The 6-index moment contributes only 2PN GWs.
The final expression for the tensor waves is considerably more
complicated than its GR equivalent; however, all deviations depend on
the same small number of parameters that characterize the equations of
motion. Most deviations appear as modifications to GR terms, except
for entirely new terms which depend on the existence of a scalar
dipole moment. The tensor
waves also show the same behavior as the equations of motion in
special cases. For BH-NS systems, the waveform is indistinguishable
from GR up to 1PN order; deviations at higher order depend only on the
single parameter described earlier. For BH-BH systems, the waveform is completely indistinguishable from GR.
Scalar radiation has recently been computed by Lang~\cite{Lang:2014osa}
using a very similar procedure. The near-zone contribution requires
the calculation of ``scalar multipole moments'' similar to the
Epstein-Wagoner moments, but involving $\tau_s$ instead of
$\tau^{ij}$. In this case, the lowest-order moment is not the
quadrupole, but the monopole. Using the standard definition of PN
orders, in which ``0PN'' waves are generated by the tensor quadrupole,
the scalar monopole moment generates a scalar field at $-1$PN order.
This field, however, turns out to be time-independent and not
wavelike. The dipole moment generates the lowest-order scalar waves,
which are of $-0.5$PN order:
\begin{equation}
\varphi = \frac{4G\mu\alpha^{1/2}}{R}\zeta\mathcal{S}_-(\mathbf{\hat{N}}\cdot \mathbf{v})
\, ,
\end{equation}
where $\mu \equiv m_1 m_2/m$ is the reduced mass, $\mathbf{\hat{N}}
\equiv \mathbf{x}/R$ is the direction from the source to the detector,
$\zeta \equiv 1/(4+2\omega_0)$, and
\begin{equation}
\mathcal{S}_- \equiv \alpha^{-1/2}(s_2-s_1) \, .
\end{equation}
Calculating the radiation up to 2PN order requires knowledge of the
monopole moment to 3PN order (relative to itself) and knowledge of the
dipole moment to 2.5PN order. Just constructing the 3PN expansion of
the source $\tau_s$ is a challenging process. Evaluating the
resulting integrals is even more difficult. For these reasons,
Lang~\cite{Lang:2014osa} computes the scalar waveform only to 1.5PN
order, with the 2PN result saved for future work. The 1.5PN waveform
turns out to be described by the same set of parameters that describes
the 2.5PN equations of motion and the 2PN tensor waveform. Other
similarities include the vanishing of the scalar waveform for binary
BH systems (so that it is indistinguishable from GR) and tremendous
simplifications in the mixed BH-NS case.
The tensor and scalar waveforms can be used to compute the total
energy carried off to infinity using the expressions
\begin{align}
\frac{dE_T}{dt} &= \frac{R^2}{32\pi}\phi_0\oint \dot{\tilde{h}}_\text{TT}^{ij}\dot{\tilde{h}}_\text{TT}^{ij} d^2\Omega \, , \label{eq:tensorfluxeqn}\\
\frac{dE_S}{dt} &= \frac{R^2}{32\pi}\phi_0(4\omega_0+6)\oint \dot{\varphi}^2 d^2\Omega \, ,
\label{eq:scalarfluxeqn}
\end{align}
for the tensor and scalar fluxes, respectively. Here TT refers to the
transverse-traceless projection of the tensor. The existence of a
$-0.5$PN piece of the scalar waveform means that the scalar waveform
must generally be known to $(N+1/2)$-th PN order to find the flux at
$N$th PN order. Lang~\cite{Lang:2014osa} computes this flux to 1PN
order. The result is
\begin{equation}
\frac{dE}{dt} = \dot{E}_{-1}+\dot{E}_0+\dot{E}_{0.5}+\dot{E}_1 \, ,
\label{eq:totalflux}
\end{equation}
where
\begin{subequations}
\begin{align}
\dot{E}_{-1} &= \frac{4}{3}\frac{\mu\eta}{r}\left(\frac{G\alpha m}{r}\right)^3\zeta\mathcal{S}_-^2 \, ,\label{eq:minus1PNflux}\\
\begin{split}
\dot{E}_0 &= \frac{8}{15}\frac{\mu\eta}{r}\left(\frac{G\alpha m}{r}\right)^3\left\{\frac{G\alpha m}{r}\left[-2\frac{\delta m}{m}\zeta \mathcal{S}_+\mathcal{S}_- \right.\right.\\
&\left.\qquad \quad +\left(-23+\eta-10\bar{\gamma}-10\bar{\beta}_++10\frac{\delta m}{m}\bar{\beta}_-\right)\zeta\mathcal{S}_-^2\right] \\
&\qquad +v^2\left[12+6\bar{\gamma}+2\zeta\mathcal{S}_+^2+2\frac{\delta m}{m}\zeta\mathcal{S}_+\mathcal{S}_-+(6-\eta+5\bar{\gamma})\zeta\mathcal{S}_-^2\right. \\
&\left.\qquad \quad -\frac{10}{\bar{\gamma}}\frac{\delta m}{m}\zeta\mathcal{S}_-(\mathcal{S}_+\bar{\beta}_++\mathcal{S}_-\bar{\beta}_-)+\frac{10}{\bar{\gamma}}\zeta\mathcal{S}_-(\mathcal{S}_-\bar{\beta}_++\mathcal{S}_+\bar{\beta}_-)\right]\\
&\qquad +\dot{r}^2\left[-11-\frac{11}{2}\bar{\gamma}+\frac{23}{2}\zeta\mathcal{S}_+^2-8\frac{\delta m}{m}\zeta\mathcal{S}_+\mathcal{S}_-+\left(-\frac{37}{2}+9\eta-10\bar{\gamma}\right)\zeta\mathcal{S}_-^2 \right.\\
&\qquad \quad -\frac{80}{\bar{\gamma}}\zeta\mathcal{S}_+(\mathcal{S}_+\bar{\beta}_++\mathcal{S}_-\bar{\beta}_-)+\frac{30}{\bar{\gamma}}\frac{\delta m}{m}\zeta\mathcal{S}_-(\mathcal{S}_+\bar{\beta}_++\mathcal{S}_-\bar{\beta}_-) \\
&\left.\left.\qquad \quad-\frac{10}{\bar{\gamma}}\zeta\mathcal{S}_-(\mathcal{S}_-\bar{\beta}_++\mathcal{S}_+\bar{\beta}_-)+\frac{120}{\bar{\gamma}^2}\zeta(\mathcal{S}_+\bar{\beta}_++\mathcal{S}_-\bar{\beta}_-)^2\right]\right\} \, ,
\label{eq:0PNflux}
\end{split}
\\
\begin{split}
\dot{E}_{0.5} &= -\frac{16}{9}\frac{\mu\eta}{r}\left(\frac{G\alpha m}{r}\right)^3(\zeta\mathcal{S}_-)^2\left(\mathcal{S}_+^2+2\frac{\delta m}{m}\mathcal{S}_+\mathcal{S}_-+\mathcal{S}_-^2\right)\frac{G\alpha m}{r}\dot{r} \\
&\quad-\frac{16}{3}\frac{\mu \eta}{r}\left(1+\frac{1}{2}\bar{\gamma}\right)\zeta \mathcal{S}_-^2\left\{2\left(\frac{G\alpha m}{r}\right)^3\frac{G\alpha m}{r}\dot{r} \right. \\
&\qquad +\frac{(G\alpha m)^3}{r} \hat{n}^k\int_0^\infty ds\ \left[\frac{G\alpha m}{r^4}\left(\left(3v^2-15\dot{r}^2-2\frac{G\alpha m}{r}\right)\hat{n}^k+6\dot{r}v^k\right)\right]_{\tau-s} \\
&\left.\qquad \quad \times \ln \frac{s}{2R+s}\right\} \, ,
\label{eq:RZflux}
\end{split}
\end{align}
\label{eq:fluxdetails}
\end{subequations}
and $\dot{E}_1$ is given in~\cite{Lang:2014osa}. Here we define $\delta m \equiv m_1-m_2$,
\begin{subequations}
\begin{align}
\mathcal{S}_+ &\equiv \alpha^{-1/2}(1-s_1-s_2) \, ,\\
\bar{\gamma} &\equiv -2\alpha^{-1}\zeta(1-2s_1)(1-2s_2) \, , \\
\bar{\beta}_\pm &= \frac{1}{2}(\bar{\beta}_1 \pm \bar{\beta}_2) \, , \\
\bar{\beta}_1 &\equiv \alpha^{-2}\zeta(1-2s_2)^2(\lambda_1(1-2s_1)+2\zeta s_1') \, , \\
\bar{\beta}_2 &\equiv \alpha^{-2}\zeta(1-2s_1)^2(\lambda_1(1-2s_2)+2\zeta s_2') \, ,\\
\intertext{and}
\lambda_1 &\equiv \frac{(d\omega/d\varphi)_0\zeta}{3+2\omega_0} \, .
\end{align}
\end{subequations}
We also note that the subscript $\tau-s$ in \eqref{eq:RZflux} means that the quantity should evaluated at time $\tau-s$, where $\tau \equiv t-R$ is the retarded time. Equations \eqref{eq:totalflux} and \eqref{eq:fluxdetails} can be used to determine the phase evolution of a
binary, the last step in producing a usable waveform for data analysis
studies.
While future work in this area will certainly involve extending the
current calculation to higher PN order, it may also be interesting to
investigate theories with multiple scalars or a potential.
A derivation of the quadrupole-order flux in tensor-multiscalar
theories, that agrees with Lang's results in the single-scalar limit,
can be found in~\cite{Damour:1992we}.
The current state-of-the-art calculation for compact binaries in the
massive Brans-Dicke theory was performed by Alsing et
al.~\cite{Alsing:2011er} (see
also~\cite{Krause:1994ar,Perivolaropoulos:2009ak}). In the notation
used by Lang, and correcting a mistake in~\cite{Alsing:2011er}, they
found that the lowest-order flux is given by
\begin{equation}
\dot{E} = \frac{4}{3}\frac{\mu\eta}{r}\left(\frac{G\alpha m}{r}\right)^3\zeta\mathcal{S}_-^2\left[\frac{\omega^2-m_s^2}{\omega^2}\right]^{3/2}\Theta(\omega-m_s) \, ,
\label{eq:Alsingflux}
\end{equation}
where $\omega$ is the orbital frequency, $m_s$ is the mass of the
scalar field, and $\Theta$ is the Heaviside function.
In massive Brans-Dicke theory, scalar dipole radiation is emitted only
when $\omega > m_s$. Alsing et al.~continued their calculation to 1PN
order; however, those terms are incomplete and we do not list them
here.
\subsubsection{Numerical relativity simulations}\label{sec:nr_st}
Numerical relativity (the use of numerical simulations to solve
Einstein's equations in full generality in the nonlinear regime) is
the most powerful tool at our disposal to understand strong
gravity. Numerical relativity had a 40-year long gestation
\cite{Sperhake:2014wpa}, and the
main motivation behind its development was the description of
high-energy astrophysical phenomena in the framework of GR. In recent
years the theory has been extended beyond GR and it found unexpected
applications in many other fields, ranging from high-energy physics to
solid-state physics~\cite{Cardoso:2014uka}.
Even within GR, obtaining numerically stable and accurate time
evolutions in the absence of high degrees of symmetry is a daunting
task that requires an understanding of many complex issues, such as
the well-posedness of the evolution system, the construction of
initial data and gauge conditions~\cite{Alcubierre:2008,Baumgarte2010}.
These same questions arise also in all proposed extensions of GR, and
at present they remain unanswered for most of the theories discussed
in this review.
Scalar-tensor theories represent a notable exception, because they can
be formulated in close analogy to GR. As discussed in
Section~\ref{subsec:ST}, the action of scalar-tensor theories in the
Einstein frame is the same as the Einstein-Hilbert action, except for
a minimal coupling with the scalar field in the gravitational sector;
a nonminimal coupling with the scalar field only appears in the matter
sector.
The field equations in the Einstein frame -- Eqs.~\eqref{ST:Einstein},
that we reproduce here for the reader's convenience --
are
\begin{subequations}
\label{ST:Einsteinbis}
\begin{align}
G^\star_{\mu \nu} &=2\left(\partial_\mu\varphi\partial_\nu\varphi-
\frac{1}{2}g^\star_{\mu\nu}\partial_\sigma\varphi\partial^\sigma\varphi\right)-
\frac{1}{2}g^\star_{\mu\nu}V(\varphi)+8\pi T^\star_{\mu\nu}\,,\label{eq:tensoreqnEbis} \\
\Box_{g^\star} \varphi &=-4\pi\alpha(\varphi)T^\star+\frac{1}{4}\frac{dV}{d\varphi}\label{eq:scalareqnEbis}\,.
\end{align}
\end{subequations}
In the Jordan frame the scalar field is minimally coupled to matter,
free particles follow geodesics of the spacetime metric and the
stress-energy tensor $T_{\mu\nu}$ of a given matter source (e.g., a
perfect fluid) has formally the same expression as in GR. In the
Einstein frame the stress-energy tensor is
\begin{equation}
T^{\star~\nu}_{\mu}=A^4(\varphi)T^{~~\nu}_{\mu}\,,
\label{transftmn}
\end{equation}
where $A(\varphi)$ is the conformal factor (see
Section~\ref{subsec:ST}). Therefore, as mentioned above, matter fields
are coupled with $\varphi$ in the Einstein frame. Energy-momentum
conservation in the Jordan frame, $\nabla_{\mu}T^{\mu \alpha}=0$,
translates to the Einstein-frame condition
\begin{equation}
\nabla_{g^\star}^{\mu} T^\star_{\mu\alpha} = \alpha(\varphi)T^\star\partial_\alpha \varphi\,.
\label{eq:evolTE}
\end{equation}
The time evolution of a physical system can then be obtained by
solving Eqs.~(\ref{eq:tensoreqnEbis}), (\ref{eq:scalareqnEbis}) and
(\ref{eq:evolTE}). Except for the addition of a minimally coupled
scalar field and -- when matter is present -- for the nonminimal
coupling of $\varphi$ with the stress-energy tensor in
Eq.~(\ref{transftmn}), this system of equations is identical to the
field equations of GR. This is the reason why scalar-tensor theories
of gravity can be attacked using relatively minor generalizations of
the numerical codes developed for GR. The evolution of the scalar
field $\varphi$ is dictated by Eq.~(\ref{eq:scalareqnEbis}), a wave
equation that manifestly preserves any hyperbolicity properties that
are satisfied when Eqs.~(\ref{eq:tensoreqnEbis}) and (\ref{eq:evolTE})
are formulated as an initial-value problem.
Salgado et al.~\cite{Salgado:2005hx,Salgado:2008xh} showed that a
strongly hyperbolic formulation can be obtained also in the physical
(Jordan) frame. However, the Einstein frame is exceptionally
convenient for applications to vacuum spacetimes. The reason is that
in vacuum ($T^{\star\,\alpha}{}_{\beta}=0$) the evolution equations
\eqref{ST:Einsteinbis} are independent of the coupling function
$A(\varphi)$. Therefore a single numerical evolution represents a
whole class of theories characterized by different functional forms of
$A(\varphi)$ for a given potential $V(\varphi)$. Different choices of $A(\varphi)$
result in different physical predictions (e.g.~in terms of
gravitational waveforms), but all of these predictions can be
calculated by {\em post-processing} data from one and the same
numerical simulation. This would not be possible in the Jordan frame,
where the coupling function explicitly appears in the system of
equations that are numerically evolved in time. At least for vacuum
spacetimes, the Einstein frame allows for a considerable reduction in
the computational cost of exploring different scalar-tensor theories.
Early numerical studies of gravitational systems in scalar-tensor
theory focused on gravitational collapse in spherical symmetry, a 1+1
dimensional problem involving only time and one radial
coordinate. These studies explored dust collapse in Brans-Dicke theory~\cite{Shibata:1994qd,Scheel:1994yn,Scheel:1994yr}, the collapse and
stability of NSs~\cite{Novak:1997hw,Novak:1998rk}, and stellar
core-collapse~\cite{Novak:1999jg} in more general scalar-tensor
theories, with particular focus on the spontaneous scalarization
phenomenon~\cite{Damour:1993hw}. The recent breakthroughs in numerical
relativity have opened up the realm of compact binary simulations in
scalar-tensor theories of gravity. We now summarize the main findings
for BH-BH and NS-NS binaries.
\begin{figure}
\capstart
\centering
\includegraphics[width=0.7\textwidth]{FP_trajectories-eps-converted-to.pdf}
\caption{BH trajectories in the Einstein frame, assuming that the
BHs are in a scalar field bubble. The
upper-left panel corresponds to the GR limit; upper-right and
lower-left panels correspond to different initial amplitudes of the scalar
field. In the lower-right panel the evolution occurs in the
presence of a nonzero quartic potential. [From~\cite{Healy:2011ef}.] }
\label{fig:trajectories}
\end{figure}
\paragraph{Black hole binaries.}
\label{sec:NRblackholes}
For BH-BH binaries, scalar-tensor theories represent conceptually simple
modifications of GR. A downside of this simplicity is that
introducing nontrivial BH binary dynamics in these theories (where by
``nontrivial'' we mean dynamics differing from pure GR) requires
somewhat contrived scenarios.
One obvious solution of the field equations \eqref{ST:Einsteinbis} in
the vacuum case ($T^{\alpha}{}_{\beta}=T^{\star\,\alpha}{}_{\beta}=0$)
is the GR solution for the metric, plus a constant scalar field. This
was realized a long time ago, and led to various no-hair theorems
stating that {\it stationary} BH solutions in Brans-Dicke theory are
the same as in GR (see
e.g.~\cite{Hawking:1972qk,1971ApJ...166L..35T,Chase:1970},
and~\cite{Chrusciel:2012jk} for a review). These results have recently
been extended to Bergmann-Wagoner theories~\cite{Sotiriou:2011dz} (see
Section~\ref{sec:ST_BHs}). Moreover, as discussed in
Section~\ref{sec:pn_st},
the dynamics of BH binaries in scalar-tensor theories was shown to be
indistinguishable from GR at all PN orders in the extreme mass-ratio
limit, and up to 2.5PN order in the equations of motion in PN theory.
This ``generalized no-hair theorem'' relies on the following
assumptions: (1) the spacetime contains no matter, (2) the potential
$V(\varphi)$ vanishes, (3) the scalar-tensor action is truncated at
second order in the derivative expansion, and (4) the metric is
asymptotically flat and the scalar field is asymptotically constant.
Deviations from GR in the radiation from BH binaries can occur if we
violate any of these four assumptions.
The most obvious way to obtain nontrivial dynamics is to violate
hypothesis (1), i.e.~to consider configurations involving matter, such
as NS-NS binaries. Leaving this possibility aside for the moment,
another way out of the no-hair theorems was suggested by Horbatsch \&
Burgess~\cite{Horbatsch:2011ye}: if the scalar field is not
asymptotically stationary, the BHs in a binary could retain scalar
hair~\cite{Jacobson:1999vr} and emit dipole radiation, as long as
their masses are not exactly equal. The introduction of higher-order
derivatives in the action would also violate the hypotheses of the
generalized no-hair theorem, but it would lead to substantially more
complicated equations, whose well-posedness remains unclear at present
(cf.~Section~1 of~\cite{Berti:2013gfa}).
\begin{figure}[t]
\begin{center}
\begin{tabular}{ll}
\includegraphics[width=0.4\textwidth,clip=true]{FP_rMPsi4_22-eps-converted-to.pdf}&
\includegraphics[width=0.5\textwidth,clip=true]{TP_rMPhi22_00-eps-converted-to.pdf}\\
\end{tabular}
\caption{Left: The $l=m=2$ multipole of the complex Weyl scalar
$\Psi_4$ (the dashed line corresponds to the real part of
$\Psi_4$, and the solid line to its modulus). Right: The $l=m=0$
multipole of the so-called breathing mode $r\,M\,\widetilde \Phi_{22}$,
where $r$ is the extraction radius and $M$ is the total mass of
the binary; solid, dashed and dotted lines correspond to
different values of the parameter $\beta_0$ (see Eq.~(\ref{DEalphabeta})).
Merger occurs at $t/M=0$. [From~\cite{Healy:2011ef}.]}
\label{fig:strain}
\end{center}
\end{figure}
Healy et al.~\cite{Healy:2011ef} investigated whether generalized
no-hair theorems carry over to the nonlinear regime, i.e.,~whether the
dynamics of BH binaries during the late inspiral and merger is the
same as in GR. Their results show that the dynamics can differ if the
scalar field evolves: the scalar field triggers
energy loss that leads to a difference in the GW polarizations. In their
evolutions, nontrivial dynamics is triggered by placing the BHs inside
a scalar field ``bubble,'' which in some cases includes a nonvanishing
scalar field potential. As the bubble collapses, the BHs accrete the
scalar field and grow. The increase in mass of the BHs has a
dramatic effect on the binary dynamics.
Figure~\ref{fig:trajectories} shows the BH trajectories (in the
Einstein frame) for the four cases considered in their study (cases A,
B, C and D from top left to bottom right). Case A represents the
binary evolution in GR. Cases B, C and D differ in the initial amplitude of
the scalar field in the bubble, and case D furthermore contains a
nonvanishing potential term: see Table~1 in~\cite{Healy:2011ef} for
details. The left panel of Figure~\ref{fig:strain} shows the $l=m=2$
multipole of the Weyl scalar $\Psi_4$ (roughly speaking, the second
time derivative of the GW signal) for each of the 4 initial
configurations. There are obvious differences between the various time
evolutions of $\Psi_4$, in particular when compared against GR (case
A). The $l=m=0$ multipole of the breathing mode
$\widetilde
\Phi_{22}$ in the Jordan frame is shown for
cases B, C and D in the right panel of Figure~\ref{fig:strain}. Notice
that the inclusion of a potential term (as in case D) introduces longer lived
dynamics in the scalar field mode; see also
Section~\ref{subsec-bh-exotic} for long-term evolutions of the
post-merger phase.
In summary, the study of Healy et al.~\cite{Healy:2011ef} supports the
view that an evolving scalar field is required to bypass the
generalized no-hair theorems for BH binaries. Inhomogeneities in the
initial scalar field configuration could provide such a mechanism.
This particular study considered a BH in a scalar field bubble, but
the conclusions can be carried over to more generic scenarios. For
the effects to be observable, the merging BHs must accrete enough
scalar field to change their masses and modify the binary evolution.
A different mechanism to circumvent the no-hair theorems was
considered by Berti et al.~\cite{Berti:2013gfa}, who relaxed
assumption (4) in the list above by introducing non-asymptotically
flat or constant boundary conditions. Conceptually, the main
motivation for relaxing this assumption comes from cosmological
considerations. Inhomogenous scalar fields have been considered in
cosmological models as an alternative to dark matter~\cite{Sahni:1999qe,Hu:2000ke}, and also as models of supermassive
boson stars~\cite{Macedo:2013qea}.
For scalar-field profiles that vary on a lengthscale much larger than
the BH binary separation, one effectively has a configuration with an
approximately constant scalar-field gradient at large separation from
the binary.
\begin{figure}
\capstart{}
\includegraphics[height=120pt,clip=true]{doublebh_s1m07-eps-converted-to.pdf}
\includegraphics[height=120pt,clip=true]{phi11derivative-eps-converted-to.pdf}
\caption{Numerical results for a 10-orbit inspiral of a nonspinning
BH binary of mass ratio $3:1$ inside a scalar-field gradient of
magnitude $M\sigma = 2 \times 10^{-7}$. Left: Real part of the
spin-weighted spheroidal harmonic components of the Newman-Penrose
scalar $\Psi_4$ extracted at coordinate radius $\tilde{r}=56~M$
for harmonic indices $l=m$ (the imaginary part is identical up to
a phase shift). Right: Time-derivative of the scalar field at the
largest and smallest extraction radii, rescaled by radius and
shifted in time. [From~\cite{Berti:2013gfa}.]}
\label{fig:bbhs2m07}
\end{figure}
Ref.~\cite{Berti:2013gfa} considered the quasi-circular inspiral of a
nonspinning BH binary (with mass ratio $3:1$) in a scalar-field
gradient of magnitude $M\sigma = 2\times 10^{-7}$ perpendicular to the
orbital angular momentum vector. The three lowest multipoles of the
Newman-Penrose scalar $\Psi_4$ extracted from the Einstein metric are
shown in the left panel of Figure~\ref{fig:bbhs2m07}. These multipoles
are effectively indistinguishable from their GR counterparts
(cf.~Figure~5 of~\cite{Berti:2013gfa}). A nonvanishing ``background''
scalar field does, however, lead to the (mostly dipolar) emission of
scalar radiation, which is not present in GR. The time derivative of
the real part of the dipole contribution is shown in the right panel
of Figure~\ref{fig:bbhs2m07}, and it displays the expected $1/r$
fall-off behavior. The oscillation frequency of this dipole mode is
{\em twice} the orbital frequency. At first glance, it may appear
surprising to see an $m=1$ multipole oscillating at twice the orbital
frequency (rather than at the orbital frequency). A simple
calculation, however, reveals that this feature is a consequence of
the interaction of the orbital motion with an $m=1$ background field:
cf.~the discussion around~Eqs.~(36)-(38) of~\cite{Berti:2013gfa}. In
summary, these simulations demonstrate that non-asymptotically flat
boundary conditions (here imposed in the form of a constant scalar-field
gradient) provide a mechanism to generate scalar radiation in BH
inspirals in scalar-tensor theories of gravity, thus circumventing the
no-hair theorems. Unfortunately, there is little hope to observe
scalar radiation of this nature in the near future for cosmologically
realistic values of the scalar-field gradients.
\paragraph{Neutron star binaries.}
\label{sec:NRneutronstars}
\begin{figure}
\centering
\includegraphics[width=5.cm,angle=0]{rhophi_t40000.pdf}
\includegraphics[width=5.cm,angle=0]{rhophi_t70000.pdf}
\vskip -1.6cm
\includegraphics[width=5.cm,angle=0]{rhophi_t90000.pdf}
\includegraphics[width=5.cm,angle=0]{rhophi_t120000.pdf}
\vskip -1.cm
\caption{Snapshots of the scalar field value (color code) and the
stellar surfaces (solid black line) at $t=[1.8,3.1,4.0,5.3]$ms for
a binary system of NSs with gravitational masses $(1.64,\,1.74)
M_\odot$ in isolation, and a scalar-tensor theory with $\beta_0 =
-4.5$. [From~\cite{Barausse:2012da}.]
\label{snapshots}}
\end{figure}
The dynamics of scalar-tensor theories of gravity is different from GR
whenever the spacetime contains matter sources. This is evident in
the Einstein frame, where the stress-energy tensor explicitly depends
on the scalar field [see Eq.~(\ref{transftmn})], but of course it is
also true in the Jordan frame, by virtue of the physical equivalence
of the two frames (cf.~Section~\ref{subsec:ST}).
Violations of the strong equivalence principle mean that
self-gravitating objects follow trajectories that depend on their
internal composition/structure: this is the well known ``Nordtvedt
effect''~\cite{Nordtvedt:1968qr,Roll:1964rd,Eardley:1975}.
For generic (Bergmann-Wagoner) scalar-tensor theories, the
dimensionless coupling $\alpha(\varphi)$ between the scalar field and matter
depends on the local value of the scalar field, and it can be Taylor
expanded as (see Eq.~(\ref{DEalphabeta}))
\begin{equation}
{\alpha}( \varphi )\equiv \frac{d\ln A(\varphi)}{d\varphi}
= \alpha_0
+{\beta_0} (\varphi-\varphi_0)
+ {\cal O} ({\varphi})^2,
\label{coupling}
\end{equation}
where $\alpha_0=1/\sqrt{3+2 \omega_{BD}}$ and $\beta_0$ are
dimensionless constants, and $\varphi_0$ is the asymptotic value of
the scalar field.
As discussed in Section~\ref{sec:NS_ST}, the constant $\alpha_0$ is
severely constrained by Solar System experiments ($\omega_{BD} >
40\,000$, or $\alpha_0<3.5\times
10^{-3}$)~\cite{Will:2014xja}. Observations of binary pulsars imply
${\beta_0}\gtrsim-4.52$, because for sufficiently negative values of
$\beta_0$ ($\beta_0\lesssim-4.35$ for a static NS) spontaneous
scalarization would set in (see Section~\ref{sec:NS_ST}), and the
motion of NSs in binary systems would be affected in ways that are
severely constrained by binary pulsar data (see
Section~\ref{subsec:pulsars}). Note however that these constraints are
somewhat degenerate with the EOS~\cite{Shibata:2013pra}, i.e., a
different EOS changes the value of $\beta_0$ below which dynamical
scalarization appears.
\begin{figure}
\capstart
\centering
\includegraphics[height=3.4cm,width=5.77cm,angle=0]{separation-eps-converted-to.pdf}\\
\vskip 0.028cm
\hskip -0.25 cm\includegraphics[height=3.4cm,width=6.cm,angle=0]{c22-eps-converted-to.pdf}
\caption{The separation and the dominant mode of the $\Psi_4$
curvature scalar (which encodes the detector's response to spin-2
GWs) for a binary system of NSs with gravitational masses $\{
1.64, 1.74\} M_\odot$ in isolation, and for different values of
$\beta_0$. [From~\cite{Barausse:2012da}.]
\label{fig:separation_gw}}
\end{figure}
Recently, Refs.~\cite{Barausse:2012da} (using fully relativistic
numerical simulations, performed in the Einstein frame)
and~\cite{Palenzuela:2013hsa} (using semi-analytical arguments)
discovered a phenomenon similar to spontaneous scalarization in the
late stages of the evolution of NS-NS binaries: ``dynamical
scalarization.'' Their results were independently confirmed by Shibata
et al.~\cite{Shibata:2013pra}. Even in cases in which the individual
NSs would \textit{not} spontaneously scalarize in isolation, the
scalar field inside each star grows when the binary separation
decreases to about $50-60~{\rm km}$
(cf. e.g. Fig.~\ref{snapshots}). This growth has a strong effect on
the binary dynamics, and produces an earlier plunge than in GR
(cf.~the upper panel of Figure~\ref{fig:separation_gw}). The plunge is
followed by the formation of a rotating bar-like matter configuration,
which sheds angular momentum in GWs before collapsing to a BH. The
resulting gravitational waveforms are significantly different from GR
at frequencies $\sim 500-600~{\rm Hz}$, as shown in the lower panel of
Figure~\ref{fig:separation_gw}, as well as in Fig.~\ref{fig:gw2} for a
different system. Deviations at even lower frequencies are also
possible for certain binary systems and theory
parameters~\cite{Palenzuela:2013hsa,Sampson:2014qqa}. Therefore, the
effects of dynamical scalarization are in principle detectable (at
least in some cases) with Advanced LIGO, Advanced VIRGO and KAGRA, for
values of the coupling parameters $\omega_0$ and $\beta$ that are
still allowed by all existing Solar System and binary pulsar tests, as
recently shown in~\cite{Sampson:2014qqa}. Deviations away from GR may
also be observable in the electromagnetic signal (driven by
magnetosphere interactions prior to merger) from binaries of
magnetized NSs~\cite{Ponce:2014hha}. While these deviations are
subtle, they might provide a way in which measurements of
electromagnetic counterparts to GW sources can increase the confidence
with which GR will be confirmed (or ruled out) by GW observations.
\begin{figure}
\centering
\includegraphics[width=0.6\textwidth,angle=0]{c22_m14.pdf}
\caption{The dominant mode of the $\Psi_4$ scalar, for $\beta_0=-4.5$
and an equal-mass binary with gravitational masses of $1.51 M_\odot$
in isolation. Note that neither star undergoes spontaneous
scalarization in isolation, but the GW signal is still different
from GR in the late inspiral/plunge because of the dynamical
scalarization of the system.
\label{fig:gw2} [Adapted from~\cite{Barausse:2012da}.]}
\end{figure}
\subsection{$f(R)$ theories}
Compact binaries in $f(R)$ gravity have been studied by considering
perturbative corrections to the Einstein-Hilbert action
(i.e.~$f(R)=R+aR^2$ with $aR\ll1$) and linear perturbations of
Minkowski spacetime. Berry and Gair~\cite{Berry:2011pb} used this
approach to compute the stress-energy pseudotensor and the parameters
of the PPN expansion. Within the same framework, but exploiting the
equivalence between $f(R)$ gravity and scalar-tensor theories, Naf and
Jetzer~\cite{Naf:2010zy} studied corrections to the periastron
precession for compact binaries, and in a follow-up
work~\cite{Naf:2011za} they also computed corrections to the GW flux
formula up to $\mathcal{O}((aR)^2)$. The flux formula derived
in~\cite{Naf:2011za} predicts that the binary would produce monopole
and dipole radiation in addition to the ordinary quadrupolar
radiation;
these contributions are expected to dominate the non-GR part of the
flux, because they enter at the lowest orders in a PN expansion.
De Laurentis and Capozziello~\cite{DeLaurentis:2011tp} derived a flux
formula at $\mathcal{O}(aR)$ in a similar perturbative expansion (see
also~\cite{DeLaurentis:2013zv}).
At this order, gravitational radiation does not contain monopole or
dipole contributions, but the quadrupole contribution has a correction
linear in $f^{\prime\prime} = 2a$.
To the best of our knowledge, no calculation of the sensitivities in
the context of $f(R)$ gravity has been performed yet. The results
should be qualitatively similar to those in scalar-tensor theory, due
to the equivalence between the two formulations.
At the moment of writing there are no numerical investigations of
compact binaries in $f(R)$ gravity, but this is not due to pathologies
in the theory. In fact, $f(R)$ gravity is equivalent to a special
scalar-tensor theory, and as such it inherits the well-posedness
properties of scalar-tensor theories (see
Section~\ref{sec:nr_st}). Preliminary work on the feasibility of
numerical relativity simulations in $f(R)$ gravity can be found
in~\cite{Paschalidis:2011ww}.
\subsection{Quadratic gravity}
\label{sec:binary-quadratic-gravity}
\newcommand{\PPE}{{\mbox{\tiny PPE}}} The compact binary problem in
quadratic gravity theories has been studied
in~\cite{Yagi:2011xp,Yagi:2013mbt}. In~\cite{Yagi:2011xp}, both
parity-even and parity-odd theories were studied, for quasicircular
orbits consisting of objects with yet-undetermined scalar monopole
moments (parity-even theories) or scalar dipole moments (parity-odd
theories). At that time, moments were only known for BHs, along with
the result (see Section~\ref{sec:NSs/quadratic-gravity}) that NSs have
no $1/r$ scalar hair in EdGB or dCS gravity. In~\cite{Yagi:2013mbt},
the authors focused on dCS. They first constructed numerical solutions
for NSs to second order in rotation (these solutions include the
leading dipole piece of the scalar field solution; see
Section~\ref{sec:NSs/quadratic-gravity} and specifically
Figure~\ref{fig:NSsCS}). With these dipole moments in hand, they were
then able to study eccentric binaries consisting of either BHs or NSs.
There are four dominant physical corrections that arise in the compact
binary problem in quadratic gravity theories.
\begin{enumerate}
\item The scalar field solutions sourced by both compact objects
interact with each other, just as electric charges or magnetic
dipoles interact through the electromagnetic field (scalar pole-pole
interaction). This modifies the binding energy of the binary and
hence the Kepler relation (orbital frequency as a function of
separation).
\item All metric multipole moments are shifted. The mass monopole
shift and mass-current dipole shift are unobservable, i.e.~these
shifts are absorbed back into the definition of physical mass and
spin angular momentum. However, higher moments' shifts can not be
absorbed and so they affect the motion. These also correct the
binding energy, the Kepler relation, and cause additional
precession.
\item The scalar field is dynamical, and sourced by a configuration of
scalar monopoles or dipoles (plus higher moments) which are
orbiting. Thus, there may be a time-varying scalar dipole or
quadrupole, which sources scalar radiation. This radiation carries
away energy and thus the system inspirals at a different rate.
\item Finally, the GWs emitted by the system are also corrected. The
change in the gravitational waveform leads to another change in the
energy flux, and thus an additional correction to the inspiral rate.
\end{enumerate}
Not all of the effects listed above can be directly physically observed. The
three primary observables relevant to the compact binary problem are:
\begin{enumerate}
\item[1.] The correction to the precession of pericenter, $\delta\langle\dot{\omega}\rangle$.
\item[2.] The correction to the orbital decay $\delta\dot{P}_{b}$, or equivalently, the
correction to the energy flux $\delta\dot{E}$.
\item[3.] The modification to the gravitational waveform, which can be
parametrized via the parametrized post-Einsteinian (PPE) parameters
$(\alpha_{\PPE},a_{\PPE})$ for
the amplitude and $(\beta_{\PPE},b_{\PPE})$ for the phase of the
waveform (see Section~\ref{sec:paramtests} for more on the PPE
parameterization).
\end{enumerate}
The first two of these corrections are observables for pulsar timing,
and the third is the observable for GW detection. As the authors
of~\cite{Yagi:2011xp} only considered quasicircular orbits, they did
not compute $\delta\langle\dot{\omega}\rangle$.
Almost the entire compact binary problem can be recovered by modeling
compact objects as point particles with scalar hair. This amounts to
replacing the scalar field's source term (in its equation of motion)
with an \emph{effective} source term $\tau_{\mathrm{eff}}$, which recovers the correct
far-field solution~\cite{Yagi:2011xp}. There are some effects which
are not captured, but they were shown to be subdominant. The
effective equation of motion for the scalar field $\phi$ is then
\begin{equation}
\square\phi = \tau_{\mathrm{eff}} =
-4\pi q_{1}\delta^{(3)}(\mathbf{x}-\mathbf{x}_{1}) + (1 \leftrightarrow 2)
\end{equation}
when the bodies have scalar monopole charges $q_{A}$ $(A=1\,,2)$, as
in the case of EdGB or other parity-even theories; or
\begin{equation}
\square\phi = +4\pi
\mu^{i}_{1}\partial_{i}\delta^{(3)}(\mathbf{x}-\mathbf{x}_{1}) + (1 \leftrightarrow 2)
\end{equation}
when the bodies have scalar dipole hair $\mu_{A}^{i}$ (the
generalization to general $\ell$-pole is covered
in~\cite{Stein:2013wza}). The quantities $q_{A}$ or $\mu_{A}^{i}$ are
found from the strong-field matching procedure, and are proportional
to the small parameters $\alpha_i$ defined in
Eq.~\eqref{eq:quadratic_expansion}.
Thus, all of the above effects depend on powers of the $\alpha_i$'s,
which are the physical parameters that can be constrained from
observations.
The scalar pole-pole interaction, correction (i), can be computed by
``integrating out'' the scalar field at the level of the action. This
gives, for the case of dCS (and correcting a sign error
in~\cite{Yagi:2013mbt}), an interaction potential
\begin{equation}
\label{eq:U-int-dCS}
U_\mathrm{int}
= 4\pi \frac{1}{r_{12}^{3}} \left[ 3(\mu_{1}\cdot n_{12})
(\mu_{2}\cdot n_{12}) - (\mu_{1}\cdot\mu_{2}) \right] \,,
\end{equation}
where $r_{12}$ is the distance between the two bodies,
and $n^i_{12}=(x_1^i-x_2^i)/r_{12}$.
The general pole-pole interaction (again with a sign error) is given
in~\cite{Stein:2013wza}. The dependence on $r_{12}$ is always as
$r_{12}^{-1-s-t}$, where $s$ and $t$ are the $\ell$'s of the dominant
scalar moments of the two bodies, e.g.~$s=0$ for monopole, $s=1$ for
dipole, etcetera. Notice that for the case of a BH-BH binary in EdGB,
where $s=t=0$, the shift in binding energy is $1/r$, just as in the
Kepler binding energy, and so this effect is completely absorbed by
redefining the gravitational constant $G$.
The corrected metric multipole moments [correction (ii)] come from the
strong-field matching calculation.
Computing corrections (iii) and (iv) now requires the far-field
radiative parts of the scalar and the metric, as sourced by the
dynamics of the binary. The far-field solution comes from the
post-Minkowskian expansion of the retarded Green function for
$\square$, e.g.
\begin{align}
\label{eq:vartheta-FZ}
\phi^{\rm FZ} =&\, \frac{1}{r} \sum_{m} \frac{1}{m!}
\frac{\partial^m}{\partial t^m} \int_{\mathcal{M}}
\frac{-\tau'_{\mathrm{eff}}}{4\pi} ({n}_{j} \; {x}'^{j})^m d^3x' \,,
\end{align}
where $\mathcal{M}$ denotes a $t-r=\text{constant}$ hypersurface.
There is a similar expression for the GW correction. The
orbit-averaged energy flux then comes from inserting $\phi^{FZ}$ into
the stress-energy tensor $T_{\mu\nu}$ and evaluating
\begin{align}
\label{eq:Edotdefinition}
\dot{E}^{(\varphi)} = \lim_{r\to\infty} \int_{S^2_r}
\left<T^{(\varphi)}_{ti} n^i \right>_{\omega} r^2 d\Omega
\end{align}
on a 2-sphere at $r\to\infty$ (here $\langle\cdot\rangle_{\omega}$
denotes the orbit-average operation). Again, there is a similar
expression for the corrected energy flux from GWs.
As seen in Eq.~\eqref{eq:vartheta-FZ}, the radiative moments of
$\phi$ come directly from time derivatives of the multipole
moments of $\tau$, and we must study which moment will dominate. The
leading-in-$\alpha_i$ part of these corrections can be computed from the
leading PN motion of the binary, i.e.~simply Keplerian motion.
For a system containing a BH in EdGB, scalar dipole radiation will
dominate, and is a pre-Newtonian effect (as in scalar-tensor theory,
it contains fewer powers of $v^{2}$ than the GR quadrupolar energy
flux). For circular orbits, this is~\cite{Yagi:2011xp}
\begin{equation}
\label{eq:Edot-vartheta-EdGB}
\delta \dot{E}^{(\phi)} = -\frac{4\pi}{3}{1\over m^4}(m_2 q_1 -m_1 q_2)^2 v^8
\end{equation}
For either BH or NS systems in dCS, the variation of the scalar dipole
moment occurs on the spin-precession timescale, while the variation of
the scalar quadrupole moment occurs on the orbital timescale, so the
latter should dominate. For a binary with semimajor axis $a$ and
eccentricity $e$, this quadrupole contribution is~\cite{Yagi:2013mbt}
\begin{equation}
\label{eq:Edot-vartheta-dCS}
\delta\dot{E}^{(\phi)} = -\frac{5}{768}\frac{\alpha_{4}^{2}}{\pi m^{4}}\left(\frac{m}{a}
\right)^{7} \frac{ 2 \Delta_{1}^{2}f_{1}(e) + 2\Delta_{2}^{2}
f_{2}(e)+\Delta_{3}^{2}f_{3(e)}}{(1-e^{2})^{11/2}}\,,
\end{equation}
where $f_{i}(e)=1+\ldots$ are $\mathcal{O}(1)$ polynomials of degree
6, $\Delta_{i}$ is an $\mathcal{O}(1)$ vector which depends on the
bodies' scalar dipole moments (and hence their spin vectors), and
where we have accounted for a difference in the convention for
$\alpha_i$ between~\cite{Yagi:2013mbt} and
Eq.~\eqref{eq:action_quadratic}.
Eqs.~\eqref{eq:Edot-vartheta-EdGB} and~\eqref{eq:Edot-vartheta-dCS}
are correction (iii) listed above. In EdGB, since correction (iii) was
pre-Newtonian, the GW effect (iv) is subdominant, so it was not computed
in~\cite{Yagi:2011xp}. However, in dCS, correction (iv) is of the same
PN order, and it is given by~\cite{Yagi:2013mbt}
\begin{align}
\label{eq:Edot-h-dCS}
\delta\dot{E}^{(\mathfrak{h})} =
-\frac{15}{16} \frac{\alpha_{4}^{2}}{\pi m^4}
\eta\left( \frac{m}{a} \right)^{7}
\chi_{1}\chi_{2}\bar{\mu}_{1}\bar{\mu}_{2}C_{1}^{3}C_{2}^{3}
\frac{g_{1}(e)\hat{S}_{1}^{x}\hat{S}_{2}^{x} + g_{2}(e)\hat{S}_{1}^{y}\hat{S}_{2}^{y}
+ 2 g_{3}(e)\hat{S}_{1}^{z}\hat{S}_{2}^{z} }{(1-e^{2})^{11/2}}\,,
\end{align}
where $\eta=\mu/m$ is the symmetric mass ratio, $g_{i}(e)=1+\ldots$
are $\mathcal{O}(1)$ polynomials of degree 6, $\chi_{A}$ is the
dimensionless spin angular momentum of body $A$, $\bar{\mu}_{A}$ is the
dimensionless magnitude of the scalar dipole moment, $C_{A}$ is the
dimensionless compactness of body $A$, $\hat{S}^{i}_{A}$ is the
normalized spin vector of body $A$, and $z$ is the direction orthogonal to the orbital plane.
We now turn to the observable signatures of the above effects.
A detailed pulsar timing model does not exist, but it is still
possible to compute the averaged additional precession of pericenter,
$\delta\langle\dot{\omega}\rangle$. The standard way to do this
averaged calculation is with the Gauss perturbation
method~\cite{will1993theory,Yagi:2013mbt}. First, the perturbing
acceleration is decomposed into components in an orthonormal frame
which co-rotates with the binary, with two axes aligned with $n_{12}$
and the unit angular momentum vector $\hat{L}$. The components are then inserted into standard formulae
which are averaged over the orbital phase of the
binary. Because~\cite{Yagi:2011xp} focused on quasi-circular orbits,
this result was only computed in~\cite{Yagi:2013mbt}. The additional
precession arising from the effect (i) (the scalar pole-pole
interaction) is
\begin{align}
\delta\langle\dot{\omega}\rangle_\phi &= \frac{75}{256}
\frac{1}{\mu}\frac{\alpha_{4}^{2}}{\pi m^4}
\frac{\chi_{1}\chi_{2}}{(1-e^{2})^{2}} C_1^3 C_2^3 \bar{\mu}_1
\bar{\mu}_2 \left( \frac{m}{a} \right)^{7/2}
\left\{
\frac{1}{2}\left(\hat{S}_{1,x}\hat{S}_{2,x}
+\hat{S}_{1,y}\hat{S}_{2,y}\right)
-\hat{S}_{1,z}\hat{S}_{2,z} \right.\nonumber\\
&\qquad\left.{}-\cot\iota \left[
\hat{S}_{1,z}\left(
\hat{S}_{2,x}\sin\omega + \hat{S}_{2,y}\cos\omega
\right)
\right]
\right\}+(1\leftrightarrow 2)\,,
\end{align}
where $\iota$ is the orbital inclination.
The same calculation can be repeated for the perturbing acceleration
arising from effect (ii), the shift in the metric quadrupole moment of
each body, $Q_1$, $Q_2$. This latter effect is estimated to be more important than
$\delta\langle\dot{\omega}\rangle_{\phi}$, and it is given by
\begin{multline}
\delta\langle \dot\omega\rangle_h =
\frac{3}{a^{7/2}\sqrt{m}(1-e^{2})^2}
Q_{1} \Big[ -1 + \frac{3}{2}\left(
\hat{S}_{1,x}^{2}+\hat{S}_{1,y}^{2} \right)\\
-\hat{S}_{1,z}\cot\iota \left( \hat{S}_{1,x}\sin\omega +
\hat{S}_{1,y}\cos\omega \right)
\Big] +(1\leftrightarrow 2)\,.
\end{multline}
The rate of a binary pulsar's pericenter precession, $\dot{\omega}$,
is measured with much more precision than the rate of orbital decay,
$\dot{P}_{b}$. Thus, observable (i) is much better for placing
constraints than observable (ii). Regardless, it is algebraically
straightforward to combine the Kepler binding energy, the shift in the
binding energy [e.g.~Eq.~\eqref{eq:U-int-dCS}], and the shift in the
energy flux
[e.g.~Eq.~\eqref{eq:Edot-vartheta-EdGB},~\eqref{eq:Edot-vartheta-dCS},
or~\eqref{eq:Edot-h-dCS}] to find the leading-order in $\alpha_i$
correction to $\delta\dot{P}_{b}$. This has been computed for a
circular BH-BH binary in dCS in~\cite{Yagi:2012vf} (this is not
sufficient for pulsar timing constraints, but the calculation is very
similar), giving $\dot{f} = \dot{f}_{GR} ( 1+ \delta C u^4 )$, where
$u\equiv(\pi m f)^{1/3}$, $f$ is the GW frequency (twice the orbital
frequency) and
\begin{align}
\label{eq:dCS-delta-C}
\delta C ={}& \frac{313345}{1107456} \frac{\alpha_{4}^{2}}{\pi m^{4}} \frac{m^2}{m_1^2} \chi_1^2 \left[ 1 - \frac{186607}{62669} \left( \hat{\bm{S}}_1 \cdot \hat{\bm{L}} \right)^2 \right] \nonumber \\
& {}+ \frac{99625}{316416} \frac{\alpha_{4}^{2}}{\pi m^{4}} \frac{\chi_1 \chi_2}{\eta} \left[ \left( \hat{\bm{S}}_1 \cdot \hat{\bm{S}}_2 \right) - \frac{8327}{3985} \left( \hat{\bm{S}}_1 \cdot \hat{\bm{L}} \right) \left( \hat{\bm{S}}_2 \cdot \hat{\bm{L}} \right) \right] + (1 \leftrightarrow 2)\,.
\end{align}
Finally, we come to observable (iii), which is the shift in the
gravitational waveform, relevant to ground-based detectors such as
Advanced LIGO. By the time a binary gets into the frequency band
relevant to these detectors, it is assumed that most eccentricity will
have been damped out, and so for quadratic theories these
calculations have only been done for quasi-circular orbits (see
Section~\ref{sec:paramtests}). Since the phase of a gravitational
waveform is measured with much more precision than the amplitude, most
of the attention has been paid to the phase. This is parameterized in
the PPE formalism in terms of $\beta_{\PPE}$ and $b_{\PPE}$ via
\begin{equation}
\label{eq:PPE-phase}
\Psi_{\PPE} = \Psi_{\rm GR} + \beta_{\PPE} (\pi \mathcal{M} f)^{b_{\PPE}}\,,
\end{equation}
where the chirp mass is $\mathcal{M}\equiv m\eta^{3/5}$.
For even-parity theories, the dominant physical effect comes from
(iii), the dipolar scalar radiation. This was computed for BH-BH binaries
in~\cite{Yagi:2011xp} as
\begin{equation}
\beta_\PPE = - \frac{5}{7168} \frac{\alpha_{3}^{2}}{\pi m^{4}} \frac{\delta
m^2}{m^2} \eta^{-18/5}\,,\qquad
b_{\PPE} = -7/3\,.
\end{equation}
Meanwhile, effects (i)-(iv) all contribute for the dCS
calculation. In~\cite{Yagi:2012vf}, this was computed for circular
BH-BH binaries as
\begin{equation}
\beta_{\PPE} = - \frac{15}{64} \; \delta C \; \eta^{-4/5}, \qquad
b_{\PPE} = -\frac{1}{3}\,,
\end{equation}
where $\delta C$ was given in Eq.~\eqref{eq:dCS-delta-C}.
\subsection{Lorentz-violating theories}
\label{lv-bins}
As discussed in Section~\ref{sec:sensitivities}, in Lorentz-violating
theories strongly gravitating objects are characterized by
``sensitivities'' (or \AE ther or khronon ``charges''). As a result,
the motion of these objects does not follow geodesics of the
background geometry, but rather depends on the numerical values of the
sensitivities, and thus ultimately on the object's nature. This means
that the universality of free fall and the strong equivalence
principle are violated in these theories. More specifically, the
sensitivities enter the equations of motion already at Newtonian
level, where the acceleration of the body $A$ in a binary system is
given by
\begin{equation}
\label{Newtonian-a}
\dot{v}_A^i= -\frac{G_N \tilde{m}_B \hat n_{AB}^i}{(1+\sigma_A) r_{AB}^2}\,,
\end{equation}
with $r_{AB}=|\boldsymbol{x}_A-\boldsymbol{x}_B|$ and $\hat
n_{AB}^i=(x^i_A-x^i_B)/r_{AB}$, and $\sigma_A$ the sensitivity parameter of
body $A$ (see Eq.~(\ref{sensitiv_lv})).
This expression can be re-written
as~\cite{Foster:2007gr}
\begin{equation}
\label{eq:New_active}
\dot{v}_A^i= -\frac{{\cal G} {m}_B \hat n_{AB}^i}{r_{AB}^2}\,,
\end{equation}
where one defines the {\textit{active}} gravitational masses
\begin{equation}
\label{active-mass}
m_B\equiv\tilde{m}_B (1+\sigma_B)
\end{equation}
and the ``effective'' gravitational constant
\begin{equation}
\label{calG}
{\cal G}\equiv\frac{G_N}{(1+\sigma_A) (1+\sigma_B)}\,.
\end{equation}
Similarly, the sensitivities appear in the equations of motion at
higher PN orders in the conservative dynamics~\cite{Foster:2007gr}.
The sensitivities enter also in the dissipative sector of the motion
of binary systems, causing both the emission of dipolar fluxes, as
well as modifications of the quadrupolar emission of GWs that takes
place already in GR. More specifically, the most relevant quantity for
binary systems (and in particular for binary pulsars) is the rate of
change of the orbital period. For systems whose orbital dynamics is
determined by Eq.~\eqref{eq:New_active}, denoting the orbital period
by $P_{b}$ and the binary's binding energy by $E_{b}$, this quantity
can be expressed as
\begin{equation}
\label{Pdot-eq}
\frac{\dot{P_{b}}}{P_{b}} = - \frac{3}{2} \frac{\dot{E}_{b}}{E_{b}}\,,
\end{equation}
which can be further manipulated by writing the binding energy's
rate of change in terms of the total flux of energy ${\cal{F}}$
carried away by GWs, i.e.~$\dot{E}_{b} = - {\cal{F}}\,$.
This flux can be calculated explicitly from the sensitivities and the
binary's orbital parameters, yielding~\cite{Yagi:2013qpa,Yagi:2013ava}
\begin{equation}
\frac{\dot{P}_{b}}{P_{b}} = - \frac{192 \pi}{5} \left(\frac{2 \pi G m}{P_{b}}\right)^{5/3}
\frac{\mu}{m\,P_{b}} \left<\mathcal{A}\right>\,,
\label{flux-AE}
\end{equation}
where as usual $\mu = m_{1} m_{2}/m$ is the reduced mass, $m = m_{1} +
m_{2}$ is the total mass, and we have defined
\begin{align}
\label{A-AE}
\left<\mathcal{A}\right> &=
\frac{5}{32} \left(s_1 - s_2\right)^2 \mathcal{A}_{4} \left( \frac{P_b}{2 \pi G m} \right)^{2/3}
\nonumber \\
&+ \left[\left(1 - s_{1}\right) \left(1 - s_{2}\right)\right]^{2/3}
\left( \mathcal{A}_1 + \mathcal{S} \mathcal{A}_2 + \mathcal{S}^{2} \mathcal{A}_3 \right)
\nonumber \\
&+ {\cal{O}}(1/c^{2})\,.
\end{align}
Also, $s_{A}=\sigma_A/(1+\sigma_A)$ are rescaled sensitivities,
${\cal{S}} = m_{1} s_{2}/m + m_{2} s_{1}/m$ and
$(\mathcal{A}_{1},\mathcal{A}_{2},\mathcal{A}_{3},\mathcal{A}_{4})$
are functions of the coupling constants [$(c_{+},c_{-})$ in
Einstein-\AE ther theory and $(\beta,\lambda)$ in khronometric
gravity].
Two comments are in order at this stage. First, in the GR limit one
obtains $\mathcal{A}=1$, thus recovering the usual quadrupole
formula. Second, for widely separated systems (such as all observed
binary pulsars) the decay rate of the orbital period is dominated by
the terms appearing at the lowest PN order, i.e.~with the least powers
of $G m/P_{b}$. Therefore, the last term in Eq.~\eqref{A-AE} (the
dipolar emission term) dominates the orbital decay rate unless $s_{1}
- s_{2} \approx 0$. This provides a way to constrain Lorentz
violations in gravity with white dwarf-pulsar systems
(cf.~Section~\ref{sec:pulsar_LV}). Nevertheless, in the case $s_1\approx
s_2$ (relevant for instance for the relativistic double pulsar,
cf.~Section~\ref{sec:pulsar_LV}), even though dipolar emission is
suppressed, the sensitivities still produce changes to the
quadrupole formula of GR (i.e., $\mathcal{A}\neq1$).
Hansen et al.~\cite{Hansen:2014ewa} computed the characteristics of
gravitational radiation from NS binary inspirals using as a starting
point the energy flux described above. The evolution of the orbital
frequency $F$ is related to the energy flux via
\begin{equation}
\label{eq:Fdot-LV}
\dot F(u) = \dot F_\mathrm{GR}(u) [1 + \delta_{\dot{F}}(u) ]\,,
\end{equation}
where $\dot F_\mathrm{GR}$ is the GR prediction, $u \equiv (2 \pi
\mathcal{G M} F)^{1/3}$ and
\begin{equation}
\delta_{\dot{F}}(u) = \frac{7}{4} \eta^{2/5} \dot E_\mathrm{-1PN} u^{-2} + \dot E_\mathrm{0PN}
\end{equation}
is the Lorentz-violating correction to the evolution of $F$. $\dot
E_\mathrm{-1PN}$ and $\dot E_\mathrm{0PN}$ represent the
Lorentz-violating correction to the energy flux due to the dipolar and
quadrupolar radiation, respectively. They are given by
\begin{equation}
\dot E_\mathrm{-1PN} = -\frac{5}{56} \mathcal{G} (s_1 - s_2)^2 \mathcal{A}_4\,,
\quad
\dot E_\mathrm{0PN} = \mathcal{G} (\mathcal{A}_1 + \mathcal{S} \mathcal{A}_2 + \mathcal{S}^2 \mathcal{A}_3)-1\,.
\end{equation}
From Eq.~\eqref{eq:Fdot-LV} one can calculate the gravitational
waveform in Fourier space using the stationary-phase approximation. In
particular, the phase is given by
\begin{equation}
\Psi = \Psi_\mathrm{GR} - \frac{3}{128} u^{-5} \left[ \dot E_\mathrm{-1PN} \eta^{2/5} u^{-2} + \dot E_\mathrm{0PN}
+ \mathcal{O}(c^{-2}) \right]\,.
\end{equation}
\subsection{Massive gravity}
Radiation from binary pulsars in the inspiral phase was studied for
the Cubic Galileon in~\cite{deRham:2012fw} and for the general case
in~\cite{deRham:2012fg}. The calculation was done by approximating the
time dependence of the source as small: $T=T_0 + \delta T$, where $T_0
= -M\delta^3(\vec{x})$ and $\delta T(\vec{x},t)$ carries the time
dependence of the inspiraling pulsars. Then upon splitting $\pi=\pi_0
+ \phi$, the background profile $\pi_0$ sourced by $T_0$ is
responsible for Vainshtein screening. The radiation in the Galileon
$\phi$ sourced by $\delta T$ in the background of $\pi_0$ was then
computed using the effective action techniques proposed
by~\cite{Goldberger:2004jt}.
For the Cubic Galileon, the dominant channel is the quadrupole
($\ell=2$)\footnote{Because the Galileon is a scalar mode, there is also
monopole radiation. However the monopole is suppressed relative to
the quadrupole order effect because the monopole enters as a
relativistic correction: $P^{\ell =
0}_{\rm cubic}/P^{\ell=2}_{\rm cubic}\sim v \sim 10^{-3}$,
see~\cite{deRham:2012fw}.}. The power radiated is given by
\begin{equation}
\frac{P^{\ell =2}_{\rm cubic}}{P^{\ell =2}_{\rm GR}} \sim v^{-1} (\Omega r_V)^{-3/2},
\end{equation}
where $v$ is the velocity of the pulsar, $\Omega$ its orbital angular
velocity and $r_V$ is the Vainshtein radius (see
Section~\ref{subsec:quadratic}). Using parameters from the
Hulse-Taylor pulsar as a fiducial example~\cite{Weisberg:2004hi}
yields $P_{\rm cubic}/P_{\rm GR} \sim10^{-7}$, well below the
observational precision $\sigma\sim10^{-3}$. It is interesting to
note, however, that the time dependence in the system makes the
Vainshtein screening less effective compared to the static case; the
force on the two pulsars is suppressed by $(\bar{r}/r_V)^{3/2}$, where
$\bar{r}$ is the separation.
For the Quartic Galileon the situation is more subtle. For a given
multipole, there is more Vainshtein suppression,
$P^\ell_{\rm quartic}/P^\ell_{\rm GR} \sim v^{-2} (\Omega r_V)^{-2}$. However,
in this case the approximation of a spherically symmetric background
is not good, because higher order multipoles are not suppressed
effectively. More work is needed to understand this case, e.g.~by
taking the time dependence into account in the background.
More recently, Narikawa et al.~\cite{Narikawa:2014fua} reported on the
prospects for GW detection from coalescing compact binaries in some
models of bimetric massive gravity~\cite{Hassan:2011zd}. They find
that, in a certain region of the parameter space, the gravitational
waveform can display large-amplitude modulations induced by the
interference between two modes. The peak amplitude can be up to an
order of magnitude larger than its GR value at a given frequency, and
such frequency depends on the parameters of the theory. By using
Bayesian methods (cf.~Section~\ref{sec:paramestimation}), Narikawa et
al.~evaluate the detectability of these deviations in the waveforms by
an advanced laser interferometer, finding that there is a region of
the parameter space of the bimetric gravity theory where the
deviations can be significant. The detectable region depends on the
specific model, but typically corresponds to a graviton mass
$\mu\gtrsim 10^{-22} {\rm eV}$~\cite{Narikawa:2014fua}. Remarkably,
this value overlaps with the bounds on the graviton mass derived
through the superradiant instability of supermassive Kerr BHs in
massive gravity~\cite{PDG,Brito:2013wya} (see
Section~\ref{sec:superradiance}). It is notable that comparable bounds
could follow from GW observations of \emph{stellar}-mass objects.
\subsection{Open problems}\label{op:binaries}
\begin{itemize}
\item It has been shown that the dynamics of a BH binary system (with
a nonextreme mass ratio) in Bergmann-Wagoner theory coincides with
that of GR up to $2.5$ PN order. Does this result hold at all PN
orders? If it does not, at which PN order does it break down?
\item Can we extend numerical relativity to modified theories of
gravity other than scalar-tensor theory? What is the signature of
nonlinear effects in the late inspiral and merger?
\item How do spontaneous scalarization and dynamical scalarization
generalize to Horndeski theories or tensor-multiscalar theories? Are
there similar nonlinear effects that could produce sensible
deviations from GR in quadratic gravity theories, Lorentz violating
theories or massive gravity?
\end{itemize}
\subsection{Black holes in general relativity}
One of the most remarkable predictions of GR is that regular,
stationary BHs in Einstein-Maxwell theory are extremely simple
objects, being defined by at most three parameters: mass, angular
momentum and electric charge. This was established by a series of
uniqueness theorems due to Hawking, Carter and Robinson
(see~\cite{Bekenstein:1996pn,Carter:1997im,Chrusciel:2012jk,Robinson}
for reviews), which imply that all isolated BHs in Einstein-Maxwell
theory are described by the Kerr-Newman family.
Astrophysical BHs are thought to be neutral to a very good
approximation because of quantum discharge
effects~\cite{Gibbons:1975kk}, electron-positron pair
production~\cite{1969ApJ...157..869G,1975ApJ...196...51R,Blandford:1977ds}
and charge neutralization by astrophysical plasmas.
Therefore the geometry of astrophysical BHs in GR is simply described
by the two-parameter Kerr metric~\cite{Kerr:1963ud}, which in standard
Boyer-Lindquist coordinates reads
\begin{equation}
\begin{aligned}
ds^2={}&-(1-2Mr/\rho^2)dt^2-4aMr\sin^2\theta/\rho^2 dt d\varphi
+\frac{\rho^2}{\Delta}\,dr^2+\rho^2\,d\theta^2\\
&{}+\left(r^2+a^2+2Ma^2r\sin^2\theta/\rho^2\right )\sin^2\theta
d\varphi^2 \,,
\end{aligned}
\label{metricKerrLambda}
\end{equation}
where
$\Delta\equiv r^2+a^2-2Mr$ and $\rho^2\equiv r^2+a^2 \cos^2\theta$.
This metric describes the gravitational field of a spinning BH of mass
$M$ and angular momentum $J=a M$.
The roots of $\Delta$ correspond to the event horizon
($r_+=M+\sqrt{M^2-a^2}$) and the Cauchy horizon
($r_-=M-\sqrt{M^2-a^2}$). The static surface $g_{tt}=0$ defines the
boundary of the ergosphere: $r_{\rm
ergo}=M+\sqrt{M^2-a^2\cos^2\theta}$. The ``angular velocity of
the event horizon'' is
\begin{equation}
\Omega_{\rm H}\equiv a/(r_+^2+a^2)\,.
\label{eq:Omega-H}
\end{equation}
Because of the uniqueness theorem, and because NSs in GR cannot be
more massive than $\sim 3M_\odot$~\cite{Rhoades:1974fn}, any
observation of a compact object with mass larger than $\sim 3M_\odot$
with metric different from the Kerr geometry would inevitably signal a
departure from standard physics (either in the gravitational or in the matter sector).
Therefore tests of strong-field gravity targeting
BH systems aim at verifying the ``Kerr hyphothesis'' in various ways.
Teukolsky~\cite{Teukolsky:2014vca} recently compiled an excellent
review on the discovery of the Kerr metric and the impact of this
discovery in astrophysics. We refer interested readers to Teukolsky's
review and standard textbooks~\cite{Shapiro:1983du,MTB,Frolov:1998wf}
for surveys of our current understanding of BHs in GR; here we
summarize some considerations on the stability and no-hair properties
of GR BHs that should be kept in mind when we
discuss BH solutions in modified theories of gravity.
The key theoretical developments after Kerr's discovery were the
derivation of a separable equation -- the ``Teukolsky master
equation'' -- describing perturbations of scalar, neutrino,
electromagnetic and gravitational fields~\cite{Teukolsky:1972my}; the
use of this master equation to assess the mode stability of the
metric~\cite{Teukolsky:1973ha,Press:1973zz,Teukolsky:1974yv}; and
Whiting's work, that conclusively excluded the possibility of
exponentially growing modes~\cite{Whiting:1988vc}. The free
oscillation modes of Kerr BHs under the boundary conditions of ingoing
waves at the horizon and outgoing waves at infinity (whose frequencies
form the so-called ``quasinormal mode'' (QNM) spectrum~\cite{pressringdown})
were investigated extensively by Leaver~\cite{Leaver:1985ax} and
several other authors. The spectrum consists of an infinite discrete
set of complex-frequency modes (hence ``quasinormal''); the nonzero
imaginary part is due to gravitational radiation damping. QNMs
find important applications in various areas of physics, ranging
from quantum gravity to the gauge-gravity duality
(see~\cite{Kokkotas:1999bd,Nollert:1999ji,Berti:2009kk,Konoplya:2011qq}
for reviews). The Teukolsky equation is not self-adjoint, and
QNMs do not form a complete set. The absence of unstable
modes may be a good enough stability proof for a physicist, but not
for a mathematician: mode stability does not imply linear stability.
A rigorous proof of linear stability was carried out by Kay and Wald
for Schwarzschild BHs~\cite{Kay:1987ax}, but the extension of this
analysis to Kerr is still work in
progress~\cite{Dafermos:2008en,Dafermos:2009uq,Dafermos:2010hb,Dafermos:2014cua,Dafermos:2014jwa},
and there is now evidence for instability in extremal Kerr
BHs~\cite{Aretakis:2012ei} (see
also~\cite{Lucietti:2012sf,Yang:2012pj,Yang:2013uba,Cook:2014cta}). For our
purposes, and with the previous caveats, we will consider the absence
of unstable modes as a physically satisfactory stability criterion.
In the rest of this chapter we will first review the properties of BHs
in various extensions of GR, and then turn to a discussion of possible
ways to verify the Kerr hypothesis.
\subsection{Scalar-tensor theories}
\label{sec:ST_BHs}
\subsubsection{Real scalars and no-hair theorems}
Theoretical studies impose remarkable constraints and limitations on
BH solutions in scalar-tensor theories. No-scalar-hair theorems for
the simplest scalar-tensor theories were proved by various authors
Refs.~\cite{Hawking:1972qk,1971ApJ...166L..35T,Chase:1970,Bekenstein:1995un},
and state that stationary BH solutions in Brans-Dicke theory are the
same as those in GR. In other words, the scalar must be trivial and
the geometry must be described by the Kerr metric.
One way to understand this property is to recall our discussion in
Section~\ref{subsec:ST}. By means of field redefinitions, it is always
possible to reformulate the action of a scalar-tensor theory as the
action~\eqref{STactionE} of a minimally coupled sigma-model,
where matter fields have a nontrivial coupling with the scalar
field. However, in vacuum (and in particular in BH spacetimes) the
matter action can be discarded, and Eq.~\eqref{STactionE}
reduces to the Einstein-Hilbert action with a minimally coupled scalar
field.
Extensions of these uniqueness theorems to multiple scalars and to
more generic scalar-tensor theories have been established more
recently~\cite{Heusler:1995qj,Sotiriou:2011dz}. These results assume
the scalar field to be time-independent, a requirement recently shown
to be unnecessary for any scalar-tensor theory with a {\it real}
scalar~\cite{Graham:2014ina}. The no-scalar-hair theorems have also
been confirmed by numerical studies of gravitational collapse to
nonrotating
BHs~\cite{Scheel:1994yr,Scheel:1994yn,Shibata:1994qd,Harada:1996wt,Novak:1997hw,Kerimo:1998qu}.
In summary, the Kerr family of vacuum BH solutions in GR is also the
most general vacuum solution in a rather general class of
scalar-tensor theories, although some exceptions exist, as we discuss in the next sections.
Furthermore, alternative theories with the same equilibrium solutions as GR have,
in general, different dynamics. The theorems summarized above imply
that Kerr BHs in GR are linearly stable, but they are unstable because
of superradiance in massive scalar-tensor theories (including
minimally coupled massive
scalars)~\cite{Damour:1976kh,Detweiler:1980uk,Zouros:1979iw,Cardoso:2004nk,Shlapentokh-Rothman:2013ysa,Cardoso:2013krh}.
Superradiance extracts energy from rotating BHs, and transfers this
energy to the perturbing field. For a monochromatic wave of frequency
$\omega$, the condition for superradiance
is~\cite{zeldovich1,zeldovich2,Cardoso:2013krh}
\begin{equation}
0<\omega<m\Omega_H\,,\label{super_cond}
\end{equation}
where $m>0$ is the azimuthal harmonic index and the angular velocity
of the BH horizon $\Omega_H$ was defined in Eq.~\eqref{eq:Omega-H}. If
the scalar is massive, superradiance triggers an
instability~\cite{Press:1972zz,Damour:1976kh,Detweiler:1980uk,Zouros:1979iw,Cardoso:2004nk,Shlapentokh-Rothman:2013ysa}:
the ergoregion amplifies the field, and the mass term ``traps it.''
The linear stages of the instability lead to the growth of a
non-spherically symmetric scalar ``cloud'' outside the horizon
[because the mechanism requires nontrivial azimuthal dependence, as
seen from (\ref{super_cond})]. For a single {\it real} scalar field,
this leads to a nonzero quadrupole moment of the cloud resulting in
periodic GW emission. Thus, the end-state is thought to be a Kerr BH
with lower
spin~\cite{Witek:2012tr,Okawa:2014nda,Cardoso:2013krh,Brito:2014wla}.
Note, however, that the instability time scale depends on the scalar
field's mass, and may be of the order of the Hubble time, leading to
what in practice amounts to hairy BH configurations.
\subsubsection{Complex scalars: new hairy rotating black holes}
\label{sec:complscal}
When the scalar is time-dependent, the assumptions behind the no-hair
theorems do not apply. Of course, the backreaction of a generic time-dependent
scalar field will lead to a time-dependent geometry and not
an equilibrium BH state. But for a \textit{complex scalar field}
(which is equivalent to two scalar fields) there is a special type of
time dependence that yields a time-independent stress-energy tensor
and hence is compatible with a stationary metric. This time dependence
is simply a phase evolution, analogous to that of stationary states in
quantum mechanics: $\Psi(t,{\bf x})=e^{-i \omega t}\phi({\bf x})$. As
shown in~\cite{Pena:1997cy}, however, no spherically symmetric BHs
exist even with this time dependence. With the wisdom of hindsight
this is easy to understand. The null generator of the horizon
$\chi=\partial_t$ does not preserve $\Psi$. As such there is scalar
flux through the horizon and hence there can be no static geometry.
The latter argument can be circumvented by introducing
rotation for the BH spacetime and thus making the geometry
axisymmetric. Then, the null generator of the horizon gains an additional term:
$\chi=\partial_t+\Omega_H\partial_\varphi$, where $\Omega_H$ is the
angular velocity of the horizon (given in Eq.~\eqref{eq:Omega-H}), and $\partial_\varphi$ the Killing
vector field which generates the axial symmetry.
We must also introduce an azimuthally dependent phase for
the scalar field, $\Psi(t,\varphi,{\bf x})=e^{-i \omega t}e^{i m
\varphi} \phi({\bf x})$, where $m\in \mathbb{Z}^{\pm}$ since
$\varphi$ is periodic with $\varphi\sim \varphi +2\pi$. Observe that,
again, the azimuthal dependence vanishes in the stress-energy
tensor. Then, $\mathcal{L}_\chi \Psi=0$, as long as
\begin{equation}
\omega=m\Omega_H \ .
\label{cloudcond}
\end{equation}
Thus there is no scalar field flux through the horizon as long
as~\eqref{cloudcond} is obeyed, regardless of the value of $\phi({\bf
x})$ on the horizon. This argument suggests the existence of
asymptotically flat rotating BHs with complex scalar hair. Such
solutions were indeed found in~\cite{Herdeiro:2014goa}.
The ultimate physical reason for the existence of these equilibrium
states -- in the sense that the geometry has an asymptotically
timelike Killing vector field, just like Kerr -- is that GW emission
is halted due to cancellations in the stress-energy tensor, which
becomes independent of the time and azimuthal variables, thus avoiding
GW emission and consequent angular momentum losses.
The condition~\eqref{cloudcond} for the existence of hairy BHs lies
precisely at the threshold of the superradiant
condition~\eqref{super_cond}. This is no accident. A test-field
analysis of the type that leads to condition~\eqref{super_cond}, for a
complex scalar field on the Kerr background, reveals that \textit{real
frequency} bound states are possible precisely in between amplified
modes, which obey the superradiant condition~\eqref{super_cond}, and
decaying modes, which obey $\omega>m\Omega_H$. These are stationary
scalar
clouds~\cite{Hod:2012px,Hod:2013zza,Herdeiro:2014goa,Benone:2014ssa,Herdeiro:2014pka}. The
hairy BHs found in~\cite{Herdeiro:2014goa} can be thought of as
nonlinear realizations of these clouds, when the scalar field becomes
``heavy'' and backreacts (see also~\cite{Herdeiro:2014ima}).
The solutions found in~\cite{Herdeiro:2014goa} correspond to a
five-parameter family of the Einstein-(massive)-Klein-Gordon
theory. Three of the parameters are continuous: the ADM mass $M$, the
ADM angular momentum $J$, and a Noether charge $Q$. The latter is
obtained by integrating the time component of the scalar field
4-current on a spacelike slice and may be regarded as measuring the
amount of scalar hair outside the horizon. In fact, it proves
convenient to introduce a normalized Noether charge $q\equiv
Q/2J$. Then, $q$ is a compact parameter in the full space of
solutions: $0\le q\le 1$. The value $q=0$ corresponds to Kerr BHs,
showing that these hairy BHs are continuously connected to the
standard Kerr family. This is why the solutions
in~\cite{Herdeiro:2014goa} were dubbed ``Kerr BHs with scalar hair.''
The value $q=1$ corresponds to asymptotically flat, rotating boson
stars~\cite{Yoshida:1997qf,Kleihaus:2005me}. These are (horizonless)
gravitating solitons, which are kept in equilibrium by a balance
between the scalar field self-gravity and its wave-like dispersive
nature. Rotating boson stars sustained by a complex, massive field
have $Q=2J$, which justifies the normalization taken. The two
remaining parameters of the solutions found in~\cite{Herdeiro:2014goa}
are discrete: the aforementioned azimuthal harmonic index $m\in
\mathbb{Z}^\pm$ and the node number $n\in \mathbb{N}_0$. The latter
counts the number of zeros of the scalar field radial profile. One may
regard $n=0$ as the fundamental configuration and $n\ge 1$ as excited
states.
The line element and scalar distribution describing Kerr BHs with scalar hair reads:
\begin{align}
\label{ansatz_hairy}
ds^2&=e^{2F_1}\left(\frac{dR^2}{N }+R^2 d\theta^2\right)+e^{2F_2}R^2 \sin^2\theta (d\varphi-W dt)^2-e^{2F_0} N dt^2,
~~~~N\equiv 1-\frac{R_H}{R}, \nonumber \\
\Psi&=\phi(r,\theta)e^{i(m\varphi-\omega t)}.
\end{align}
In this ansatz there are five functions of $(R,\theta)$:
$F_0,F_1,F_2,N,\phi$. To obtain them one numerically solves five
nonlinear, coupled PDEs, with appropriate boundary conditions that
ensure both asymptotic flatness and regularity at the horizon. The
latter requirement actually implies condition~\eqref{cloudcond}. The
parameter $R_H$ is the location of the event horizon in this
coordinate system. We remark that these are \textit{not}
Boyer-Lindquist coordinates in the Kerr limit. In order to write Kerr
in the form~\eqref{ansatz_hairy} one must change the radial coordinate
$r$ in~\eqref{metricKerrLambda} by the transformation
$R=r-{a^2}/{r_H}$, where $r_H=M+\sqrt{M^2-a^2}$ is the event horizon
location in Boyer-Lindquist coordinates.
The parameter and phase space for the solutions with $n=0$, $m=1$ were
discussed in detail in~\cite{Herdeiro:2014goa} and are summarized in
Figure~\ref{hairy-pm}. There is a region of overlap of hairy and Kerr
BHs with the same $(M,\,J)$. In that sense there is nonuniqueness. The
degeneracy seems to be raised, however, by the introduction of $q$: no
two solutions were found with the same $(M,\,J,\,q)$. In the region of
nonuniqueness, hairy BHs have larger entropy than the corresponding
Kerr BHs. As such the former cannot decay into the latter
adiabatically. Also, hairy BHs can violate the Kerr bound: $J\le
M^2$. This violation is not surprising since it is known to occur for
rotating boson stars~\cite{Ryan:1996nk}, and hairy BHs are
continuously connected to boson stars. It is indeed a generic feature
that hairy BHs are more star-like, i.e.~less tightly constrained in
their physical properties than Kerr BHs. This observation also has
implications for possible astrophysical phenomenology of hairy BHs, an
aspect of special relevance for this review. It was observed
in~\cite{Herdeiro:2014goa} that both the quadrupole moment and the
angular frequency at the ISCO can differ significantly for hairy BHs,
as compared to the standard Kerr values. Finally, hairy BHs have a
richer structure of ergo-regions than Kerr, with the occurence of
\textit{ergo-Saturns}, besides ergo-spheres, in a region of parameter
space~\cite{Herdeiro:2014jaa}.
\begin{figure}[h!]
\centering
\includegraphics[height=2.48in]{BH-w-M.pdf}\\
\caption{%
Domain of existence of hairy BHs for $n=0$,
$m=1$ in $M$-$\omega$ space (shaded blue region). The black solid
curve corresponds to extremal Kerr BHs, which obey
$M={1}/(2\Omega_H)$; Kerr BHs exist below it. For $q=0$, the domain
of existence connects to Kerr solutions (dotted blue line). For
$q=1$, $R_H$ vanishes and hairy BHs reduce to boson stars (red solid
line). The final line that delimits the domain of existence of the
hairy BHs (dashed green line) corresponds to extremal BHs, i.e.~with
zero temperature. (Inset) Boson star curves for $m=1,2$. The units
in the axes are normalized to the scalar field mass $\mu$. [Adapted
from~\cite{Herdeiro:2014goa}.]}
\label{hairy-pm}
\end{figure}
In Figure~\ref{functions_plot} we plot the five functions
in~\eqref{ansatz_hairy} for an example of a Kerr BH (left panel, for
which case $\phi=0$) and also for a hairy Kerr BH solution (right
panel), on the equatorial plane $\theta=\pi/2$ and in terms of a
compactified radial coordinate $1-R_H/R$. The behavior observed here
is quite generic. All metric functions are monotonic functions of
$R$. The scalar field profile function is nonzero on the horizon, has
one maximum and tends to zero asymptotically, also a generic behavior
for $n=0$ solutions. A set of ten example solutions (including the two
just mentioned) are available online as a supplement to this
review~\cite{ref:webpage}. The files provide all five metric functions
$(F_0,\,F_1,\,F_2,\,N,\,\phi)$ on a fine grid (see
also~\cite{Herdeiro:2015gia} for a detailed computation of these
solutions).
The stability of these solutions and the formation mechanism of hairy
BHs that deviate significantly from Kerr remain urgent open issues. A
recent analysis suggests that, should these solutions arise from a
superradiant instability of the Kerr metric, the energy-density of the
scalar field would be negligible and the geometry would be well
described by the Kerr solution~\cite{Brito:2014wla}.
\begin{figure}[h!]
\centering
\includegraphics[height=1.78in]{Kerr_functions.pdf}
\includegraphics[height=1.78in]{hairy_functions.pdf}\\
\caption{Left panel: The metric functions in \eqref{ansatz_hairy} for
a Kerr BH in the region of nonuniqueness. We have chosen its mass
and angular momentum to be $M=0.415$, $J=0.172$; this corresponds to
$r_H= 0.066$. Right panel: The metric and
scalar field functions for a hairy BH in the region of
nonuniqueness with the same $M,J$ as the Kerr BH. This hairy BH is
Kerr-like and has $ \omega=0.975$ and $r_H=0.2$. The value of the
scalar field profile function has been multiplied by a factor of
100.}
\label{functions_plot}
\end{figure}
\subsubsection{Evading no-hair theorems in Horndeski/Gauss-Bonnet gravity}
\label{sec:BH_Horndeski}
Hawking's no-hair theorem for stationary BHs in Brans-Dicke
theory~\cite{Hawking:1972qk} was recently extended by Sotiriou and
Faraoni to more general scalar-tensor theories~\cite{Sotiriou:2011dz}.
Hui and Nicolis~\cite{Hui:2012qt}
further extended these proofs to the most general scalar-tensor
theory leading to second-order field equations, i.e.~Horndeski's
theory (introduced in Section~\ref{subsec:horndeski}). Hui and Nicolis
claimed that static, spherically symmetric, asymptotically flat BHs in
vacuum have no hair in Horndeski's theory, provided that the scalar
exhibits shift symmetry -- i.e., symmetry under $\phi\to\phi+{\rm
constant}$. The conclusion follows from the fact that the scalar
field equation can be written as a conservation equation for the
Noether current $J^\mu$ associated with the shift symmetry, namely
\begin{equation}
\nabla_\mu J^\mu=0\,. \label{eq:Noether}
\end{equation}
In a nutshell, the argument is the following. If the scalar field
respects the symmetries of the metric,
\begin{equation}
ds^2=-f(r)dt^2+f(r)^{-1}dr^2+R(r)^2(d\theta^2+\sin^2\theta d\varphi^2)\,,
\end{equation}
then the only nonvanishing component of $J^\mu$ is the radial one,
which gives the value of the invariant $J^\mu J_\mu=(J^r)^2/{f}$. Because
$f=0$ at the horizon, $J^r$ must also vanish there. Then
the conservation equation~\eqref{eq:Noether} implies that $J^r$ must be zero
everywhere. Finally, the last step of the proof is to argue that
$J^r=0$ everywhere implies that the scalar field must be constant, and
therefore the metric must satisfy Einstein's equations in vacuum.
This last step was criticized in~\cite{Sotiriou:2013qea}, where it was
shown that there exists a counterexample where the scalar field has a
nontrivial profile even though $J^r=0$. This happens precisely when
the scalar field is \emph{linearly coupled to the Gauss-Bonnet
invariant} $R_{\rm GB}$ (see Section~\ref{subsec:quadratic}),
i.e.~for a theory with action
\begin{equation}
S=\frac{1}{16\pi}\int\sqrt{-g} d^4x \left[R-2\nabla_a\phi\nabla^a\phi+\alpha \phi R_{\rm GB}^2\right]\,. \label{EdGBaction2}
\end{equation}
This special case of Horndeski's theory is also a special case of the
quadratic gravity theories discussed in Section~\ref{subsec:EdGB}.
BHs in this theory are indeed endowed with a nontrivial scalar profile
(cf.~Section~\ref{sec:BHquadratic}). The scalar charge is not an
independent quantity, but it depends on the BH mass, so these BHs are
said to have ``hair of the second kind,'' but they are nevertheless
different from their Schwarzschild counterparts. These solutions were
discussed in~\cite{Sotiriou:2014pfa}. They are the only vacuum,
spherically symmetric hairy BHs in Horndeski's theory with shift
symmetry for which the scalar field respects the symmetries of the
metric.
If the scalar field is time-dependent, it is possible for BHs to
develop hair both in scalar-tensor theories~\cite{Jacobson:1999vr} and
in Horndeski theories~\cite{Babichev:2013cya}. Furthermore, as
discussed below, BHs in scalar-tensor theories can grow hair in the
presence of matter.
\subsubsection{BHs surrounded by matter}
Isolated BHs in scalar-tensor theories are the same as in GR, but the
situation changes completely in the presence of
matter. Refs.~\cite{Cardoso:2013fwa,Cardoso:2013opa} have investigated
the effects of simple models of accretion disks and halos around BHs
in generic scalar-tensor theories. In these theories the Klein-Gordon
equation on a Kerr BH surrounded by matter takes the form
\begin{equation}
[\square-\mu_{\rm eff}^2]\Psi=0\,,
\end{equation}
where the effective mass $\mu_{\rm eff}$ depends on the specific
scalar-tensor theory, and it is proportional to the trace of the
stress-energy tensor.
Depending on the sign of the scalar coupling, $\mu_{\rm eff}^2$ can
either be positive or negative. In the latter case the system is prone
to a \emph{tachyonic} instability and spontaneously develops a scalar
hair supported by matter. This phenomenon is akin to the spontaneous
scalarization phenomenon in NSs (cf.~Section~\ref{sec:NS_ST}
below). On the other hand, when $\mu_{\rm eff}^2>0$ the scalar field
acquires a real effective mass and can trigger a ``spontaneous
superradiant instability'' similar to the one discussed
previously. The instability is much stronger than in the vacuum case,
because the presence of matter drastically affects the amplification
of scalar waves. In fact, superradiant amplification from spinning BHs
is strongly enhanced when Breit-Wigner resonances
occur~\cite{Cardoso:2013fwa,Cardoso:2013opa}. Possible astrophysical
implications of this amplification have not been investigated yet, but
they may yield phenomenological constraints on the parameter space of
scalar-tensor theories.
The effect of a matter distribution around BHs in scalar-tensor theory
was studied in~\cite{Davis:2014tea} in the context of theories with a
screening mechanism. In scalar-tensor theory, screening occurs when
the conformal factor $A(\varphi)$ and/or the potential $V(\varphi)$
suppress the effect of modified gravity in dense enviroments (such as
those of the Solar System or of our Galaxy), allowing for
modifications on a cosmological scale which are not in conflict with the
bounds from Solar System and binary pulsar tests~\cite{Brax:2012gr}
(see Section~\ref{subsec-cosm-nonlinear}). Using a simple spherically
symmetric model of an accretion disk or a galactic halo, and
neglecting the effect of matter and of the scalar field on the
spacetime metric, Ref.~\cite{Davis:2014tea} shows that matter induces
a nontrivial, spherically symmetric scalar field profile. Their
estimates suggest that the effect of the ``scalar fifth force'' on
test particles should be much smaller than effects due to the
quadrupole emission of GWs, and therefore that it is unlikely to
reveal the presence of a fifth force in this context.
\subsubsection{Stability}
\label{sec:bh-st-stability}
The stability of BHs in scalar-tensor theories is a nontrivial
issue. No-hair theorems do not apply to BH dynamics, which is
different from GR even in the simplest scalar-tensor theories.
To our knowledge, there are very few works on BH stability in
scalar-tensor theories. Kobayashi, Motohashi and Suyama studied the
linear perturbations of static, spherically symmetric BHs in Horndeski
gravity~\cite{Kobayashi:2012kh,Kobayashi:2014wsa} finding a set of
necessary conditions for BH stability. Quite interestingly, these
conditions impose restrictions on the general Horndeski action.
The stability of static, spherically symmetric BHs with respect to
linear odd-parity perturbations was demonstrated in the case of GR
minimally coupled with a scalar field (with an arbitrary
potential)~\cite{Anabalon:2014lea}; their result also applies to
Bergmann-Wagoner theory (see Section~\ref{subsec:ST}). On the other
hand, it has long been known that rotating BHs in GR minimally coupled
with a {\em massive} scalar field are unstable. We will return to this
point in Section~\ref{sec:superradiance}.
\subsection{f(R) theories}
Scalar-tensor theories include $f(R)$ theories as a special case.
The vacuum Kerr spacetime is a solution in $f(R)$ gravity by virtue of
the theorems that apply to generic scalar-tensor
theories~\cite{Sotiriou:2011dz}. Typically $f(R)$ theories propagate
massive degrees of freedom~\cite{Nzioki:2014oaa}. As a consequence,
rotating BHs may be prone to superradiant instabilities: this
possibility was discussed as early as 1985 by Hersh and
Ove~\cite{Hersh:1985hz}. Interestingly, in this case the effective
scalar field is related to the scalar curvature of the metric, which
grows exponentially through superradiance. This suggests that, at
variance with the case of real massive fields previously discussed,
the end-state of superradiant instabilities in $f(R)$ gravity might be
different from a Kerr BH~\cite{Hersh:1985hz}.
\subsection{Quadratic gravity}
\label{sec:BHquadratic}
New BH solutions can be found in theories with quadratic curvature
terms in the action. We first discuss the perturbative approach in
generic quadratic theories before focusing on results specific to the
EdGB and dCS theories.
\subsubsection{Perturbative solutions in the slow-rotation limit}
\label{subsec:BHquadraticPert}
Consider the action~\eqref{eq:action_quadratic}, that includes EdGB
and dCS as special cases. BH solutions in this theory are not known
in full generality (with the exception of EdGB gravity, see
Section~\ref{sec:BH_EdGB} below), but perturbative solutions were
obtained analytically when the coupling functions $f_i$ admit the
expansion~\eqref{eq:quadratic_expansion} and the BH is slowly rotating
(numerical solutions for rapid rotation will be discussed below). The
solution describing a static, spherically symmetric BH is a limiting
case (vanishing rotation) of this family.
Consider the following metric ansatz for the stationary, slow-rotation
limit:
\begin{equation}
\begin{aligned}
ds^2={}&{}-f(r,\theta)dt^2+g(r,\theta)^{-1}dr^2-2\omega(r)\sin^2\theta
dtd\varphi \\
&{}+r^2\Theta(r,\theta)d\theta^2+r^2\sin^2\theta\Phi(r,\theta) d\varphi^2
\,,
\end{aligned}
\label{metric_slowrot_quadratic}
\end{equation}
and let the scalar field have the dependence $\phi=\phi(r,\theta)$.
By solving the field equations one finds the following metric
functions that describe a slowly rotating BH
solution~\cite{Pani:2011gy}:
\begin{align}
f(r,\theta)={}& f^{(0)}+\frac{\alpha_3^2}{4}\left[-\frac{49}{40 M_0^3 r}+
\frac{1}{3 M_0 r^3}+\frac{26}{3 r^4}+\frac{22 M_0}{5 r^5}+\frac{32 M_0^2}{5 r^6}-\frac{80 M_0^3}{3 r^7}\right]\,,\nonumber\\
g(r,\theta)={}& g^{(0)}+\frac{\alpha_3^2}{4}\left[-\frac{49}{40 M^3_0 r}+
\frac{r+M_0}{M_0^2 r^3}+\frac{52}{3 r^4}+\frac{2 M_0}{r^5}+\frac{16 M_0^2}{5 r^6}-\frac{368 M_0^3}{3 r^7}\right]\,,\nonumber\\
\omega(r)={}&\frac{2 a M_0}{r}-\frac{a\alpha_3^2}{4}\left[\frac{3}{5 M_0 r^3}+
\frac{28}{3 r^4}+\frac{6 M_0}{r^5}+\frac{48 M_0^2}{5 r^6}-\frac{80 M_0^3}{3 r^7}\right]\nonumber\\
&{}-a\alpha_4^2\frac{5}{2}\left[\frac{1}{r^4}+\frac{12 M_0}{7r^5}+\frac{27 M_0^2}{10r^6}\right] \,, \nonumber\\
\Theta(r,\theta)={}& 1+\frac{\cos^2\theta}{r^2}a^2\,,\qquad \Phi(r,\theta)=
1+\frac{r+2M_0\sin^2\theta}{r^3}a^2 \,.\nonumber
\end{align}
The scalar field solution is given by
\begin{align}
\phi(r,\theta)={}&{\alpha_3}\left[\frac{1}{2M_0 r}+\frac{1}{2r^2}+\frac{2 M_0}{3 r^3}\right]+
a\alpha_4 \frac{5\cos\theta}{8M_0}\left[\frac{1}{r^2}+\frac{2 M_0}{r^3}+\frac{18 M_0^2}{5r^4}\right]\nonumber\\
&{}-\frac{\alpha_3 a^2}{2}\left[\frac{1}{10r^4}+\frac{1}{5M_0r^3}+\frac{M_0+r}{4M_0^3r^2}+
\cos^2\theta \left(\frac{48M_0^2+21M_0r+7r^2}{5M_0r^5}\right)\right] \,.\nonumber
\end{align}
Here $f^{(0)}\equiv 1-2 M_0/r+2a^2M_0\cos^2\theta/r^3$ and
$g^{(0)}\equiv 1-2M_0/r+a^2\left(r-(r-2M_0)\cos^2\theta\right)/r^3$ are
the expansions of the Kerr metric coefficients of
Eq.~\eqref{metricKerrLambda} up to terms of order ${\cal O}(a^2)$.
Note that, as discussed in Section~\ref{subsec:quadratic}, $\alpha_3$
is the EdGB coupling constant, and $\alpha_4$ is the dCS coupling
constant.
The curvature invariants are regular in the exterior spacetime. The
angular momentum is $J=aM_0$, whereas the physical (ADM) mass of the BH
is~\cite{Mignemi:1992nt,Yunes:2011we}
\begin{equation}
M=M_0 \left[1+\frac{{49}\alpha_3^2}{320M_0^4}\right] \,. \label{EdGB:mass}
\end{equation}
The above solution is accurate up to order
${\cal O}(a^2/M^2,\alpha_i^2/M^4,a\alpha_i^2/M^5)$ in the metric,
and up to order
${\cal O}(a^2/M^2,\alpha_i^2/M^4,a\alpha_i^2/M^5,a^2\alpha_i/M^3)$ in
the scalar field.
Using the second-order in spin corrections obtained
in~\cite{Yagi:2012ya,Ayzenberg:2014aka}, the corrections to the Kerr
quadrupole moment $Q_{\rm Kerr}$ arising in quadratic gravity read
\begin{equation}
Q=Q_{\rm Kerr}\left(1+\frac{4463}{2625}\frac{\alpha_3^2}{M^4}-\frac{201}{448}
\frac{\alpha_4^2}{M^4} \right)\,, \label{quadrupole_quadratic}
\end{equation}
where the quadrupole moment is defined through a large-distance
expansion of the metric as
\begin{equation}
g_{tt}\to -1+\frac{2M}{r}+\frac{\sqrt{3}}{2}\frac{Q}{r^3} Y_{20}(\theta)\,,
\end{equation}
and $Y_{20}$ is a spherical harmonic with $l=2$ and $m=0$.
Constraints on BH solutions are mostly derived by understanding how
matter moves in the vicinity of the BH.
Geodesic motion can be derived from the matter action for a point
particle:
\begin{equation}
S_\text{mat}=-m\int dt\,\sqrt{-\gamma(\phi)g_{\mu\nu}\dot x^\mu\dot x^\nu}\,,
\label{eq:mataction}
\end{equation}
where $m$ is the mass of the particle, and $\gamma(\phi)$ is the
coupling function between the matter and the scalar field. For the
low-energy limit of heterotic string theory, $\gamma=e^{\phi}$.
In the small-coupling limit we have
\begin{equation}
\gamma(\phi)=1+2b\phi+{\cal O}(\phi^2)\,,
\end{equation}
where $b=0$ for minimal coupling and $b=1/2$ in heterotic string
theory. We focus on equatorial motion ($\theta=\pi/2$,
$\dot\theta\equiv0$).
Expanding geodesic quantities to the same order as the metric itself,
we find the following expressions for the ISCO location and the
frequency at the ISCO (both normalized by the physical mass
$M$):
\begin{align}
\frac{R_\text{ISCO}}{M}={}&6-4 \sqrt{\frac{2}{3}} \frac{a}{M_0}-\frac{7 a^2}{18 M_0^2}+
\frac{8}{9}\frac{b\alpha_3}{M_0^2}-\frac{17}{54}\sqrt{\frac{2}{3}}\frac{ba\alpha_3}{M_0^3}\nonumber\\
&{}-\left(\frac{16297}{38880}-\frac{22267 a}{17496 \sqrt{6} M_0}\right) \frac{\alpha_3^2}{M_0^4}+
\frac{77 a}{216 \sqrt{6} M_0^5}\alpha_4^2,\nonumber\\
M\Omega_\text{ISCO}={}&\frac{1}{6 \sqrt{6}}+\frac{11 a}{216 M_0}+\frac{59 a^2}{648 \sqrt{6} M_0^2}
-\frac{12113 a}{5225472 M_0^5}\alpha_4^2\nonumber\\
&{}-\frac{29}{432 \sqrt{6}}\frac{b\alpha_3}{M_0^2}-\frac{169}{7776}\frac{ba\alpha_3}{M_0^3}+
\left(\frac{32159}{2099520 \sqrt{6}}-\frac{49981 a}{75582720 M_0}\right) \frac{\alpha_3^2}{M_0^4}\,.\label{eq_romega_isco}
\end{align}
We have kept only dominant terms in $b$, and for simplicity we focused
on corotating orbits (the result for counterrotating orbits is
trivially obtained by inverting the sign of $a$). Note that $a/M_0$ is
not the physical dimensionless angular momentum, $J/M^2$, but
it can be easily related to the latter to second order in
$\alpha_3/M^2$ using
Eq.~\eqref{EdGB:mass}~\cite{Maselli:2013mva}.
The behavior of the ISCO frequency depends on several coupling
parameters. For $b=0$, the dominant correction is of order ${\cal
O}(\alpha_3^2)$ and tends to increase the ISCO frequency. The first
corrections proportional to the BH spin are ${\cal O}(a\alpha_3^2)$
and ${\cal O}(a\alpha_4^2)$, and they contribute to lower the
frequency. However, when a nonminimal coupling is turned on, its
effect is dominant~\cite{Pani:2009wy}. When $b\neq0$, the ISCO frequency gets
corrections of order ${\cal O}(b\alpha_3).$\footnote{%
Note that a nonminimal coupling to the
matter sector violates the weak equivalence principle. Since the
latter is tested within the astonishing precision of 1 part in
$10^{13}$~\cite{Will:2014xja}, a very stringent bound on $b$ can be
derived: $b\alpha_i/M^2<10^{-13}$.}
For null geodesics, the light-ring frequency $\Omega_{\rm LR}=L_{\rm
LR}/E_{\rm LR}$ (related to the real part of the ringdown frequency
of the BH in the eikonal limit~\cite{Cardoso:2008bp}) and the
light-ring radius $R_{\rm LR}$ do not depend on the coupling $\gamma$:
\begin{align}
\frac{R_{\rm LR}}{M}={}&3-\frac{2 a}{\sqrt{3} M_0}-\frac{2 a^2}{9 M^2_0}+
\frac{31}{81 \sqrt{3}}\frac{a\alpha_4^2}{M_0^5}-\left(\frac{961}{3240}-
\frac{33667 a}{174960 \sqrt{3} M_0}\right) \frac{\alpha_3^2}{M_0^4}\,,\\
M\Omega_{\rm LR}={}&\frac{1}{3 \sqrt{3}}+\frac{2 a}{27 M_0}+\frac{11 a^2}{162 \sqrt{3} M_0^2}
-\frac{131}{20412} \frac{a\alpha_4^2}{M^5_0}+\left(\frac{4397}{262440 \sqrt{3}}+
\frac{24779 a}{4723920 M_0}\right) \frac{\alpha_3^2}{M^4_0} \,.
\end{align}
The dominant correction is ${\cal O}(\alpha_3^2)$ and increases the
frequency, whereas the ${\cal O}(a\alpha_3^2)$ and ${\cal
O}(a\alpha_4^2)$ corrections have opposite relative signs.
\subsubsection{EdGB theory}\label{sec:BH_EdGB}
\paragraph{Static and slowly rotating solutions.}
Perturbative BH solutions in EdGB gravity can be obtained as special
cases of the results discussed above. Besides these perturbative results, an exact
static solution in EdGB gravity (i.e., a solution going beyond the perturbative level in $\alpha_3$) is also
known~\cite{Kanti:1995vq}. It has the form
\begin{equation}
ds^2=-f(r)dt^2+g(r)^{-1}dr^2+r^2d\theta^2+r^2\sin^2\theta d\varphi^2
\,,
\end{equation}
where the metric functions $f(r)$, $g(r)$ and the scalar field $\phi(r)$ can be found by solving a system of ordinary
differential equations. The solution (regular at the horizon and at infinity) only exists
when~\cite{Kanti:1995vq,Pani:2009wy}
\begin{equation}
0<\alpha_3/M^2\lesssim0.691\,.
\end{equation}
Ref.~\cite{Pani:2009wy} extended these results to the slow-rotation
case, analyzing geodesics and QNMs in the eikonal
limit. Ref.~\cite{Maselli:2014fca} studied the epicyclic frequencies
of this solution, with the aim of constraining EdGB gravity through
observations of BH quasi-periodic oscillations.
Ref.~\cite{Ayzenberg:2014aka} analyzed corrections of second order in
the spin working in the small-coupling limit. The character of the
solution changes quite dramatically: the solution at first order in
spin is algebraically special (i.e.~of Petrov type D, just like the
Kerr metric), while the second-order solution is of Petrov type I. The
Petrov type of the metric is very important, because it is related to
the separability of the field equations. This example illustrates the
importance of obtaining {\em exact} BH solutions (rather than
perturbative expansions) when analyzing their features in modified
theories of gravity.
\paragraph{Rapidly rotating solutions.}
Rapidly rotating BHs in EdGB theory have been obtained
numerically~\cite{Kleihaus:2011tg} using the Lewis-Papapetrou ansatz
for a
stationary, axially symmetric spacetime. The line element can be
parametrized as
\begin{equation}
ds^2
= - e^{2 \nu_0}dt^2
+e^{2(\nu_1-\nu_0)}\left[e^{2 \nu_2}\left(dr^2+r^2 d\theta^2\right)
+r^2 \sin^2\theta
\left(d\varphi-\omega dt\right)^2\right] ,
\label{metric2}
\end{equation}
where $\nu_0$, $\nu_1$, $\nu_2$ and $\omega$ are functions of $r$ and
$\theta$ only. BH solutions are asymptotically flat and possess the
expansion
\begin{align}
\nu_0 & = -\frac{M}{r} + \frac{D_1 M}{3r^3} - \frac{M_2}{r^3} P_2(\cos\theta) +{\cal O}(r^{-4}) ,
\label{exnu0} \\
\nu_1 & = \frac{D_1}{r^2} +{\cal O}(r^{-3}) ,
\label{exnu1} \\
\nu_2 & = -\frac{4 M^2+16 D_1+ q^2}{8 r^2}\sin^2\theta +{\cal O}(r^{-3}) ,
\label{exnu2} \\
\omega & = \frac{2 J}{r^3} +{\cal O}(r^{-4}) ,
\label{exom} \\
\phi & = \frac{q}{r} +{\cal O}(r^{-2}) ,
\label{exdil}
\end{align}
where $P_2(\cos\theta)$ is a Legendre polynomial, and $\phi$ denotes the dilaton field. The expansion constants $M$,
$J$, and $q$ denote the (ADM) mass, the angular momentum and the dilaton charge, respectively. The expansion
(\ref{exnu0})-(\ref{exdil}) also depends on the constants $D_1$, $M_2$.
The quadrupole moment $Q$ of EdGB BHs reads~\cite{Kleihaus:2014lba}
\begin{equation}
Q
= -M_2 +\frac{4}{3}\left[\frac{1}{4}+\frac{D_1}{M^2}
+\frac{q^2}{16M^2}\right] M^3 .
\label{Q}
\end{equation}
This has been obtained by extending the formalism of Geroch and
Hansen~\cite{Geroch:1970cd,Hansen:1974zz}.\footnote{The scalar field
of the static BH in EdGB gravity decays as $1/r$ at large distances,
similarly to the electric field of a Reissner-Nordstr\"om BH. The
Geroch-Hansen formalism to compute multipole moments was extended to
stationary electrovacuum spacetimes in
Refs.~\cite{0264-9381-7-10-012,Sotiriou:2004ud}. Using the fact that
the Gauss-Bonnet curvature term $R_{\rm GB}$ decays very quickly at large distances, the
structure of the first multipoles can be shown to be equivalent to
that of a Reissner-Nordstr\"om BH with a suitable indentification of
the scalar charge.}
\begin{figure}
\begin{center}
\begin{tabular}{cc}
\includegraphics[width=.5\textwidth, angle=0, clip=true]{KKRfig1b-eps-converted-to.pdf}&
\includegraphics[width=.5\textwidth, angle=0, clip=true]{KKMfig1a-eps-converted-to.pdf}\\
\end{tabular}
\end{center}
\caption{The domain of existence of EDBG BHs (shaded area). We plot
the scaled horizon area $a_{\rm H}=A_{\rm H}/M^2$ (left panel) and
the scaled quadrupole moment $\hat{Q}=Q M/J^2$ (right panel) as
functions of the scaled angular momentum $j=J/M^2$. Different
curves correspond to families of EDBG BHs with fixed scaled horizon
angular velocity $\Omega_{\rm H}
\alpha_3^{1/2}$. [From~\cite{Kleihaus:2011tg,Kleihaus:2014lba}.]}
\label{KKRfig1}
\end{figure}
The domain of existence of EdGB BHs is illustrated by the shaded area
in the left panel of Figure~\ref{KKRfig1}. The figure shows the scaled
horizon area $a_{\rm H}=A_{\rm H}/M^2$ (where $A_{\rm H}$ is the BH area)
as a function of the scaled
angular momentum $j=J/M^2$. The upper-left edge corresponds to
Schwarzschild BHs, which are all mapped to the point $a_{\rm H}=1$,
$j=0$. Likewise, Kerr BHs lie on a curve in this plot, i.e.~the upper
boundary of the domain of existence (except for $j\approx 1$: see
inset).
The lower boundary of the domain of existence corresponds to critical
BH solutions. These arise when the argument of a square root in the
expansion of the dilaton function at the horizon
vanishes~\cite{Kanti:1995vq,Torii:1996yi,Kleihaus:2011tg}. For a
given value of the coupling constant $\alpha_3$ and of the mass, EdGB BHs possess lower horizon area than
Kerr BHs. A remarkable feature is that EdGB BHs can slightly exceed
the Kerr bound ($j \le 1$) for the dimensionless angular momentum.
Only the metric functions for EdGB solutions with $j\ge 1$ are well
defined, while the dilaton field diverges at the poles at the horizon.
The right panel of Figure~\ref{KKRfig1} shows the rescaled quadrupole
moment $\hat{Q}=Q M/J^2$~\cite{Kleihaus:2014lba}. $\hat{Q}$ is
largest for slow rotation, and decreases with increasing $j$. The
deviations of $\hat{Q}$ from the corresponding Kerr values can be up
to 20\% and more. Superspinning EdGB BHs with $j>1$ always have $\hat
Q>1$.
\begin{figure}
\begin{center}
\includegraphics[width=9.2cm, angle=0, clip=true]{KKMfig1b-eps-converted-to.pdf}
\end{center}
\caption{The scaled moment of inertia $\hat{I} = J/(\Omega_{\rm H}
M^3)$ is shown versus the scaled quadrupole moment $\hat{Q}$ for
fixed values of $j$. The Kerr BHs are
indicated by the fat dots on the $\hat{I}$-axis. The straight
dotted lines represent the perturbative results
of~\cite{Ayzenberg:2014aka}. The critical BHs are represented by
the dotted curve. [From~\cite{Kleihaus:2014lba}.]
}
\label{KKMfig1}
\end{figure}
Figure~\ref{KKMfig1} exhibits the scaled moment of inertia $\hat{I} =
J/(\Omega_{\rm H} M^3)$ versus the scaled quadrupole moment $\hat{Q}$
for fixed values of $j$~\cite{Kleihaus:2014lba}. The Kerr values
$\hat{I}_{\rm Kerr}=2(1+\sqrt{1-j^2})$ and $\hat{Q}_{\rm Kerr}=1$ are
indicated by dots on the vertical axis of the plot, and represent the minimum
possible values for $\hat{Q}$. Families of EdGB solutions terminate at
the critical solutions represented by the dotted curve. For
comparison, straight, dotted lines show the perturbative results
of~\cite{Ayzenberg:2014aka}, derived for small $\alpha_3$ and small
$j$.
The inset shows the region $j>1$, not present in GR. The extraction
of higher multipole moments from the numerical solutions is still an
open problem.
The study of geodesic motion around rapidly rotating EdGB BHs unveiled
some interesting features~\cite{Pani:2009wy,Kleihaus:2011tg}. Timelike
geodesics for circular motion are obtained from the Lagrangian
$2 {\cal L} = e^{2 b \phi} g_{\mu\nu}\dot x^\mu \dot x^\nu =-1,$
where again the constant $b$ fixes the coupling between the matter and
the dilaton field ($b=1/2$ corresponds to the low-energy limit of
heterotic string theory).
\begin{figure}
\begin{center}
\begin{tabular}{cc}
\includegraphics[width=.47\textwidth, angle=0, clip=true]{KKRfig2b-eps-converted-to.pdf}&
\includegraphics[width=.47\textwidth, angle=0, clip=true]{KKRfig2c-eps-converted-to.pdf}\\
\end{tabular}
\end{center}
\caption{%
(Left) The scaled circumferential ISCO radius $R_{\rm ISCO}/M$
and (right) the scaled ISCO frequency $\nu_{\rm ISCO}/\Omega_{\rm H}$
are shown versus the scaled angular
momentum $j=J/M^2$ for dilaton matter coupling constant $b=1/2$
for families of EDBG BHs with fixed scaled horizon angular
velocity $\Omega_{\rm H} \alpha_3^{1/2}$. [From~\cite{Kleihaus:2014lba}.] }
\label{KKRfig2}
\end{figure}
Figure~\ref{KKRfig2} shows the scaled circumferential ISCO radius
$R_{\rm ISCO}/M$ versus the scaled angular momentum $j=J/M^2$ for a
coupling constant $b=1/2$. The Kerr solutions and the extremal
EdGB solutions possess the smallest values for the scaled ISCO radius
$R_{\rm ISCO}/M$, whereas the maximal values of $R_{\rm ISCO}/M$ are
found for the critical EdGB solutions. Note that the ISCO radius is
not given in Boyer-Lindquist coordinates.
When the rescaled angular momentum is large, the deviation of $R_{\rm
ISCO}/M$ from the Kerr value can be as large as 10\% for
$b=1/2$. Similarly, for large angular momentum the orbital
frequencies at the ISCO exhibit deviations from
the Kerr frequencies as large as 60\% [see~\eqref{eq_romega_isco} for a small-spin expansion].
Note, however, that the weak equivalence principle imposes $b=0$, so that
the corrections to the geodesic quantities are expected to be smaller.
Future space-based observations of the gravitational signal emitted by
extreme mass-ratio inspirals~\cite{Ryan:1995wh} and observations of
the electromagnetic signal associated to quasi-periodic oscillations
in low mass X-ray binaries~\cite{Feroci:2012qh,Pappas:2012nt,Vincent:2013uea} can be used to map the
spacetime around a BH. The key information is encoded in the orbital
frequency $\Omega_{\rm ISCO}$ and, more generally, in the epicyclic
frequencies ($\Omega_r$, $\Omega_\theta$, $\Omega_\phi$). These
quantities have been computed in~\cite{Vincent:2013uea} in the case of
dCS gravity, and in~\cite{Maselli:2014fca} for the case of EdGB gravity.
\paragraph{Stability.}
In the special cases investigated so far, BHs in EdGB gravity were
found to be linearly stable. Ref.~\cite{Kanti:1997br} studied radial
perturbations of static BHs, and Ref.~\cite{Pani:2009wy} considered
axial gravitational perturbations. The stability against polar
gravitational perturbations is an open problem.
\subsubsection{dCS theory}\label{sec:BH_dCS}
\begin{figure}[htb]
\centering
\includegraphics[width=0.7\columnwidth]{power-theta-j}
\caption{``Power'' in different multipole moments of the scalar
field $\tilde{\theta}$ (rescaled with $\alpha_4/M^2$)
around a rotating BH in dCS. The horizontal axis is the multipole
moment number $j$, i.e.~the coefficient of $P_{j}(\cos\theta)$
(only the odd coefficients are plotted). The vertical axis is the
$L^{2}$ norm of $\tilde{\theta}_{j}$ on a log scale. As spin
increases, the exponential convergence slows down, and there is
more power in higher multipole moments. [From~\cite{Stein:2014xba}.]}
\label{fig:BHs/dCS/power-theta}
\end{figure}
The field equations of dCS gravity in spherical symmetry reduce to
GR~\cite{Grumiller:2007rv,Molina:2010fb}, so static BH solutions are
given by the Schwarzschild metric.
Spinning BHs in dCS gravity are more interesting, because they are
endowed with a nontrivial scalar field sourced by a nonvanishing
Pontryagin density (${{}^{*}\!RR}\neq0$). Spinning solutions at first
order in a slow-rotation expansion were computed
in~\cite{Yunes:2009hc,Konno:2009kg}. These solutions can be obtained from the
general slowly rotating BH solution for quadratic gravity discussed in
Section~\ref{subsec:BHquadraticPert} by setting $\alpha_3=0$, and they
have been extended to second order in the BH spin~\cite{Yagi:2012ya}.
In the slow-rotation limit, the scalar field is dominated by a dipole
moment which is proportional to the spin of the BH. The correction to
the metric quadrupole moment was computed in~\cite{Yagi:2012ya}. The
geometry is of Petrov Type D at first order in rotation, and of Petrov
Type I at higher order. For arbitrary rotation, the scalar field
profile and trace of the metric deformation were computed
in~\cite{Stein:2014xba}. As rotation increases, the higher multipole
moments of the scalar field are sourced more strongly, as seen in
Figure~\ref{fig:BHs/dCS/power-theta}.
\paragraph{Stability.}
The linear stability of Schwarzschild BHs in dCS gravity was studied
in Refs.~\cite{Cardoso:2009pk,Molina:2010fb}, where it was found that
these solutions are mode-stable against all gravitational and scalar
perturbations. The stability of slowly rotating solutions has not
been studied yet, but see~\cite{Garfinkle:2010zx,Ayzenberg:2013wua}
for a high-frequency analysis.
\subsection{Lorentz-violating theories\label{sec:BH_LV}}
The notion of a BH in Lorentz-violating gravity is at first glance less clear-cut than in GR. As mentioned in Section~\ref{lv-theories},
in the infrared limit the most generic Lorentz-violating gravity theories are Einstein-\AE ther and khronometric theory, which allow for spin-2 gravitons (like in GR),
but also for spin-0 and (in Einstein-\AE ther theory but not in khronometric gravity) also spin-1 gravitons. These propagating gravitational modes have speeds that are functions of the coupling parameters $c_i$ of the theories, and are
generally different from the speed of light appearing in the Maxwell equations and regulating the propagation of the electromagnetic field.
As a result, BHs in these theories, provided they exist, will present multiple horizons, namely: a ``matter horizon''
for the electromagnetic field and the other matter fields, which do not couple directly to the Lorentz violating \ae ther or khronon field (so as to
enforce the weak equivalence principle, cf.~Section~\ref{lv-theories});
a spin-2 horizon for the spin-2 gravitons; a spin-0 horizon for the spin-0 gravitons; and (for Einstein-\AE ther theory only) a spin-1 horizon for the spin-1 gravitons.
These horizons will generally lie at different locations, depending on the propagation velocity of the corresponding field. However, because of the
requirement that there be no gravitational Cherenkov radiation in these theories (cf.~discussion in Section~\ref{lv-theories}), for viable
values of the coupling constants the propagation speeds of the spin-2, spin-0 and (when present) spin-1 modes will be larger than (or equal to) the speed of
light, and therefore the spin-2, spin-1 and (when present) spin-0 horizons will be enclosed by the matter horizon.
The situation gets even more complicated if one interprets Einstein-\AE ther and khronometric theory as low-energy
limits of a more generic Lorentz-violating gravity theory containing higher-order spatial derivative terms in the
action. This is the case for instance in Ho\v rava gravity, whose action \eqref{SBPSHfull} reduces to that of
khronometric gravity in the infrared limit, but which contains fourth- and sixth-order spatial derivative terms
that are crucial for the power-counting renormalizability of the theory. The presence of those terms, as mentioned in
Section~\ref{lv-theories}, causes gravitons to obey nonlinear dispersion
relations [see Eq.~\eqref{mdsr}].
The matter degrees of freedom (and photons in particular) will also satisfy similar
nonlinear dispersion relations, although the coefficients
of the nonlinear terms
may be smaller than for gravitons,
and in particular sufficiently small to satisfy particle physics tests of Lorentz invariance~\cite{Kostelecky:2003fs,Kostelecky:2008ts,Mattingly:2005re,Jacobson:2005bg,Liberati:2013xla}, if the theory
efficiently suppresses percolation of the Lorentz violations from gravity to the matter sector (cf.~discussion in
Section~\ref{lv-theories}). From a conceptual point of view, however, Eq.~\eqref{mdsr} makes the very concept of an
event horizon meaningless in the ultraviolet limit, because it implies diverging propagation speeds $d\omega/dk$ in the limit
$k\to\infty$. Therefore, the question arises of whether the multiple event horizons discussed above are simply
low-energy artifacts.
To answer this question, Ref.~\cite{Barausse:2011pu} (building on Ref.~\cite{Eling:2006ec}) looked first at BH solutions in the infrared limit of Lorentz-violating gravity, i.e.~in
Einstein-\AE ther and khronometric theory. As mentioned in Section~\ref{lv-theories}, the two theories have exactly the same solutions for
static, spherically symmetric, asymptotically flat BHs. More specifically, using ingoing Eddington-Finkelstein coordinates, the most generic
static and spherically symmetric ansatz for the metric and the \ae ther is given by
\begin{align}
\label{efmetric}
ds^2&=-f(r)dv^2+2 B(r)dv dr+r^2d\Omega^2\,,\\
\label{efaether}\boldsymbol{u}&=A(r) \partial_v-\frac{1-f(r) A^2(r)}{2 B(r) A(r)}\,\partial_r\,.
\end{align}
Solving the field equations perturbatively near spatial infinity, and imposing asymptotic flatness, one obtains the series-expanded solution~\cite{Barausse:2011pu,Eling:2006ec}
\begin{align}
f(r) &= 1-\frac{r_g}{r}-\frac{c_{1}+c_{4} }{48} \frac{r_g^3}{r^3}+ \cdots\,, \label{asyF}\\
B(r) &= 1+\frac{c_{1}+c_{4}}{16} \frac{r_g^2}{r^2}+\frac{c_{1}+c_{4}}{12} \frac{r_g^3}{r^3}+\cdots\,, \label{asyB} \\
A(r) &= 1+\frac12 \frac{r_g}{r}+\frac{A_2 r_g^2}{r^2}-
\left(\frac{1}{16} -\frac{c_1+c_4}{96}-A_2\right) \frac{r_g^3}{r^3}+\cdots\,, \label{asyA}
\end{align}
where $r_g=2 G_N M/c^2$ [the locally measured gravitational constant $G_N$
being related to the ``bare'' one appearing in the action by Eq.~\eqref{eq:GN}], $M$
is the mass of the BH as measured by an observer far from the system, and $A_2$ is a dimensionless ``\ae{}ther charge.''
The latter can in principle take arbitrary values, but if one attempts to construct the BH solution corresponding
to a given $A_2$ value by integrating the field equations inwards starting from the asymptotic solution \eqref{asyF}--\eqref{asyA},
one obtains a solution that presents a finite-area singularity on the spin-0 horizon~\cite{Eling:2006ec,Barausse:2011pu}. Only for a specific value $A_2=A_2^{\rm reg}$ (a function
of the theory's coupling constants) is the BH regular everywhere except for the central $r=0$ singularity.
We stress that fully nonlinear numerical simulations~\cite{Garfinkle:2007bk} have shown that spherically symmetric gravitational collapse does indeed select the regular $A_2=A_2^{\rm reg}$
BH solution.
In practice, this regular BH solution is found
by solving the field equations perturbatively near the spin-0 horizon, imposing that the solution be regular there, and then selecting the solution that
matches the asymptotically flat perturbative solution~\eqref{asyF}--\eqref{asyA} by a shooting procedure (see~\cite{Barausse:2011pu,Barausse:2013nwa} for more details).
After the shooting procedure has selected the asymptotically flat solution, the behavior in the interior can be obtained by integrating inwards from the spin-0 horizon.
\begin{figure}[tb]
\capstart
\begin{center}
\begin{tabular}{lr}
\includegraphics[width=6.2cm,clip=true]{wiscoAE}
& \includegraphics[width=6.2cm,clip=true]{wiscoHL}
\end{tabular}
\caption{Fractional deviation of the dimensionless combination $\omega_{_{\rm ISCO}} r_g$ from its GR value, in Einstein-\AE ther (left) and khronometric theory (right).
Negative values denote smaller values in Lorentz-violating gravity than in GR. The quantities $c_\pm$, $\mu$, $\beta$ are defined in Section~\ref{subsec:lorentz-viol}.
[From~\cite{Barausse:2013nwa,Barausse:2011pu}.]\label{isco}}
\end{center}
\end{figure}
\begin{figure}[htb]
\capstart
\begin{center}
\begin{tabular}{lr}
\includegraphics[width=6.2cm,clip=true]{bphAE}
& \includegraphics[width=6.2cm,clip=true]{bphHL}
\end{tabular}
\caption{Fractional deviation of the dimensionless combination $b_{\rm ph}/r_g$ from its GR value, in Einstein-\AE ther (left) and khronometric theory (right).
Postive values denote larger values in in Lorentz-violating gravity than in GR. The quantities $c_\pm$, $\mu$, $\beta$ are defined in Section~\ref{subsec:lorentz-viol}.
[From~\cite{Barausse:2013nwa,Barausse:2011pu}.]\label{bph}}
\end{center}
\end{figure}
The resulting solutions will therefore describe the BHs of infrared Lorentz-violating gravity and present multiple horizons, as discussed above. However, in spite of
the causal structure differing from GR, the BH geometry outside the outermost horizon (i.e.~the matter one) is very similar to GR as far as
astrophysical tests are concerned. As two representative examples, Figures~\ref{isco} and~\ref{bph} show the fractional
deviation from GR of the dimensionless combinations $\omega_{_{\rm ISCO}} r_g$ and $b_{\rm ph}/r_g$,
where $\omega_{_{\rm ISCO}}$ is the orbital frequency of the innermost stable circular orbit (ISCO) and
$b_{\rm ph}$ is the impact parameter of the circular photon orbit. The former is measurable, at least in
principle, with GW observations of the inspiral of binary BH systems or with observations of iron-K$\alpha$ emission lines from
accretion disks, while the latter regulates the size of the BH ``shadow'' observable with future electromagnetic telescopes,
as well as the BH ringdown frequencies, in principle measurable with GW detectors.
For values of
the couplings allowed by binary pulsar observations (cf.~the purple region in Figure~\ref{fig:LVconstraints}), deviations from GR are below the percent level,
and thus outside the reach of electromagnetic observations (cf.~e.g.~\cite{Bambi:2011jq}), although probably within the reach of space-based detectors such as eLISA~\cite{Seoane:2013qna}.
The behavior of these BH solutions is however very different from GR inside the matter horizon. Figure~\ref{causal} shows a
spacetime diagram that captures schematically the causal stucture of the BH solutions studied in~\cite{Barausse:2011pu}. Hypersurfaces of constant preferred time $T$ are represented by green lines that get darker and darker as the value of $T$ increases (i.e.~a darker green means that the curve
is farther in the future). Also shown (in red) are two very special hypersurfaces of constant $T$, namely ones that are also hypersurfaces of constant radius. Those
hypersurfaces lie within the matter, spin-2, spin-0 and (when present) spin-1 horizons, and act as \textit{universal} horizons for signals of \textit{arbitrary} speed~\cite{Barausse:2011pu,Blas:2011ni,Barausse:2013nwa}. This can be
understood because a signal emitted at the universal horizon must propagate into the future, as defined by the preferred time $T$. As a result, as can be seen from Figure~\ref{causal},
such a signal must propagate inwards (i.e.~towards smaller radii).
Note that the solutions of~\cite{Barausse:2011pu} present multiple universal horizons, but we are truncating Figure~\ref{causal}
to show just the outermost two. Also, it is clear that the concept of a universal horizon only makes sense in the presence of the \textit{ultraviolet}
higher-order spatial derivative terms in the action,
which produce the nonlinear dispersion relation \eqref{mdsr}, thus allowing for the infinite-speed signals for which the universal horizon is relevant.
As mentioned earlier, the BH
solutions of~\cite{Barausse:2011pu} were instead derived by solving the field equations for the \textit{infrared} limit of Lorentz-violating gravity theories. Nevertheless, the universal horizon of those
solutions lies very close to the matter horizon and far from the central singularity, and thus in a region of small curvature for the
BH masses that are relevant in astrophysics. Indeed, simple dimensional arguments show that for astrophysical BHs the effect of the higher-order derivative terms is tiny at the location of the
universal horizon, leading to corrections $\lesssim 10^{-16} (M_{\odot}/M)^2$ ($M$ being the BH mass) away from the results obtained with the infrared limit of the theory. Universal horizons have been shown to be compatible with the
first~\cite{Berglund:2012bu} -- and possibly the
second~\cite{Cropp:2013sea} -- law of BH thermodynamics.
Finally, we stress that while the presence of such a universal horizon is a remarkable feature of the theory, clues of its instability to nonlinear perturbations
have been reported in the decoupling limit (i.e.~neglecting the backreaction of the \ae ther/khronon on the metric) and for the low-energy limit of
Ho\v rava gravity (i.e.~khronometric theory)~\cite{Blas:2011ni}. A fully nonlinear analysis accounting for the \ae ther's/khronon's backreaction
has not been performed yet, and it is needed to draw definitive conclusions about stability. Similarly, the effect of the higher-order spatial derivative terms on the stability of the universal horizon is unknown.
\begin{figure}[tb]
\capstart
\begin{center}
\includegraphics[width=9.2cm,clip=true]{causal}
\caption{Sketch of the causal structure of a BH possessing a universal horizon, marked with a red vertical line.
The green curves are hypersurfaces of constant preferred time $T$ (the darker the color, the larger $T$).
The universal horizon itself is a hypersurface of constant $T$. As can be seen, a signal emitted at the universal horizon has to travel inwards,
simply because it has to propagate in the future direction as defined by the
preferred time $T$. Note that multiple universal horizons are generally present,
but here we are truncating the region between the first two and the central singularity.
[From~\cite{Barausse:2013nwa}.]\label{causal}}
\end{center}
\end{figure}
\subsection{Massive gravity}
BH solutions in massive-gravity theories are still largely
unexplored. In massive bigravity theories there are asymptotically flat
solutions for which the reference metric $f_{\mu\nu}$ equals the
spacetime metric ($f_{\mu\nu}=g_{\mu\nu}$). One such family of
solutions includes the Kerr metric.
The superradiant instability responsible for hairy BH solutions in
theories of minimally coupled massive gravity
(cf.~Section~\ref{subsec:massgrav}) also destabilizes Kerr BHs in
massive gravity~\cite{Brito:2013wya}. Because of this instability,
astrophysical BH spin measurements imply that the graviton mass $\mu$
in any theory of massive gravity must be smaller than
$5\times 10^{-23}$ eV~\cite{Brito:2013wya}
(cf. Section~\ref{sec:superradiance} for a discussion of similar
constraints on the masses of ultralight scalar and vector fields).
Graviton masses $\mu$ {\it smaller} than a threshold value $\mu
M\leq0.438$ (in $G=c=\hbar=1$ units, with $M$ the BH mass) trigger
yet another instability against monopole fluctuations, that plagues
even nonrotating
BHs~\cite{Babichev:2013una,Brito:2013wya,Brito:2013yxa}. Quite
remarkably, the mass coupling $\mu M$ is well within the instability
region for values of $M$ and $\mu$ that are phenomenologically
relevant. In a cosmological context it is natural to consider the
graviton mass to be of the order of the Hubble constant, i.e.~%
$\mu\sim H\sim 10^{-33}{\rm eV}$~\cite{Hinterbichler:2011tt}. Such a
tiny graviton mass would destabilize any Schwarzschild BH with mass
smaller than $10^{22} M_\odot$!
Instabilities often signal the existence of a new family of
equilibrium solutions. Because there are no complex fields in massive
gravity, the superradiant instability presumably drives rotating BHs
to slower rotation rates. However, the monopole instability affecting
nonrotating BHs hints at the existence of a truly new family of BH
solutions~\cite{Brito:2013yxa,Brito:2013xaa}. These ``hairy''
solutions have metrics of the form
\begin{align}
g_{\mu\nu}dx^{\mu}dx^{\nu}&=-F(r)^2\, dt^2 + B(r)^{-2}\, dr^2 + r^2 d\Omega^2\,,\nonumber\\
f_{\mu\nu}dx^{\mu}dx^{\nu}&=-p(r)^2\, dt^2 + \left[U'(r)\right]^{2}/Y(r)^2\, dr^2 + \left[U(r)\right]^2 d\Omega^2\,,\nonumber
\end{align}
where $'\equiv d/dr$. Such asymptotically flat, hairy BH solutions were indeed
found numerically in~\cite{Brito:2013xaa}; their properties depend on the
particular theory under consideration, i.e.~on the values of the parameters
$\alpha_3$ and $\alpha_4$ as defined in Eq.~\eqref{eq:S-dRGT}.
Notebooks to generate these hairy
solutions are available online~\cite{DyBHo:web}.
Additional BH solutions may exist when the fiducial metric is not
proportional to the spacetime metric. The only asymptotically flat
solutions found so far belong to the Kerr
family~\cite{Babichev:2014tfa,Volkov:2014ooa} and monopole
fluctuations are {\it stable} for these
configurations~\cite{Babichev:2014oua}. Stability against nonradial
modes and superradiant amplification has not been studied at the
time of writing.
\subsection{Gravity with auxiliary fields\label{BI_BF}}
As discussed in Section~\ref{subsec:auxiliary}, gravitational theories
that modify GR by adding solely nondynamical fields are
\emph{equivalent} to Einstein's theory in vacuum. In these theories,
vacuum BH solutions and their dynamics are the same as in GR. In
particular, any stationary, regular and asymptotically flat geometry
is described by the Kerr family. However, corrections to GR appear in
the coupling with matter. An interesting aspect of these theories is
that singularities may not form during
gravitational collapse~\cite{Pani:2011mg} and in early cosmology~\cite{Banados:2010ix}.
Nonvacuum solutions -- such as charged BHs -- are generally different from
GR~\cite{Banados:2010ix}. Furthermore, similarly to the regularization
of the Coulomb field generated by a point charge in Born-Infeld
electromagnetism, the curvature singularity hosted in the BH interior
in EiBI gravity may be replaced by a regular, wormhole-like geometry
due to nonperturbative effects~\cite{Olmo:2013gqa}.
\subsection{Parametrized phenomenological deviations from the Kerr metric}
BH solutions and their properties are obviously dependent on the
theory they are derived from. Although many theories -- some of which
were described previously -- share the Schwarzschild and Kerr geometry
as stationary solutions, even in these cases their dynamical
properties (stability, GW emission, etcetera) depend on the field content
of the theory. Unfortunately, in the context of alternatives to
Einstein's theory, the possibilities are endless. Each theory has its
own family (or families) of BH solutions, and in the absence of
observational data in the strong-field regime, choosing one's favorite
theory is largely a matter of taste.
Thus, some efforts focus on {\it parametrizing} generic spacetimes,
rather than exploring specific theories. These efforts are in many
ways parallel to the PPN expansion, designed to parametrize
asymptotically flat spacetimes in the weak-field
regime~\cite{EddingtonBook,Will:1972zz,Nordtvedt:1972zz}. The PPN
approach facilitates tests of the weak-field regime of GR and is
particularly well suited to perform tests in the Solar System, which
translate into constraints on alternative theories of gravity. For
example, one can show that any metric theory of gravity yielding an
asymptotically flat spacetime admits the
expansion~\cite{Will:2014xja}:
\begin{align}
-g_{tt}&\to 1-\frac{2{M}}{r}+2(\beta-\gamma)\frac{{M^2}}{r^2}+{\cal O}(1/r^3)\,, \label{gttPPN}\\
g_{ij}&\to \delta_{ij}\left[1+2\gamma\frac{M}{r}+{\cal O}(1/r^2)\right]\,, \label{grrPPN}
\end{align}
where $M$ is the ADM mass, and the indices $(i,j)$ run over
asymptotically Cartesian coordinates.
The PPN parameters are very well constrained by observations,
$|\gamma-1|\lesssim10^{-5}$ and $|\beta-1|\lesssim
2.3\times10^{-4}$~\cite{Will:2014xja}. The success of the PPN approach
is rooted in the existence of an extensively studied, unique reference
metric, the Minkowski geometry. Because the metric is
post-Minkowskian, the meaning of the coordinates is clear, and so are
the physical predictions one can draw from the metric.
A comparable ``reference metric'' is lacking in the strong-field
regime. For this reason, developing a parametrized approach to
quantify deviations from GR is a nontrivial problem. Existing
attempts
deal with the construction of a generic parametrization of spinning
geometries which can be matched continuously onto the Kerr metric in
the strong- {\it and} in the weak-field regime. This is a formidable
task with no unique solution.
Several approaches have been proposed, each of them with their own
limitations, but all very similar in spirit (see
e.g.~\cite{Johannsen:2013rqa,Cardoso:2014rha,Rico_thesis}). The
original ``bumpy BH'' formalism assumes Einstein's equations, and
perturbs the Kerr metric to find BHs {\it in GR} distorted by small
amounts of unspecified
matter~\cite{Collins:2004ex,Vigeland:2009pr}. The metric computed
within this approach is supposed to be valid only in vacuum. This
formalism cannot be extended in a straightforward manner to test
alternative theories of gravity (but see~\cite{Gair:2011ym} for some
improvements over the analysis of~\cite{Vigeland:2011ji}).
A similar approach was used to build ``quasi-Kerr'' spacetimes, by
expanding generic slowly rotating spacetimes up to the lowest
nontrivial quadrupole moment~\cite{Glampedakis:2005cf}. These
solutions are not regular at the horizon. Stationary, axisymmetric
and asymptotically flat solutions of the vacuum Einstein equations
which do {\it not} describe BHs, most notably the so-called
Manko-Novikov spacetimes~\cite{1992CQGra...9.2477M}, have also been
used by several authors to model spacetimes in alternative theories
and to parametrize deviations from the Kerr
geometry~\cite{Gair:2007kr,Bambi:2011jq}.
To overcome some of the limitations of the parametrizations
above -- while introducing others -- it was recently proposed to build
on the
``Newman-Janis algorithm''\footnote{The Newman-Janis algorithm allows
one to generate the Kerr family of BH starting from the
Schwarzschild family~\cite{Newman:1965tw}. This approach works in GR, but is
bound to fail in general for modified theories of gravity.} to
generate spinning BH solutions in arbitrary theories of
gravity~\cite{Johannsen:2011dh}. Using suitable choices of parameters,
these solutions consist of small deformations of the Kerr geometry. At
variance with previous studies, this approach does not assume the
validity of Einstein's equations, nor the existence of an approximate
Carter constant~\cite{Vigeland:2011ji}. Even though the procedure
makes use of the -- unjustified, because the field equations are
unknown -- Newman-Janis transformation (see e.g.~\cite{Hansen:2013owa}
for some criticism), the final transformed metric could as well be the
ad-hoc starting point for the investigation of deviations from
GR~\cite{Rico_thesis}. Such parametrized metrics can {\it in
principle} be suitable for tests involving observations of the
images of inner accretion flows, X-ray observations of
relativistically broadened iron lines or of the continuum spectra of
accretion disks, for which a regular behavior very close to the event
horizon is crucial~\cite{Bambi:2014sfa}.
The generalized deformed Kerr metric in this approach is~\cite{Cardoso:2014rha}
\begin{align}
g_{tt}&=-F(1+h^t), \label{eq:JP2_1}\\
g_{rr}&=\frac{(r^2+a^2\cos^2\theta) (1+h^r)}{\Delta+a^2 \sin^2\theta h^r},\label{eq:JP2_2}\\
g_{\theta\theta}&=r^2+a^2\cos^2\theta,\label{eq:JP2_3}\\
g_{\phi\phi}&=\sin^2\theta \left\{r^2+a^2\cos^2\theta + a^2 \sin^2\theta \left[2 H - F (1+h^t) \right] \right\},\label{eq:JP2_4}\\
g_{t\phi}&=-a\sin^2\theta \left[H- F (1+h^t)\right], \label{eq:JP2_5}
\end{align}
where $F\equiv 1-2M_0r/\Sigma$, we have introduced $H\equiv\sqrt{(1+h^r)(1+h^t)}$,
\begin{equation}
h^i(r,\theta)\equiv\sum^\infty_{k=0}\left(\epsilon_{2k}^i+\epsilon_{2k+1}^i\frac{M_0r}{\Sigma}\right)\left(\frac{M_0^2}{\Sigma}\right)^k \label{hJP}
\end{equation}
are the small deformation quantities parametrizing deviations
from the Kerr geometry in terms of dimensionless numbers $\epsilon_k^i$, and $\Sigma=r^2+a^2\cos^2\theta$, $\Delta=r^2+a^2-2M_0r$.
\begin{figure}
\begin{center}
\begin{tabular}{c}
\includegraphics[width=.7\textwidth, angle=0, clip=true]{fig2.pdf}
\end{tabular}
\end{center}
\caption{
Relative corrections $\delta\Omega_k/\Omega_0$ to the ISCO frequency as a function of $J/{M^2}$ for the metric~\eqref{eq:JP2_1}--\eqref{eq:JP2_5} to linear order in $\epsilon_k\ll1$ up to $k=9$. The ISCO frequency reads $\Omega=\Omega_0+\sum_k\delta \Omega_k\epsilon_k$, where $\Omega_0$ is the ISCO frequency of a Kerr geometry. The small-coupling approximation requires $(\delta \Omega_k/\Omega_0)\epsilon_k\ll1$ for consistency.
Each $\epsilon_k-$line is built by setting to zero all other $\epsilon_i,\,i\neq k$.
The two panels refer to the corrections associated to $\epsilon_k^t$ (upper panel) and $\epsilon_k^r$ (lower panel), respectively. For ease of comparison, the range of the vertical axis is the same for both panels. In this case the total ISCO frequency reads $\Omega=\Omega_0+\sum_k\delta \Omega_k^t\epsilon_k^t+\sum_k\delta \Omega_k^r\epsilon_k^r$. The small-coupling approximation requires $(\delta \Omega_k^i/\Omega_0)\epsilon_k^i\ll1$ for consistency. [From~\cite{Cardoso:2014rha}.]
}
\label{fig:deltaOmegaISCO2}
\end{figure}
Imposing asymptotic flatness requires only
$\epsilon_0^t=\epsilon_0^r=0$, but does not imply any constraint on
$\epsilon_1^t$ and $\epsilon_1^r$. Expanding the metric
elements~\eqref{eq:JP2_1} and \eqref{eq:JP2_2} at infinity and
comparing with the PPN expansions~\eqref{gttPPN} and \eqref{grrPPN},
we can identify
\begin{align}
{M}&=M_0\left(1-{\epsilon_1^t}/{2}\right)\,,\label{mass2}\\
\epsilon_1^r&=-2-\gamma(\epsilon_1^t-2)\,,\\
2\epsilon_2^t&=(\beta-\gamma)(\epsilon_1^t-2)^2+4\epsilon_1^t\,.
\end{align}
Therefore, even imposing the GR values $\beta=\gamma=1$ supported by
observations, the parameters $\epsilon_1^t$, $\epsilon_2^r$ and all
the $\epsilon_k^i$'s with $k>2$ $(i=t,r)$ are left unconstrained.
Figure~\ref{fig:deltaOmegaISCO2} shows the shifts of the ISCO frequency for the generalized metric~\eqref{eq:JP2_1}--\eqref{eq:JP2_5} in the small-$\epsilon_k^i$ limit.
For low rotation rates the corrections associated to $\epsilon_k^t$
are larger than those associated to $\epsilon_k^r$, while the converse
is true for rapid rotation, i.e.~when $a/M\gtrsim 0.85$. An exception
to this behavior are the $\epsilon_1^i$ parameters, for which the
$t$--correction is larger than the $r$--correction for any spin.
The dominant corrections are the ones associated with $\epsilon_1^t$,
although in the fast-spinning case the corrections $\delta\Omega_k^r$
for different values of $k$ are all comparable to each other, and they
are also comparable to $\delta\Omega_1^t$. However, at least for
moderately large spin, the corrections $\delta\Omega_1^t$ and
$\delta\Omega_2^r$ are dominant. Note that both $\epsilon_1^t$ and
$\epsilon_2^r$ are currently unconstrained by observations, so that
their contribution would likely dominate the near-horizon geometry of
the deformed Kerr metric~\eqref{eq:JP2_1}--\eqref{eq:JP2_5}. A more
detailed analysis of this parametrization has recently appeared
in~\cite{Bambi:2014sfa,Bambi:2014mla}.
The approach summarized above relies on a Taylor expansion of the
unknown functions $h^t(r)$ and $h^r(r)$ in powers of
$M/r$. Continued-fraction resummations based on a compactified radial
coordinate have been recently proposed and explored for nonrotating
BHs~\cite{Rezzolla:2014mua}.
\subsection{BH mimickers}
Despite growing experimental evidence (see
e.g.~\cite{Broderick:2005xa}), at the moment an incontrovertible proof
that dark compact objects are indeed BHs (i.e.~that they possess an
event horizon or, at least, an apparent horizon) is lacking. In fact,
concerns have been raised on whether such a proof is possible at all
with electromagnetic observations~\cite{Abramowicz:2002vt}.
Our current understanding of stellar evolution strongly suggests that
even extreme forms of matter cannot support the enormous self-gravity
of massive and ultracompact objects, so that the latter are naturally
expected to be BHs. The above picture has been challenged by the
construction of exotic objects -- so-called ``BH mimickers'' -- relying
on different support mechanisms. These objects are all (almost) as
compact as BHs, but do not possess horizons. Among others, they
include \emph{boson stars}, consisting of self-gravitating massive
scalar fields~\cite{Liebling:2012fv,Macedo:2013qea}; gravitational
condensate stars or \emph{gravastars}~\cite{Mazur:2001fv}, supported
by an exotic EOS of the form $P(\rho)\approx -\rho$; and
\emph{superspinars}~\cite{Gimon:2007ur}, objects with angular momentum
exceeding the Kerr bound and with some form of matter replacing the
singular BH interior.
The key observational distinction between genuine BHs and
``mimickers'' is the presence of a surface. Experimental tests of this
property are challenging in the electromagnetic spectrum, but they should become
simpler in the context of future GW observations: the oscillation
modes of BHs have a very precise and well-known structure, which can
be tested against
observations~\cite{Berti:2005ys,Berti:2009kk,Berti:2006qt}, while the
presence of a surface will leave an imprint on the GWs generated
during the merger of two
objects~\cite{Kesden:2004qx,Macedo:2013qea,Pani:2009ss} (but see the
discussion about QNMs and ringdown modes in
Section~\ref{sec:environment} and Ref.~\cite{Barausse:2014tra}).
Some BH mimickers can be ruled out by purely theoretical arguments,
that generally rely on instabilities related to the absence
of the event horizon. Mimickers can be ruled out when these
instabilities grow on time scales much shorter than the age of the
Universe.
The theoretical foundation for the presence of these instabilities is
the work of Friedman, that showed how \emph{any spacetime with an
ergoregion but without a horizon is linearly
unstable}~\cite{1978CMaPh..63..243F}. The instability is due to
long-lived modes that exist for ultracompact objects whose radius is
$R\lesssim 3M$, and that might turn unstable because of the effects of
rotation~\cite{SchutzComins1978, Cardoso:2014sna}. Ultracompact
objects such as gravastars and boson stars become linearly unstable
when they possess an
ergoregion~\cite{SchutzComins1978,Cardoso:2007az}, with an instability
time scale that depends strongly on the compactness and
spin~\cite{Chirenti:2008pf}. The same instability affects also
superspinars~\cite{Cardoso:2008kj,Pani:2010jz}.
In addition to the ergoregion instability, a new mechanism could
exclude {\it any} ultracompact ``star'' on the grounds that such an
object would be nonlinearly unstable~\cite{Keir:2014oka}. In this case
the instability is due to the existence of long-lived modes in the
linearized spectrum. These modes are trapped between the center of the
object and the light ring, and they are localized near a second, {\em
stable} null geodesic~\cite{Cardoso:2014sna}. The long-lived modes
may become unstable under fragmentation via a
Dyson-Chandrasekhar-Fermi mechanism at the nonlinear
level. Alternatively, nonlinear interactions over their long life time
may lead to the formation of small BHs close to the stable light
ring~\cite{Cardoso:2014sna}.
If confirmed, the nonlinear instability results could soon give further support to the
BH hypothesis: the mere observation of a light ring -- a much simpler
task than the observation of the event horizon, and something that is
within the reach of upcoming
facilities~\cite{Johannsen:2015qca,Lu:2014zja,GRAVITY} -- would be
conclusive evidence for the existence of BHs.
\subsection{BHs as strong-gravity laboratories for exotic
fields}\label{subsec-bh-exotic}
Besides being the optimal testbed for tests of GR in the strong-curvature
regime, BHs can also be used to study exotic fields, as those appearing in
extensions of the Standard Model of particle physics and as dark-matter
candidates. This possibility stems from a surprising connection between
strong-field gravity and particle physics. Although not immediately related with
tests of GR, we conclude this chapter by discussing two examples in which the
interplay between BHs and exotic fields is particularly dramatic.
We consider the dynamics of scalar fields in the framework of Einstein's GR but
-- as it will be clear below -- the qualitative aspects of this analysis are
mostly independent of the underlying theory of gravity.
\subsubsection{Collapse of self-interacting scalar fields}
\label{sec:scalarCollapse}
One of the most important phenomena in GR where the nonlinearity of
the theory plays a crucial role is that of gravitational collapse (for
a review, see~\cite{Gundlach:2007gc} and references therein). A
particularly intriguing result in this context has recently been
discovered numerically by Bizo{\'n} and
Rostworowski~\cite{Bizon:2011gg}, namely the collapse to a BH of
arbitrarily small spherically symmetric, massless scalar field
configurations in asymptotically anti-de Sitter (AdS) spacetimes. The
AdS boundary plays a key role for the dynamics because, in contrast to
asymptotically flat spacetimes, the scalar field pulses reach spatial
infinity in finite time and get reflected back onto the coordinate
origin. This effective {\em confinement} of the spacetime combined
with the nonlinear interaction of the wave modes results in a resonant
transfer of energy to higher frequencies, i.e.~shorter
wavelengths~\cite{Bizon:2011gg,Dias:2011ss,Buchel:2012uh} (see
also~\cite{Bizon:2013xha}). On the other hand, there exist
asymptotically AdS scalar-field solutions which do not collapse into a
BH, such as time-periodic solutions or boson
stars~\cite{Buchel:2013uba,Dias:2012tq,Maliborski:2013jca}.
Gravitational collapse in these spacetimes could also be prevented by
the formation of nonlinear bound states of {\em massive} fields. Such
bound states have been studied extensively in asymptotically flat
spacetimes~\cite{Seidel:1991zh,Seidel:1993zk,Page:2003rd,Cardoso:2005vk,Dolan:2007mj,Pani:2012vp,Pani:2012bp,Witek:2012tr,Okawa:2014nda}.
To study the possibility of
asymptotically flat spacetimes being unstable in the context of
confinement mechanisms~\cite{Choptuik:1992jv,Brady:1997fj}, consider the action
\begin{equation}
S = \int d^4x \sqrt{-g}
\left( \frac{R}{16\pi}
-\frac{1}{2}\partial^{\mu}\varphi\, \partial_{\mu}\varphi
-\frac{1}{2}\mu^2\varphi^2 \right)\,.
\label{eq:actionOkawa}
\end{equation}
This choice corresponds to the special case $V(\varphi)/16\pi =
\frac{1}{2}\mu^2 \varphi^2$ and to a one-dimensional (hence flat)
target space in Eq.~\eqref{STactionE}, i.e., to a minimally coupled
massive scalar field of mass $\mu$. Applying the
Arnowitt-Deser-Misner (ADM) decomposition to the field equations
resulting from (\ref{eq:actionOkawa}) for the special case of
spherical symmetry, one obtains two constraint equations and a set of
evolution equations which are given explicitly in Eqs.~(3), (4)
of~\cite{Okawa:2013jba}. Initial data are constructed by analytically
solving the constraint equations for a Gaussian scalar pulse in a
Minkowski background.
\begin{figure}
\capstart{}
\centering
\includegraphics[width=5.cm]{phase_diagram-eps-converted-to.pdf}
\includegraphics[width=7.5cm]{Phi0_all_phase_v3-eps-converted-to.pdf}
\caption{Left panel: Qualitative phase diagram for the spherically symmetric
collapse of a massive scalar field in the amplitude $(A)$
vs. width $(w)$ plane. Right panel: The scalar field amplitude
at the coordinate origin as a function of time is shown for
selected values of the amplitude $A$. $M_0$ denotes the
ADM mass of the spacetime. [Adapted from~\cite{Okawa:2013jba}.]
}
\label{fig:okawa}
\end{figure}
The results of the numerical time evolutions are summarized in
Figure~\ref{fig:okawa}. The left panel of the figure shows the phase
diagram for the collapse of a massive scalar field in the
$w\mu$-$A/\mu$ plane, where $w\mu$ and $A/\mu$ represent the initial
pulse width to Compton wavelength ratio and the pulse amplitude,
respectively. For small values of the width, the phase diagram
reveals a behavior similar to the massless case: collapse above a
threshold initial amplitude, and dispersion below that threshold. For
large $w\mu$, however, the phase diagram exhibits a much richer
phenomenology. This is also demonstrated in the right panel of the
figure, where the scalar field amplitude at the coordinate origin is
shown as a function of time for several configurations with varying
initial amplitude. For large initial amplitude the scalar field
collapses promptly to a BH, irrespective of whether a mass term
$U(\phi) = \frac{1}{2}\mu^2 \phi^2$ is included in the action or
not. For smaller amplitudes, however, this mass term leads to a
delayed collapse, similar to that observed in the AdS case. The mass
term introduces an effective confinement of the scalar field, which
gets reflected off the potential barrier instead of escaping to
infinity, and thus collapses after some number of reflections: cf.~the
top three plots in the right panel. Additionally, there exist
meta-stable, long-lived oscillations (second lowest plot in the right
panel) which occur for smaller values of the initial
amplitudes. Finally, for very small $A$, the scalar field decays
(bottom plot).
In summary, the numerical evolutions demonstrate that for a massive
scalar field coupled to gravity there exist nonlinear bound states,
i.e.~meta-stable oscillations~\cite{Seidel:1991zh,Page:2003rd,Grandclement:2011wz,Okawa:2013jba}. This
implies that the mass term of the scalar field can lead to a
confinement-induced gravitational collapse similar to the AdS case,
but that in the asymptotically flat case, energy can escape to some
extent from the potential, which results (for some values of the
initial parameters) in a bound state, rather than BH formation through
gravitational collapse. Indeed, a rather generic class of arbitrarily
small initial data evolving in a totally confined geometry seems to be
generically unstable to BH formation~\cite{Okawa:2014nea}. If
confirmed, this result might have interesting implications for the
nonlinear stability of compact stars, whose fluid perturbations are
effectively confined within the stellar surface.
\subsubsection{Superradiant instabilities: black holes as observatories for beyond-standard-model physics}
\label{sec:superradiance}
One of the main reasons why BHs represent interesting
laboratories for the exploration of the properties of light bosonic fields is
the superradiant instability of spinning
BHs~\cite{zeldovich1,Press:1972zz,Cardoso:2004nk} (for a recent exhaustive
overview on the subject, see Ref.~\cite{Brito:2015oca}). As previously
discussed, superradiance
occurs in the interaction between BHs and fundamental fields
with frequencies $\omega \le m \Omega_{\rm H}$, where $\Omega_{H}$ is the
angular velocity of the BH horizon, and $m$ denotes the
azimuthal mode number. The interaction provides a {\em classical}
mechanism to reduce the mass and spin of the rotating hole as the
field taps into the rotational energy and gets amplified. A gedanken
experiment first proposed by Press and Teukolsky~\cite{Press:1972zz}
suggests that a superradiant system can be rendered unstable by a
run-away amplification of the field if it is immersed inside a
reflective cavity: this is a superradiant instability, or ``BH
bomb.''
Such a configuration naturally arises for the case of massive fields,
as in the action~\eqref{eq:actionOkawa}: as discussed above, the mass
term $\mu$ leads to a potential barrier, and thus acts as a
``mirror''~\cite{Damour:1976kh,Detweiler:1980uk,Zouros:1979iw}. In
that case, modes with $\omega \lesssim \mu$ are trapped inside the
potential well, and the rotating BH can become (superradiantly)
unstable against these modes. The growth rate of the BH-bomb
instability is regulated by the coupling between the field's mass
$\mu$ and the BH mass $M$, and it is strongest when these parameters
satisfy the condition $M \mu \sim \mathcal{O}(1)$. In other words, the
interaction is maximized when the Compton wavelength of the field is
comparable to the size of the BH. More specifically, perturbative
calculations predict that the strongest growth rates of scalar fields
surrounding a BH with dimensionless spin parameter $a/M \sim 0.99$ are
realized for the dipole mode with a coupling $M\mu\sim0.42$. The
$e$-folding times in this case are $\tau \sim 50
M/M_{\odot}$~\cite{Dolan:2007mj,Cardoso:2005vk}. That time
scale can decrease by several orders of magnitude if we consider
massive vector fields~\cite{Rosa:2011my,Pani:2012vp,Witek:2012tr} or
gravitons~\cite{Brito:2013wya}.
While the BH bomb mechanism is negligible for known composite or
fundamental scalar particles interacting with astrophysical BHs (the
mass coupling is $M \mu \ge 10^{18}$, yielding time scales longer than
the age of the Universe), it can play a significant role if the
field's mass is $10^{-22}~{\rm eV} \le \mu \le 10^{-8}~{\rm eV}$, as
might be the case for dark-matter candidates, ultra-light
axions~\cite{Arvanitaki:2009fg,Peccei:1977hh} or fundamental fields in
modified gravity
theories~\cite{Sotiriou:2008rp,Clifton:2011jh,Yunes:2013dva}. Given
that the superradiant amplification provides a mechanism to reduce the
energy and spin of a BH, one can argue that BHs with certain
parameters $(M,\,a/M)$ should not exist if they interact with fields
of mass $\mu$. Conversely, the observation of
BHs~\cite{McClintock:2009as,Reynolds:2013qqa,AmaroSeoane:2012km}
within these exclusion regions allows us to constrain the allowed
field masses. This effect has indeed been used to constrain the mass
of Proca fields by comparing the superradiant instability time scale
with the Salpeter time scale, which gives (roughly) the time it takes
to spin up and feed a BH through accretion. The bound obtained from
this study constrains the mass of a hypothetical light vector field,
$\mu_{\gamma} \le 10^{-20}~{\rm eV}$~\cite{Pani:2012vp}. Strictly
speaking, superradiant instabilities only exclude mass intervals
(superradiance is ineffective at large boson
masses~\cite{Brito:2015oca}), but the quoted upper limit on the vector
field mass takes into account previous constraints obtained by other
means. In principle this bound also applies to a hypothetical massive
photon, but in this case it may be necessary to model the interaction
of the photons with the surrounding accretion disk and plasma.
More solid bounds are in place for massive
gravitons~\cite{Brito:2013wya}, which are only weakly coupled to
matter. The superradiant instability under massive spin-2
perturbations is the strongest instability of the Kerr metric known to
date and, together with observations of rapidly spinning supermassive
BHs, imposes the constraint $\mu_g\lesssim5\times10^{-23}{\rm eV}$ on
the mass $\mu_g$ of the graviton~\cite{PDG}.
\begin{center}
\begin{table}[t]
\begin{tabular}{ccccc}
Field & & Bounds & & Reference \\
\hline
\hline
Scalar & $\mu \lesssim 5\times 10^{-20} {\rm eV} $ & $\cup$ & $ \mu \gtrsim 10^{-11} {\rm eV}$ & \cite{Arvanitaki:2010sy,Arvanitaki:2014wva} \\
Vector & $\mu_\gamma \lesssim 5\times 10^{-21} {\rm eV} $ & $\cup$ & $ \mu_\gamma \gtrsim 10^{-11} {\rm eV}$ & \cite{Pani:2012vp} \\
Tensor & $\mu_g \lesssim 5\times 10^{-23} {\rm eV} $ & $\cup$ & $ \mu_g \gtrsim 10^{-11} {\rm eV}$ & \cite{Brito:2013wya} \\
\hline
\end{tabular}
\caption{Current bounds on the mass of ultralight bosonic degrees of
freedom arising from BH superradiant instabilities within a
linearized approximation (cf.~\cite{Brito:2015oca} for details).}\label{tab:SR}
\end{table}
\end{center}
Table~\ref{tab:SR} summarizes the current bounds on the mass of
ultralight bosonic degrees of freedom arising from BH superradiant
instabilities (cf.~\cite{Brito:2015oca} for details).
These results follow from perturbative calculations, and leave various
questions unanswered. (i) What is the fate of the system if we include
back reaction? Does the instability persist or is the system driven
towards a stable regime? (ii) Is it possible to form in this manner a
``gravitational atom'' or ``pulsar,'' as has been suggested
in~\cite{Arvanitaki:2009fg,Arvanitaki:2010sy}? (iii) What would be the
observational signatures, including modifications of GW signals and
the radiation emitted by the field itself?
Reference~\cite{Brito:2014wla} has recently addressed these questions
by performing a quasi-adiabatic, fully relativistic evolution of the
superradiant instability of a Kerr BH including GW emission and gas
accretion. It turns out that GW emission does not have a significant
effect on the evolution of the BH, although it contributes to
dissipate the dipolar bosonic cloud that forms as a result of the
instability. The mass of the cloud can be a sizeable fraction of the
total BH mass, but its energy density is very low, because the cloud
typically extends over very large distances. This implies that
backreaction effects are always negligible: even in the presence of
effective bosonic ``hair'' (both for real and for complex fields), the
geometry remains close to Kerr. Thus, the prospects of imagining
deviations from Kerr due to superradiantly produced bosonic clouds in
the electromagnetic band are low, but such systems are a primary
source for observations aiming at testing the Kerr hypothesis through
GW detection.
Finally, the role of gas accretion is very important. On the one hand,
accretion competes against superradiant extraction of mass and angular
momentum. On the other hand, accretion might produce the optimal
conditions for superradiance, for example by increasing the BH spin
before the instability becomes effective or by increasing the
superradiant coupling $M \mu$.
In order to verify the theoretical bounds on the existence of light bosons~\cite{Arvanitaki:2010sy,Kodama:2011zc,Pani:2012vp,Brito:2013wya}, a relevant problem concerns the \emph{final} BH state at the time of observation in realistic situations. In other words, given the observation of an old BH and the measurement of its mass and spin, would these measurements be compatible with the evolution driven by superradiance, accretion and GW emission?
\begin{figure}[t]
\begin{center}
\begin{tabular}{c}
\includegraphics[width=0.7\textwidth]{ReggeMC-eps-converted-to.pdf}
\end{tabular}
\end{center}
\caption{\label{fig:ReggeMC} The final BH mass and spin in the Regge
plane~\cite{Arvanitaki:2010sy} for initial data consisting of
$N=10^3$ BHs with initial mass and spin randomly distributed between
$\log_{10}M_0\in[4,7.5]$ and $J_0/M_0^2\in[0.001,0.99]$. The BH
parameters are then extracted at $t=t_F$, where $t_F$ is distributed
on a Gaussian centered at $\bar t_{F}\sim 2\times 10^9{\rm yr}$ with
width $\sigma=0.1\bar t_{F}$. As an example we considered
$\mu=10^{-18}{\rm eV}$, but similar results hold for other
masses. The dashed blue line is the prediction of the linearized
analysis obtained by comparing the superradiant instability time
scale with the accretion time
scale~\cite{Arvanitaki:2010sy,Pani:2012vp,Brito:2013wya}, whereas
the solid green line is a new prediction computed
in~\cite{Brito:2014wla}. Old BHs do not populate the region above
the green threshold curve, especially for high accretion rates. The
experimental points with error bars refer to the supermassive BHs
listed in~\cite{Brenneman:2011wz}. [From~\cite{Brito:2014wla}.] }
\end{figure}
This problem is addressed in Figure~\ref{fig:ReggeMC}, which shows the
final BH mass and spin in the Regge plane~\cite{Arvanitaki:2010sy}
(i.e.~a BH mass-spin diagram) for $N=10^3$ Monte Carlo evolutions for
a scalar field mass $\mu=10^{-18}{\rm eV}$. We consider three
different accretion rates $f_{\rm Edd}$ (defined as the fraction of
mass accretion rate relative to the Eddington limit) and, in each
panel, we superimpose the bounds derived from the linearized analysis,
i.e.~the threshold line when the instability time scale equals the
accretion time scale. As a comparison, in the same plot we include the
experimental points for the measured mass and spin of some
supermassive BHs listed in~\cite{Brenneman:2011wz}. These results
confirm that a very solid prediction of the existence of ultralight
bosons is the appearance of ``holes'' in the Regge
plane~\cite{Arvanitaki:2010sy}, i.e.~regions of the BH mass-spin
diagram which should not be populated by old BHs. We refer
to~\cite{Brito:2014wla} for a detailed discussion.
A quasi-adiabatic evolution is well suited to studying superradiant
instabilities because of the existence of two very different
scales~\cite{Brito:2014wla}. One is dictated by the oscillation time
$\tau_S\sim1/\mu$, the other by the instability growth time scale,
$\tau\gg\tau_S$. In the most favorable case for the instability (that
of a massive scalar field), $\tau\sim 10^6\tau_S\sim 10^6 M$ is the
minimum evolution time scale required for the superradiant effects to
become noticeable. Thus, fully numerical simulations that capture the
effects of the instability are extremely challenging to perform.
\begin{figure}[t]
\capstart{}
\centering
\includegraphics[width=0.49\textwidth]{Plot_a95_relMspin-eps-converted-to.pdf}
\includegraphics[width=0.49\textwidth]{Plot_KerrMassive-eps-converted-to.pdf}
\caption{\label{fig:NonLin} Nonlinear evolutions of a scalar cloud
with $M\mu=0.3$ around a BH with initial spin $a/M=0.95$. Left:
fractional variation in the BH mass (top) and dimensionless spin
of the hole (bottom). At early times (shown in the inset), both
quantities {\em decrease} hinting at superradiance scattering. At
later times, the superradiance condition is no longer satisfied,
resulting in an increase of the BH mass due to accretion of the
scalar cloud. Right: dominant scalar (top) and induced
gravitational waveforms (bottom). [From~\cite{Okawa:2014nda}.]}
\end{figure}
The first fully nonlinear evolutions of massive scalars coupled to
Kerr BHs~\cite{Okawa:2014nda} have found evidence for superradiant
scattering in the early stages of the interaction, which has also been
observed in simulations modeling the infall of GWs into a rotating
BH~\cite{East:2013mfa}. After this scattering, however, the spin of
the BH decreases so much that the system is driven out of the
superradiant regime, and slow accretion of the scalar field dominates
the ensuing evolution.\footnote{Note that these evolutions focused
only on an isolated BH-scalar field system, neglecting the effects
due to the presence of ordinary matter or accretion disks.} These
features are illustrated in the left panel of
Figure~\ref{fig:NonLin}. Because of the mass term and the resulting
potential barrier the scalar field is trapped inside a region near the
BH, and therefore forms a scalar cloud which continuously leaks into
the rotating hole. This induces long-lived scalar and gravitational
radiation: cf.~right panel in Figure~\ref{fig:NonLin} (as well as the
animations available online at~\cite{DyBHo:web}). Although the
gravitational radiation itself is not sensitive to the mass potential
barrier, the continued interaction between the BH and the scalar cloud
excites a GW signal with about twice the scalar-field frequency, which
follows the beating pattern exhibited in the scalar
modes~\cite{Witek:2012tr,Okawa:2014nda}. Moreover, the frequencies in
both the scalar and GWs are of the order $f\sim
\mathcal{O}(10)~{\rm kHz} (M/M_{\odot})^{-1}$, implying that the
signals generated by stellar-mass or supermassive BHs are potentially
observable with the Advanced LIGO/VIRGO network~\cite{Aasi:2013wya} or
future space-based detectors such as eLISA~\cite{AmaroSeoane:2012km},
respectively (see~\cite{Arvanitaki:2014wva} for recent work on GW
signatures of bosonic clouds around BHs).
In addition to interesting constraints on particle masses,
superradiant mechanisms have the potential to test the existence and
geometry of extra dimensions~\cite{Rosa:2012uz}. Finally, the
possibility of indirect observation of superradiance in a BH-pulsar
system has recently been proposed. The idea is that the pulsar's GW
and electromagnetic luminosities may exhibit a characteristic
modulation due to superradiant scattering that depends on the pulsar
position relative to the BH~\cite{Rosa:2015hoa}. If observed, this would
be the first -- albeit indirect -- observation of rotational
superradiance, which forms the basis of all the superradiant
instabilities discussed here~\cite{Brito:2015oca}.
\subsection{Open problems}\label{op:BHs}
Here we give a (necessarily biased) list of open problems regarding BH
physics in the context of tests of gravity:
\begin{itemize}
\item The stability of the hairy BH solutions found
in~\cite{Herdeiro:2014goa}, as well as
their formation mechanism in astrophysical scenarios, remain an urgent open
issue.
\item In the context of stationary solutions, BHs in dCS gravity are
known only for low spin and were obtained analytically within a
perturbative scheme. Highly spinning dCS BHs have not been
constructed numerically yet.
\item The stability of BHs in quadratic gravity has been only
partially investigated. Notable missing investigations include:
polar gravitational perturbations of static EdGB BHs, any
perturbations of slowly rotating EdGB BHs, and the gravito-scalar
perturbations of slowly rotating dCS BHs. The latter might be
relevant in the context of Ostrogradski instabilities and the
well-posedness of the dCS gravity.
\item Similarly, the stability of BHs in Lorentz-violating theories
under gravitational perturbations has not been studied, even in
the static case. Such an analysis might be interesting in the
context of stability of the universal horizons.
\item The phenomenology of BHs in Horndeski theory, Galileon theory and
massive gravity has not been studied in detail yet. In theories
admitting BH solutions other than Kerr, it would be interesting to
understand whether such solutions are formed as the result of
gravitational collapse.
\item Despite various attempts, a solid strong-field parametrization
of spinning BHs in generic modified gravity is not yet available.
\item The endpoint of the superradiant instability of spinning BHs
triggered by massive bosons in full GR is unknown, due to the
long time scales of the problem. Nonlinear effects, such as
bosenova collapse~\cite{Yoshino:2012kn}, should be taken into
account in numerical simulations.
\end{itemize}
\subsection{General-relativistic stellar models} \label{sec:NS_GR}
Before discussing equilibrium stellar solutions in modified gravity,
it is convenient to present a brief summary of the basic properties of
relativistic stars in Einstein's theory (we refer the reader
to~\cite{Stergioulas:2003yp,Shapiro:1983du,FriedmanStergioulas} for
excellent treatments of the subject).
In GR, the equilibrium of spherically symmetric, static (nonrotating)
stars is governed by the Tolman-Oppenheimer-Volkoff (TOV) equations,
that follow from
Einstein's equations for a perfect-fluid stress energy tensor. When
supplemented with an equation of state (EOS) relating the fluid's
density and pressure, Einstein's equations form a closed system of
ordinary differential equations. The NS EOS encodes the
thermodynamical behavior of matter in the extreme conditions
prevailing in the NS interior. Despite recent progress (see
e.g.~\cite{Lattimer:2004pg,Lattimer:2006xb,Ozel:2010fw,Steiner:2010fz,Steiner:2012xt,Hebeler:2013nza,Psaltis:2013fha}),
the EOS is still largely unknown at the supranuclear densities
characterizing the NS core.
The solutions to the TOV equations are obtained, in general, by
numerical integration. They form a single-parameter family, where the
parameter labeling different solutions can be chosen to be (say) the
central density. Relativistic equilibrium configurations are
characterized by a maximum mass and a maximum compactness that depend
on the EOS. Uncertainties in the EOS translate into uncertainties in
the NS mass-radius relation. For example, for a typical NS mass $M\sim
1.4 M_\odot$, EOSs compatible with our current knowledge of nuclear
physics predict radii ranging from $\sim$6 to $\sim$16
kilometers~\cite{Steiner:2012xt}.
Generic rotating stellar models are more difficult to
construct. However, ``old'' NSs are expected to rotate rather slowly,
unless they are spun up by accretion from a companion; therefore,
perturbative calculations using a slow-rotation expansion are reliable
in many situations of astrophysical interest. The formalism to
construct slowly rotating NS models, developed in the seminal work by
Hartle and Thorne~\cite{Hartle:1967he,Hartle:1968ht}, has been pushed
up to fourth order in rotation~\cite{Yagi:2014bxa}. Various works~\cite{Berti:2004ny,Benhar:2005gi,Yagi:2014bxa} have shown that the
equilibrium properties of slowly rotating solutions compare favorably
with numerical codes that solve Einstein's equations in full
generality to construct models of relativistic stars with arbitrary
rotation rates (cf.~\cite{Stergioulas:2003yp,FriedmanStergioulas} for
reviews).
Linear perturbations of stellar configurations are complex and interesting,
even for static objects. Because of GW emission, the modes of
relativistic NSs (just like the modes of BHs) have a dissipative
component, i.e.~they are QNMs. In addition to the
standard fluid modes, that have well studied counterparts in the
Newtonian limit~\cite{1989nos..book.....U}, compact stars also possess
characteristic modes of oscillation (the so-called $w$-modes)
associated to pure spacetime perturbations, rather than fluid
displacements
(see~\cite{Kokkotas:1999bd,Ferrari:2007dd,Andersson:2009yt,FriedmanStergioulas}
for reviews). These $w$-modes are similar in nature to BH QNMs.
Therefore NS QNMs carry information about the
stellar geometry (as in the BH case), but in addition they can also be
used to infer the properties of the NS EOS. In fact, one of the main
scientific goals of third-generation Earth-based GW detectors is their
potential to fulfill the promise of ``GW
asteroseismology''~\cite{Andersson:1997rn,Benhar:2004xg,Andersson:2009yt}:
accurate GW measurements of the oscillation frequencies would allow us
to reconstruct the properties of the NS (something that is routinely
done in helioseismology) and therefore to constrain nuclear physics in
regimes that are out of reach for laboratory experiments. Rotating
compact stars are characterized by various instabilities (most
notably, the Chandrasekhar-Friedman-Schutz
instability~\cite{Chandrasekhar:1992pr,Friedman:1978hf} and the
related r-mode instability~\cite{Andersson:2000mf}) that have
important implications for their spin rate and evolution. This topic
is largely unexplored in modified gravity, but there is a very large
body of work on these instabilities and their implications in GR~\cite{Kokkotas:1999bd,Andersson:2006nr,Andersson:2009yt,FriedmanStergioulas}.
\subsection{Scalar-tensor theories} \label{sec:NS_ST}
It should come as no surprise that most of the work on NSs in modified
theories of gravity has focused on the simplest and arguably most natural extensions of GR,
namely scalar-tensor theories. In these theories, the properties of
static and spinning NSs and of their oscillation modes are well
understood.
The modified TOV equations of hydrostatic equilibrium in Brans-Dicke
theory were first studied by Salmona~\cite{Salmona:1967zz}. Soon
after, Nutku~\cite{1969ApJ...155..999N} explored the radial stability
of stellar models using a PN treatment. Hillebrandt and
Heintzmann~\cite{1974GReGr...5..663H} analyzed incompressible
(constant density) configurations.
These studies found that corrections to NS structure are typically
suppressed by a factor $1/\omega_{\rm BD}$, where $\omega_{\rm BD}$ is
the Brans-Dicke coupling constant. The current best bound $\omega_{\rm
BD}>40,000$~\cite{Will:2014xja} implies that the bulk properties of
NSs in the original Brans-Dicke theory deviate from GR by unmeasurable
amounts.
However, as pointed out by Damour and
Esposito-Far\`ese~\cite{Damour:1993hw,Damour:1996ke}, for a particular
class of scalar-tensor theories that is indistinguishable from GR in
the weak field regime [more precisely, when $\alpha_0=0$ and
$\beta_0<0$ in the expansion of the Einstein-frame coupling function
\eqref{DEalphabeta}], a nonlinear phenomenon called ``spontaneous
scalarization'' can occur, introducing macroscopically (and
observationally) significant modifications to the structure of the
star\footnote{For a comprehensive study of analytic solutions and an
extensive bibliography, see~\cite{Horbatsch:2010hj}. Note in
particular that Tsuchida et al.~\cite{Tsuchida:1998jw} extended the
Buchdahl inequality ($M/R\leq 4/9$ for incompressible stars) to
generalized scalar-tensor theories.}. In addition the solutions
become nonunique: for certain ranges of the parameter space, NS
solutions in GR coexist with scalarized NSs. One of the most
interesting observations in~\cite{Damour:1993hw,Damour:1996ke} is that
scalarized configurations are energetically favored over their GR
counterparts. Scalarization occurs also for BHs in the presence of
matter
fields~\cite{Stefanov:2008,Doneva:2010,Cardoso:2013fwa,Cardoso:2013opa}.
A simple way to illustrate the principle behind spontaneous
scalarization is by taking the limit in which the scalar field
$\varphi$ is a small perturbation around a GR solution. Expanding
around the constant value $\varphi_{0}$ to first order in
$\hat\varphi\equiv\varphi-\varphi_{0}\ll1$, the field equations in the
Einstein frame~(\ref{eq:tensoreqnE}), (\ref{eq:scalareqnE}) read (see
e.g.~\cite{Yunes:2011aa})
\begin{align}
&G_{\mu\nu}^\star=8\pi T_{\mu\nu}^\star
\,, \label{Einsteinlin} \\
&\square^\star\hat\varphi=-4\pi \alpha_0 T^\star-4\pi \beta_0\hat\varphi T^\star\,. \label{KGlin}
\end{align}
Here we have assumed analyticity around $\varphi\sim \varphi_{0}$ and
used Eq.~\eqref{DEalphabeta}, where $A(\varphi)$ is the nonminimal
coupling to the matter fields in the Einstein frame, as defined by the
action~\eqref{STactionE}.
It is clear from Eq.~\eqref{KGlin} that $\alpha_0$ controls the
effective coupling between the scalar and matter. Various
observations, such as weak-gravity constraints and tests of violations
of the strong equivalence principle, require $\alpha_0$ to be
negligibly small when the scalar tends to its asymptotic value~\cite{Damour:1998jk,Damour:1996ke,Freire:2012mg}. This implies that a
configuration in which the scalar $\varphi\approx\varphi_0$ and $\alpha_0\approx 0$
should be at least an approximate solution in most viable
scalar-tensor theories. A detailed study
of the connection between the perturbative and nonperturbative scalarized
solutions can be found in~\cite{Stefanov:2008,Doneva:2010}.
With $\alpha_0=0$, any background GR solution solves the field
equations above at first order in the scalar field. At this order, the
Klein-Gordon equation reads
\begin{equation}
\left[\square^\star-\mu_s^2(x^\nu )\right]\hat\varphi=0\,,\qquad \mu_s^2(x^\nu)\equiv -4\pi\beta_0 T^\star\,. \label{effectivemass}
\end{equation}
Thus, the coupling of the scalar field to matter is equivalent to an
effective position-dependent mass. Depending on the sign of $\beta_0
T^\star$, the effective mass squared can be negative. Because $-T^\star\approx
\rho^\star>0$, this happens when $\beta_0<0$. When $\mu_s^2<0$ in a
sufficiently large region inside the NS, scalar perturbations of a GR
equilibrium solution develop a tachyonic instability (i.e., the
perturbations propagate superluminally, as particles with imaginary
mass). This instability is associated with an exponentially growing
mode, which causes the growth of scalar hair in a process akin to
ferromagnetism~\cite{Damour:1993hw,Damour:1996ke}.
Spherically symmetric NSs develop spontaneous scalarization for
$\beta_0\lesssim-4.35$~\cite{Harada:1998}. Detailed investigations of
stellar structure~\cite{Damour:1996ke,Salgado:1998sg}, numerical
simulations of collapse~\cite{Shibata:1994qd,Harada:1996wt,Novak:1997hw} and stability studies~\cite{Harada:1997mr,Harada:1998} confirmed that spontaneously
scalarized configurations would indeed be the end-state of stellar
collapse in these theories. In fact, spontaneously scalarized
configurations may also be the result of semiclassical vacuum
instabilities~\cite{Lima:2010na,Pani:2010vc,Mendes:2013ija,Landulfo:2014wra}.
The nonradial oscillation modes of spontaneously scalarized,
nonrotating stars were studied
in~\cite{Sotani:2004rq,Sotani:2005qx,Sotani:2014tua,Silva:2014ora}. The
bottom line of these studies is that the oscillation frequencies can
differ by a large amount from their GR counterparts if spontaneous
scalarization modifies the equilibrium properties of the star (e.g.,
the mass-radius relation) by appreciable amounts. However, current
binary pulsar observations yield very tight constraints on spontaneous
scalarization, and the oscillation modes of scalarized stars for
viable theory parameters are unlikely to differ from the corresponding
GR modes by any measurable amount.
\begin{figure*}[t]
\begin{center}
\begin{tabular}{ll}
\includegraphics[width=0.5\textwidth]{MR_Jordan-eps-converted-to.pdf}&
\includegraphics[width=0.5\textwidth]{CM_Jordan-eps-converted-to.pdf}\\
\includegraphics[width=0.5\textwidth]{I-Q-M_Jordan-eps-converted-to.pdf}&
\includegraphics[width=0.5\textwidth]{lambda-lambdaR-M_Jordan-eps-converted-to.pdf}\\
\end{tabular}
\caption{NS configurations in GR (solid lines) and in two
scalar-tensor theories defined by Eq.~\eqref{STactionE} with
$A(\varphi)\equiv e^{\frac{1}{2}\beta_0 \varphi^2}$ and $V(\varphi)\equiv0$. Dashed
lines refer to $\beta_0=-4.5$, $\varphi_0^\infty/\sqrt{4\pi}=10^{-3}$; dash-dotted
lines refer to $\beta_0=-6$, $\varphi_0^\infty/\sqrt{4\pi}=10^{-3}$. Each panel
shows results for three different EOS models (\texttt{FPS},
\texttt{APR} and \texttt{MS1}).
Top-left panel, left inset: relation between the nonrotating mass
$M$ and the radius $R$ in the Einstein frame. Top-left panel, right inset:
relative mass correction $\delta M/M$ induced by rotation as a
function of the mass $M$ of a nonspinning star with the same central
energy density.
Top-right panel, left inset: scalar charge $\tilde{q}/M$ as a function
of $M$. Top-right panel, right inset: relative correction to the
scalar charge $\delta \tilde{q}/\tilde{q}$ induced by rotation as a
function of $M$.
Bottom-left panel: Jordan-frame moment of inertia $\tilde{I}$ (left
inset) and Jordan-frame quadrupole moment $\tilde{Q}$ (right inset)
as functions of $M$.
Bottom-right panel: Jordan-frame tidal ($\tilde{\lambda}$) and
rotational ($\tilde{\lambda}^{\rm rot}$) Love numbers as functions
of $M$. [From~\cite{Pani:2014jra}.]
\label{fig:NS_ST2}}
\end{center}
\end{figure*}
\paragraph{Spontaneous scalarization and quantum instabilities in scalar-tensor theories with a conformal coupling.}
An interesting class of scalar-tensor theories that has been
recently investigated in the context of NS physics is the following:
\begin{equation}
S=\frac{1}{16\pi}\int d^4x\sqrt{-g}\left[R
-2g^{\mu\nu}\varphi_{,\mu}\varphi_{,\nu}-\xi\,R\,\varphi^2\right]+S_{\rm perfect\,fluid}\,,
\end{equation}
where $\xi$ is the conformal coupling parameter. For $\xi =1/12$ the
scalar field equations are invariant under conformal transformations
($g_{\mu\nu}\rightarrow
\gamma^2g_{\mu\nu}\,,\varphi\rightarrow\gamma^{-1}\phi$), whereas for
$\xi=0$ one recovers the usual minimally coupled massless scalar.
The theory above can be obtained as a particular case of the
action~\eqref{STactionJ} after a field redefinition.
Lima, Matsas and Vanzella showed that the vacuum expectation value of
nonminimally coupled scalar fields can grow exponentially in
relativistic stars~\cite{Lima:2010na}. At the classical level, this
quantum instability can be interpreted in terms of the spontaneous
scalarization discussed above~\cite{Pani:2010vc}. The instability can
occur for both positive and negative values of $\xi$. When $\xi<0$ and
$|\xi|$ is large enough, the instability can occur even for Newtonian
stars. For a detailed analysis of the approach to the classical limit
and of the relation between the quantum and classical nature of the
final state, see~\cite{Mendes:2013ija,Landulfo:2014wra}.
\paragraph{Slowly rotating solutions.}
Spinning NSs at first order in the Hartle-Thorne slow-rotation
approximation were studied by Damour and
Esposito-Far\`ese~\cite{Damour:1996ke} and later by
Sotani~\cite{Sotani:2012eb}. At first order in rotation, the scalar
field only affects the moment of inertia, mass and radius of the
NS. Second-order calculations~\cite{Pani:2014jra} are necessary to
compute corrections to the spin-induced quadrupole moment, tidal and
rotational Love numbers, as well as higher-order corrections to the NS
mass and to the scalar charge. Figure~\ref{fig:NS_ST2} shows
representative examples of the properties of NSs in a scalar-tensor
theory with spontaneous scalarization at second order in the rotation
parameter.
\paragraph{Rapidly rotating solutions.}
Rapidly rotating NSs in scalar-tensor theories were recently
constructed in~\cite{Doneva:2013qva} by extending the {\tt RNS}
code~\cite{Stergioulas:2003yp}. The results shown in
Figure~\ref{fig:NS_ST_fast} illustrate that scalarization effects are
stronger for rapidly rotating stars, and deviations from GR are
sensibly larger for fast-spinning NSs. One of the reasons is that the
stress-energy tensor (which acts as a source for the scalar field)
gets a contribution from the rotational energy of the star. The
nontrivial scalar field has a strong effect also on the NS angular
momentum and moment of inertia, which can differ by as much as a
factor of two from their GR values at the Kepler limit~\cite{Doneva:2013qva}. In addition,
there exists a larger range of parameters for which scalarization
occurs, and the critical value of the coupling constant $\beta_0$
where a nontrivial scalar field can develop increases substantially.
For example, for the polytropic EOS considered in~\cite{Doneva:2013qva} the critical value of $\beta_0$ increases from
$\beta_0\gtrsim-4.35$ in the nonrotating case to $\beta_0\gtrsim-3.9$
for rapid rotation. For realistic EOSs scalarization can occur for
even larger $\beta_0$~\cite{Doneva:2014uma}. Binary pulsar
observations imply $\beta_0\gtrsim-4.5$ (see Section~\ref{sec:BP}).
For a marginally allowed $\beta_0$, nonrotating scalarized NSs would
not differ considerably from the GR solutions, whereas rapid rotation
can produce significant deviations that can potentially set even
stronger astrophysical constraints on scalar-tensor theories~\cite{Doneva:2014uma,Doneva:2014faa}. Other proposed mechanisms that
can amplify the effects of scalarization include anisotropy~\cite{Silva:2014fca} and ``dynamical scalarization'' for merging NSs
in scalar-tensor theories, that will be discussed in
Section~\ref{sec:CB}~\cite{Barausse:2012da,Palenzuela:2013hsa,Shibata:2013pra,Taniguchi:2014fqa}. In
the last stages before merger the rotational frequencies of each NS
may approach the Kepler limit.
\begin{figure}[ht!]
\centering
\begin{tabular}{cc}
\includegraphics[width=0.5\textwidth]{M_eps_-eps-converted-to.pdf}&
\includegraphics[width=0.5\textwidth]{M_R_-eps-converted-to.pdf}
\end{tabular}
\caption{The NS mass as a function of the central energy density (left
panel) and of the radius (right panel) for static sequences of NSs
(solid lines) and sequences of stars rotating at the mass-shedding
limit (dotted lines). The trivial solutions coincide with the GR
limit ($\beta=\beta_0=0$). For $\beta=\beta_0=-4.2$ nontrivial
solutions do not exist in the nonrotating
case. [From~\cite{Doneva:2013qva}.]}
\label{fig:NS_ST_fast}
\end{figure}
\subsection{f(R) theories} \label{NS_fofR}
In principle $f(R)$ theories can be mapped to a specific form of the
action in scalar-tensor theory~\cite{Sotiriou:2008rp,DeFelice:2010aj},
but this mapping involves subtleties and technicalities that justify a
separate discussion of NS solutions in metric $f(R)$ gravity. In fact
the literature on NS solutions in metric $f(R)$ gravity is quite
extensive, and it contains several apparently controversial claims
~\cite{Frolov:2008uf,Kobayashi:2008tq,Upadhye:2009kt,Babichev:2009td,Babichev:2009fi,Jaime:2010kn}.
The recent interest in $f(R)$ theories is due to their potential to
explain cosmological observations without introducing dark matter or
dark energy. In terms of compact objects, this means that one is
usually interested in matching the stellar interior to a de Sitter
metric with an effective cosmological constant
\begin{equation}
\Lambda_\text{eff}=R_{\rm dS}/4\,,\label{Lambdaeff}
\end{equation}
where $R_{\rm dS}$ is the curvature at the de Sitter point, and $R\to
R_{\rm dS}$ far from the star. The problem involves two completely
different density (or curvature) scales, because the central density
of a NS ($\rho_0\sim10^{14} \text{g cm}^{-3}$) is enormously larger
than the density associated to the cosmological constant
($\rho_\Lambda=\Lambda/(8\pi G)\sim 10^{-29}\text{g cm}^{-3}$):
$\rho_\Lambda/\rho_0 \sim10^{-43}\label{ratio_f(R)}$.
In practice, only much larger values
($\rho_\Lambda/\rho_0\sim10^{-10}-10^{-6}$) can be used in
numerical codes. This issue is not specific to $f(R)$ theories: it
would also arise in GR with a positive cosmological constant if one
tries to match a NS interior with a de Sitter exterior. In fact, the
large disparity in density (or curvature) scales is not a problem if
one assumes that the cosmological scale has no sensible influence on
local physics. In other words, one would expect local observables such
as the NS mass and radius to be insensitive to $\rho_\Lambda/\rho_0$,
as long as this ratio is small enough:
$\rho_\Lambda/\rho_0\sim10^{-10}$ (say) would be practically
indistinguishable from $\rho_\Lambda/\rho_0\sim10^{-43}$, except for
giving an unrealistically large cosmological constant.
Calculations of NS structure in $f(R)$ theory used different
approaches, reaching different conclusions on the very existence of
relativistic compact stars. Here we try to clarify some critical
issues in the literature, pointing the reader to the original
references for more details.
\paragraph{Singular potential.}
When $f(R)$ is reformulated as a scalar-tensor theory, the potential
for the scalar degree of freedom can, and in general will, be
singular~\cite{Frolov:2008uf}. The scalar-field
equation~\eqref{scalareqKM} can be recast in the form
\begin{equation}
\square\phi=V_{\rm KM}'(\phi)-{\cal F}\,,
\end{equation}
where a prime denotes a derivative with respect to $\phi$ and
${\cal F}=8\pi/3 (\rho-3P)$ plays the role of a matter-driven force
term. For various solutions describing late-time cosmology or compact
objects, the ``force'' ${\cal F}$ pushes the scalar field towards the
unprotected curvature singularity, which is at finite distance (in
field and energy space) from the equilibrium configuration. As
discussed below, this happens precisely for those $f(R)$ models which
are otherwise theoretically and observationally
viable~\cite{Frolov:2008uf}.
This singular character of the potential can cast doubts on the
viability of $f(R)$ gravity and on the existence of compact objects in
these theories. However, this premature conclusion depends on the
choice of a specific scalar-tensor formulation of $f(R)$ gravity.
Kobayashi and Maeda~\cite{Kobayashi:2008tq} reported that the field
equations inside relativistic stars are plagued by singularities, but
subsequent work~\cite{Upadhye:2009kt,Babichev:2009td} claimed that
such singularities were unphysical and due to numerical
instabilities. Indeed, the scalar field in the interior of compact
objects can be very close to the value that corresponds to the
singular potential, but does not necessarily need to end up in the
singularity. This makes the integration challenging from a numerical
point of view, but it does not necessarily imply a pathology in the
underlying theory, as we show below.
First of all, note that any $f(R)$ model that meets the minimum
requirements to satisfy Solar System constraints -- i.e., it satisfies
the requirements of Eq.~\eqref{limit_f} -- is such that the first and
second derivatives of the scalar potential defined
in~\cite{Kobayashi:2008tq,Upadhye:2009kt,Babichev:2009td} are {\em
divergent} in the $R\to\infty$ limit. This limit corresponds to a
\emph{finite} value of the scalar field and to a \emph{finite} value of the
potential at the singular point. In the following, we shall
generically denote the scalar degree of freedom (in the various formulations discussed in Section~\ref{subsec:f(R)}) by
$\Phi$ and the scalar potential by $V_\Phi$, whereas $\Phi_s$ denotes
the value of the scalar at the singular point,
i.e.~$\Phi_s=\Phi(R\to\infty)$. Let us parametrize the large-curvature
expansion by using the rather generic expression (see Eq.~\eqref{limit_f})
\begin{equation}
f(R)\sim R+R_c\left[a+b\left(\frac{R_c}{R}\right)^c +
d\log\left(\frac{R_c}{R}\right)\right]\,,\quad R\gg R_c\,, \label{expansion}
\end{equation}
where $(a,b,c,d)$ are dimensionless real constants, $c\geq0$,
$d\geq0$, $R_c$ is some curvature scale of the order of $R_{\rm dS}$,
and we have kept only the dominant terms in a large-curvature
expansion.
Eq.~\eqref{expansion} is a good approximation in the interior of a NS,
where the curvature is much larger than the cosmological curvature
$R_{\rm dS}$, and indeed most of the models considered in the
compact object literature belong to this class. Using
Eq.~\eqref{expansion}, it is straightforward to prove that
\begin{equation}
V_\Phi(R\gg R_c)\to R_c\left[{\rm const}+d\log\left(\frac{R_c}{R}\right)\right]
\end{equation}
up to a constant (that can be adjusted to eliminate the constant term
above). In conclusion, in the limit $\Phi\to\Phi_s$ the potential
$V_\Phi$ is finite if $d=0$ and diverges logarithmically if $d\neq0$,
but its derivative $V_\Phi'\to\infty$ in any case.
Therefore, the energy density needed to make the singularity energetically
accessible is roughly
\begin{equation}
V_\Phi'(\Phi\to\Phi_s)\sim R\,.
\end{equation}
In models in which corrections to GR are relatively small for
$R\gg R_c$, this quantity is parametrically of the same order as the
matter energy density $\rho_c$ in the interior of a NS. This can be
seen by taking the trace of the modified Einstein equations [see
discussion around Eq.~\eqref{Rmin} below]. This simple argument seems
to suggest that the singularity should be accessible, as discussed
in~\cite{Frolov:2008uf}. However, as we will see below, in $f(R)$
theories a subtle mechanism prevents such singularities to be
accessible, at least in various situations. The price to pay is that
numerical integrations for realistic values of the theory parameters
are extremely challenging.
To this end, it is important to remark that the singular behavior of the scalar
potential is not an \emph{intrinsic} ingredient of the theory, but
just a prerogative of specific formulations of $f(R)$ gravity.
Indeed, this severe problem does not arise in the approach developed
by Jaime et al.~\cite{Jaime:2010kn}, in which the potential defined in
Eq.~\eqref{Jaime2} is regular for finite curvature.
When $R\gg R_c$, using Eq.~\eqref{expansion} we get
\begin{equation}\label{VJPS}
V_{\rm JPS}(R)\sim \left\{\begin{array}{ll}
R_c^{3}\left({R}/{R_c}\right)^{c+4}\,, &d=0\\
R_c^{3}\left({R}/{R_c}\right)^4\,, &d\neq0
\end{array} \right.\,.
\end{equation}
Remarkably, in this formulation the singularity at $R\to\infty$ is protected by an \emph{infinite}
potential well for any viable model satisfying Eq.~\eqref{limit_f}.
Furthermore, dimensional arguments show that in this case the energy density needed
to reach the singularity is of order
\begin{equation}
V_{\rm JPS}'(R)R^{-1}\sim \left\{\begin{array}{ll}
R_c\left({\rho_c}/{R_c}\right)^{c+2}\,, &d=0\\
R_c\left({\rho_c}/{R_c}\right)^2\,, &d\neq0
\end{array} \right.\,,
\end{equation}
which is always much larger than the internal energy density of a
NS. In other words, in this formulation the singularity at infinity is protected by an infinite potential
barrier~\cite{Jaime:2010kn}, as in any well-behaved mechanical model (e.g.~the harmonic
oscillator).
\paragraph{Chameleon mechanism.}
The existence of singular scalar configurations accessible at finite
energies is potentially dangerous. As we just discussed, in the case
of $f(R)$ gravity it is not the theory itself to be potentially
problematic, but only some particular formulations of the theory.
Clearly, the viability of the theory cannot depend on the particular
formulation chosen in~\cite{Jaime:2010kn}, so there must exist a
mechanism that prevents singular behavior also in other formulations.
Indeed, even in these potentially ill-defined formulations, a
subtle~\emph{chameleon mechanism}~\cite{Khoury:2003aq,Gubser:2004uf}
keeps the scalar field away from the singularity. The chameleon
mechanism is related to the generation of an infinitely large mass
term of the scalar field, due to self-interactions and to interactions
with other matter fields (see e.g.~\cite{Upadhye:2009kt}).
At high curvature $dV_\Phi/d\Phi \sim R$ and $dV_\Phi^{\rm
eff}/d\Phi\to R+8\pi T$,
so that the effective potential has a minimum at
\begin{equation}
R_{\rm min}\sim -8\pi T\,,\label{Rmin}
\end{equation}
which corresponds to some $\Phi_{\rm min}=\Phi(R_{\rm min})$. For any
$f(R)$ gravity theory satisfying the viability
conditions~\eqref{limit_f}, assuming $T\sim{\rm const}$, the scalar
mass in the large-curvature limit reads
\begin{equation}\label{chameleon_R}
m^2_\Phi\equiv\left.\frac{d^2V_\Phi^{\rm eff}(\Phi)}{d\Phi^2}\right|_{\Phi=\Phi_{\rm min}} \sim \left\{\begin{array}{ll}
R_c\left({-8\pi T}/{R_c}\right)^{c+2}\,, & d=0\\
R_c\left({-8\pi T}/{R_c}\right)^2\,, & d\neq0
\end{array} \right.\,.
\end{equation}
It is crucial to realize that the dimensionless quantity $m_\Phi R^{-1/2}$
is an estimate of the ratio between the curvature lengthscale and the
Compton wavelength of the graviton; therefore, it
is proportional to the number of integration steps needed to resolve the
dynamics of the chameleon field in a region of approximately constant
curvature $R$~\cite{Upadhye:2009kt}. Evaluating this quantity inside a
NS, where $R\sim -8\pi T\sim\rho_c$, one obtains
\begin{equation}\label{chameleon_R2}
m_\Phi^2 \rho_c^{-1}\sim \left\{\begin{array}{ll}
\left(x_{\rm dS}{\rho_c}/{\Lambda_{\rm eff}}\right)^{c+1}\,, & d=0\\
\left(x_{\rm dS}{\rho_c}/{\Lambda_{\rm eff}}\right)\,, & d\neq0
\end{array} \right.
\,,
\end{equation}
where $x_{\rm dS}=R_{\rm dS}/R_c$. The larger the ratio
${\rho_c}/{\Lambda_{\rm eff}}$, the larger the effective chameleon
mass and the number of steps needed for the integration in the stellar
interior. For a realistic NS embedded in a de Sitter universe the
effective mass is extremely large in cosmological units, and so is the
number of integration steps.
Such a heavy field is challenging to treat numerically. Indeed, as
discussed by~\cite{Upadhye:2009kt}, Yukawa-like error modes grow as
$e^{m_\Phi r}/r$, and dominate if the Compton scale $m_\Phi^{-1}$
becomes much smaller than the computational domain. Hence, in the ``wrong'' formulations the
integration of the field equations becomes practically impossible in
realistic situations, but this does not imply that the solutions are singular.
Note also that the chameleon mass depends on powers of $x_{\rm dS}$,
which may be a large quantity, making the integration even more
challenging. For example, the bound $x_{\rm dS}\gtrsim10^3$ must be
imposed for a popular model (the simplest version of the Starobinsky model~\cite{Starobinsky:2007hu}, where $f(R)=R[1-\mu R_c R/(R_c^2+R^2)]$) to
satisfy local tests~\cite{Hu:2007nk,Upadhye:2009kt}. Curiously, this
fact is often overlooked and unrealistic values $x_{\rm dS}={\cal
O}(1)$ are commonly used (but see~\cite{Upadhye:2009kt} for an
exception).
In summary, the challenge of constructing compact models in $f(R)$
gravity depends to a large extent on the formulation adopted and, in many cases, is usually overcome by restricting the discussion to
unrealistic stars with $\rho_c\sim 10^2 \Lambda_{\rm eff}$, which is
much smaller than the realistic density one should use ($\rho_c\sim
10^{43} \Lambda_{\rm eff}$), and even in this case numerical
integrations can be very difficult.
The chameleon mechanism provides an elegant way to keep the scalar
field away from the singular point. In fact, it is remarkable that the
same nonlinearities that make the scalar potential singular also keep
the field away from the singularity. As Babichev and Langlois put it
in their paper, the chameleon mechanism forces the scalar field to
stay ``attached to a track very near a precipice, without falling into
it''~\cite{Babichev:2009fi}. The distance of the track from the
precipice can be tiny (as small as $10^{-43}$ in cosmological units),
so that it is practically impossible to follow the evolution of the
scalar field numerically without being contaminated by the nearby
singularity.
\paragraph{Multivalued potential.}
Another known and controversial issue with the approaches
by~\cite{Upadhye:2009kt,Babichev:2009td} is multivaluedness. When we
recast $f(R)$ theories as scalar-tensor theories, the scalar-field
potential can be
multi-valued~\cite{Frolov:2008uf,Kobayashi:2008tq,Jaime:2010kn}. This
conclusion is model-dependent, but it applies e.g.~to the Starobinsky
model (cf.~Figure~1 of~\cite{Frolov:2008uf}) and to any model for
which $f_R$ is not a monotonic function of $R$. In particular, the
cosmological branch of the Starobinsky potential is discontinuous at
the singular point, i.e.~as $\Phi\to\Phi_s$.
This problem is usually ignored on the grounds that stellar structure
calculations only refer to the structure of the potential around a
local minimum, and that possible multiple branches are harmless if the
entire cosmological evolution of the scalar field is confined within a
single-valued branch of the potential. While this is true,
multivaluedness might seem to make $f(R)$ theories less natural and attractive,
especially if the scalar field is extremely close to the singular and
discontinuous point of the potential (as in the interior of compact
objects).
However --~similarly to the issue with the singular scalar potential~-- also multivaluedness is
formulation-dependent, as pointed out in~\cite{Jaime:2010kn}. Indeed, it turns
out that the potential~\eqref{VJPS} is \emph{not}
multivalued. This is simply due to the fact that the ``true''
dynamical degree of freedom is the curvature $R$, not $f_R$, and no
inversion $R=R(\Phi)$ is needed in the approach
of~\cite{Jaime:2010kn}.
\paragraph{Main results.}
The mapping of $f(R)$ gravity to scalar-tensor theories -- both in the
Jordan and in the Einstein frame -- is plagued by several potentially
dangerous issues, including singularities and multivaluedness in the
scalar potential and a diverging effective mass for the field. These
issues are intertwined, and in fact they can be seen as ``features,''
as they are needed for the theoretical safety of the theory (an
example being the chameleon mechanism that keeps the scalar field away
from the singularity). The numerical challenges they introduce may
also serve as a motivation to develop more efficient integration
methods. These same issues make the study of compact objects in $f(R)$
gravity particularly difficult, especially for realistic
configurations
~\cite{Cooney:2009rr,Arapoglu:2010rz,Alavirad:2013paa,Astashenok:2013vza,Astashenok:2014pua,Yazadjiev:2014cza,Staykov:2014mwa,Yazadjiev:2015zia}.
Most studies of NS configurations in $f(R)$ gravity consider
\emph{perturbative} corrections to the Einstein-Hilbert action of the
form $f(R)=R+\epsilon f_1(R)$, with $\epsilon\ll1$~\cite{Cooney:2009rr,Astashenok:2013vza,Arapoglu:2010rz,Alavirad:2013paa,Astashenok:2014pua}. This
expansion is similar in spirit to the EFT approach discussed elsewhere
in this review, and it bypasses some of the difficulties listed
above. Some of these models~\cite{Astashenok:2013vza} predict NSs with
large compactness (NS radii as small as $\sim9$ km), which are
difficult to obtain in GR, even taking into account current
uncertainties in the EOS.
Yazadjiev et al.~\cite{Yazadjiev:2014cza} recently went beyond the
perturbative level constructing static equilibrium models of NSs in a
theory of the form $f(R)=R+\lambda R^2$, where the coupling $\lambda$
is not necessarily small. They found that deviations from GR are
comparable with the variations due to uncertainties in the EOS, even
for large values of $\lambda$. Subsequent work by Staykov et
al.~\cite{Staykov:2014mwa} extended the analysis to first order in the
slow-rotation approximation, finding that the NS moment of inertia can
be up to $30\%$ larger than its GR counterpart. This correction is
larger than that introduced by uncertainties in the EOS, and (in
principle) it can be used to break the EOS degeneracy
(cf.~Section~\ref{subsec:ILQ} below). Yazadjiev et
al.~\cite{Yazadjiev:2015zia} constructed rapidly rotating NSs in
nonperturbative $f(R)=R+\lambda R^2$ gravity. For fast rotation, the
maximum NS mass and moment of inertia can be up to $\sim 16\%$ and
$60\%$ larger than in GR, respectively. These corrections to the NS
properties are large enough that, if observed, they may be used to
constrain the parameter $\lambda$.
\subsection{Quadratic gravity}
\label{sec:NSs/quadratic-gravity}
A remarkable consequence of recent work is that, from the point of
view of the properties of compact objects, scalar-tensor theories and
quadratic gravity are ``orthogonal'' extensions of GR. Isolated BHs in
scalar-tensor theories are described by the Kerr solution, just like
in GR, whereas NS configurations can acquire a scalar charge through
spontaneous scalarization. On the contrary, BH solutions in quadratic
gravity theories are endowed with a scalar field that is supported by
the higher-order curvature terms, but a ``no-scalar-monopole-hair''
theorem holds for NSs in EdGB and dCS gravity (at least in the
\emph{perturbative} regime).
There is a simple heuristic proof of the NS no-hair theorem in
quadratic gravity~\cite{Yagi:2011xp}. Integrating the scalar equation
of motion in EdGB gravity (within the perturbative EFT expansion)
yields
\begin{equation}
\int d^4x \sqrt{-g}\, \square\phi \propto \int d^4x\sqrt{-g} R_{\rm GB}^2 \,,
\label{proofNSs}
\end{equation}
where $R_{\rm GB}^2$ is the Gauss-Bonnet term (a similar conclusion applies to
dCS gravity, as long as we replace $R_{\rm GB}^2$ with the Pontryagin
density ${}^*RR$). Because $R_{\rm GB}^2$ (and ${}^*RR$) are
topological invariants, the right-hand side of Eq.~\eqref{proofNSs}
vanishes identically for any simply connected, asymptotically flat
domain. Furthermore isolated NSs are stationary, so the time
dependence of the left-hand side can be neglected, yielding an
integral over the volume. Finally, Stokes' theorem yields
\begin{equation}
\int \sqrt{-g} (\partial_i\phi) n^i dS = \int \sqrt{-g}(\partial_r \phi)
dS=0\,,
\label{StokesQuad}
\end{equation}
where the unit vector $n^i$ is normal to the surface $S$, taken to be
a 2-sphere at infinity. In order to have finite energy, the scalar
field must go to zero at large distances ($r\to \infty$). If we had
$\phi\to Q/r$, where $Q$ is a constant related to some hypothetical
scalar charge, Eq.~\eqref{StokesQuad} would imply $Q=0$. Therefore the
scalar field must decay faster than $1/r$ at infinity, and isolated
NSs in EdGB gravity and dCS gravity have no scalar monopole charge.
\subsubsection{EdGB theory}
The proof given above is valid only in the \emph{perturbative} regime,
i.e.~when the $R_{\rm GB}^2$ or ${}^*RR$ terms in the action are
coupled \emph{linearly} to the scalar field, so that the right-hand
side of Eq.~\eqref{proofNSs} vanishes identically. If we consider the
EdGB action of Eq.~\eqref{EdGBaction} with a generic coupling to the
Gauss-Bonnet term, ``baldness'' does not necessarily apply to NS
solutions. Furthermore, even in the perturbative regime the proof does
not exclude the possibility that NSs may be endowed with a nontrivial
scalar field profile that just happens to have vanishing scalar
monopole charge. Indeed, it is easy to check that the source term of
the scalar field equation does not vanish in general ($R_{\rm
GB}^2\neq0$): a constant (or zero) scalar field is not a solution of
the field equations, and therefore we would expect the NS properties
to be affected by a nontrivial scalar field.
These modifications were studied in~\cite{Pani:2011xm} for the case of
standard EdGB gravity with an exponential coupling of the form
\begin{equation}
f_1(\phi)=\alpha_{\rm GB} e^{\frac{\beta}{\sqrt{8\pi}}\phi}\,
\end{equation}
(in the notation of~\eqref{EdGBaction}), where $\alpha_{\rm GB}$ and
$\beta$ are coupling constants. For $\beta/\sqrt{8\pi}=2$ the model
reduces to the bosonic sector of heterotic string theory~\cite{Gross:1986mw}.
\begin{figure*}[htb]
\begin{center}
\begin{tabular}{cc}
\includegraphics[width=0.5\textwidth]{NSsEDGBa-eps-converted-to.pdf}&
\includegraphics[width=0.5\textwidth]{NSsEDGBb-eps-converted-to.pdf}\\
\includegraphics[width=0.5\textwidth]{NSsEDGBc-eps-converted-to.pdf}&
\includegraphics[width=0.5\textwidth]{NSsEDGBd-eps-converted-to.pdf}
\end{tabular}
\caption{Compact star models in EdGB gravity for different values of
the parameters $\alpha$ (denoted by $\alpha_{\rm GB}$ in the text)
and $\beta$, using the APR EOS. The bottom-left panel shows the NS
binding energy as a function of its central energy density. The
bottom-right panel displays the moment of inertia as a function of
the NS mass, together with the observation of a NS with $M\approx
2M_\odot$~\cite{Demorest:2010bx} (see
also~\cite{Antoniadis:2013pzd}) and a putative future observation of
the moment of inertia in agreement with GR within
$10\%$~\cite{Lattimer:2004nj}. Curves terminate when the
condition~\eqref{alpha_max} is not
fulfilled. [From~\cite{Pani:2011xm}.]
\label{fig:NSsEdGB}}
\end{center}
\end{figure*}
\paragraph{Static and slowly rotating solutions.}
Some properties of static and rotating solutions (computed at first
order in the slow-rotation approximation) are shown in
Figure~\ref{fig:NSsEdGB}.
Regardless of the EOS and for any value of $\alpha_{\rm GB}$, the
coupling to the dilaton tends to {\em reduce} the importance of
relativistic effects: this is again in contrast with the case of
scalar-tensor theory, where spontaneous scalarization increases the
relevance of relativistic effects (cf.~e.g.~the mass-radius curves in
Figure~\ref{fig:NS_ST2}). This trend is confirmed in the left panel of
Figure~\ref{fig:NSsEdGB2}, showing that the maximum gravitational mass
$M_{\rm max}$ decreases monotonically as a function of the product $\alpha_{\rm GB}\beta$ of the EdGB
coupling parameters.
\begin{figure}[htb]
\begin{center}
\begin{tabular}{cc}
\includegraphics[width=0.5\textwidth]{NSsEDGB2a-eps-converted-to.pdf}&
\includegraphics[width=0.5\textwidth]{NSsEDGB2b-eps-converted-to.pdf}
\end{tabular}
\caption{Left panel: Maximum mass as a function of the product
$\alpha_{\rm GB}\beta$ of the EdGB coupling parameters, for
different EOS models and in the nonrotating case . To the left of
the filled circle, this maximum mass corresponds to the radial
stability criterion; to the right, it corresponds to the maximum
central density for which we can construct static equilibrium
models. The maximum observed NS mass $M\approx2 M_\odot$ is marked
by a horizontal
line~\cite{Demorest:2010bx,Antoniadis:2013pzd}. Right panel:
Exclusion plot for the EdGB coupling in the small-field limit and
for nonrotating models. In the region above the dotted lines no
compact star solution can be constructed
(cf.~Eq.~\eqref{alpha_max}). In the region above the thick lines
(marked as ``RI,'' Radial Instability), static configurations are
unstable against radial perturbations. Markers indicate the maximum
central density of radially stable stars in
GR. [From~\cite{Pani:2011xm}.]
\label{fig:NSsEdGB2}}
\end{center}
\end{figure}
Thus in EdGB gravity -- as in GR -- soft EOS models should be ruled
out by observations of high-mass NSs. As we will see, similar
conclusions apply to other theories.
An interesting feature of EdGB gravity is that, for fixed values of
$\alpha_{\rm GB}$ and $\beta$, there exists a constraint on the
central density $\rho_c$ (or central pressure $P_c$) that allows for
the existence of NSs. In the small-field limit, the constraint
reads~\cite{Pani:2011xm}
\begin{equation}
\label{alpha_max}
\begin{split}
\alpha_{\rm GB}^2\beta^2<\frac{1}{7776 \pi P_c^4 \rho_c}
\Bigg[&128 \rho_c^3-27 P_c^2 \rho_c+288 P_c \rho_c^2+54 P_c^3 \\
&{}-2\sqrt{ (3 P_c+\rho_c)
\left(3 P_c-8 \rho_c\right)^2 (3 P_c+4 \rho_c)^3}\Bigg].
\end{split}
\end{equation}
If we fix $\alpha_{\rm GB}\beta$, the condition above implies that no
NSs exist above some critical maximum central density.
Quite interestingly, the requirement that the theory should support a
maximum mass $M_{\rm max}$ larger than some fiducial observational
value can place rather stringent upper bounds on the EdGB
coupling. Under rather mild assumptions, Pani et
al.~\cite{Pani:2011xm} estimated that $\alpha_{\rm GB}\lesssim {\cal
O}(10) M_\odot^2$, the precise number depending on the EOS and on
the value of $\beta$. This bound is slightly more stringent than the
purely theoretical bound that results from requiring the existence of
BHs in the theory~\cite{Kanti:1995vq,Pani:2009wy}.
Unfortunately, future observations of the moment of inertia are
unlikely to place even tighter bounds on the theory. This is because
deviations of the moment of inertia from its GR value are at most
$\sim 5\%$, at least in the slow-rotation limit~\cite{Pani:2011xm},
while the precision of future observations is expected to be $\sim
10\%$ in optimistic scenarios~\cite{Lattimer:2004nj}. At least for
EdGB gravity, we expect the most stringent constraints to come from
mass measurements, rather than from measurements of the moment of
inertia.
\begin{figure}[t!]
\begin{center}
\begin{tabular}{cc}
\includegraphics[width=0.7\textwidth, angle=0, clip=true]{KKMfig2a-eps-converted-to.pdf}
\end{tabular}
\end{center}
\caption{The quadrupole moment $Q$ (in units of $M_\odot \cdot {\rm
km}^2$) is shown versus the angular velocity $\Omega$ (in Hz) for
the scaled angular momentum $j=0.4$ and EdGB couplings $\alpha\equiv
\alpha_{\rm GB}/M_\odot^2=0$, 1 and 2~\cite{Kleihaus:2014lba}. The
upper and lower sets of curves refer to the polytropic EOS
DI-II~\cite{Diaz-Alonso:1985} and to an approximation of EOS
FPS~\cite{Haensel:2004nu}, respectively.}
\label{KKMfig2}
\end{figure}
\paragraph{Rapidly rotating solutions.}
The previous analysis was limited to small (first-order) corrections
in the NS spin. Rapidly rotating NSs in EdGB gravity were considered
by Kleihaus et al.~\cite{Kleihaus:2014lba}.
Figure~\ref{KKMfig2} shows the quadrupole moment $Q$ in units of
$M_\odot \cdot {\rm km}^2$ for NSs at fixed angular momentum ($j\equiv
J/M^2=0.4$) versus the angular velocity $\Omega$. The EdGB results
with coupling
$\alpha_{\rm GB}/M_\odot^2=1,2$ are compared to the case of GR ($\alpha_{\rm GB}=0$) for two
EOSs: the polytropic EOS from~\cite{Diaz-Alonso:1985} (DI-II) and the
FPS EOS~\cite{Haensel:2004nu} (fitted by a polytrope).
The quadrupole moment is more sensitive to the EOS than to the Gauss-Bonnet
coupling $\alpha_{\rm GB}$, so that even a putative measurement of the
quadrupole moment of a fast-spinning NS
can not be used to constrain the theory.
\begin{figure}[tb]
\begin{center}
\begin{tabular}{c}
\includegraphics[width=0.7\textwidth,clip=true]{DI_full.pdf}\\
\includegraphics[width=0.7\textwidth,clip=true]{Q-M-APR-Dup-eps-converted-to.pdf}
\end{tabular}
\caption{\label{fig:NSsCS} Properties of slowly rotating NSs in dCS
gravity.
The CS coupling $\zeta$ is related to $\alpha_{\rm CS}$, introduced in
Eq.~\eqref{CSaction}, by $\zeta\equiv 4 \sqrt{2} \alpha_{\rm CS}
M/R^3$ in the top panel, and $\zeta \equiv 4 \alpha_{CS}^2 M^2/R^6$
in the bottom panels.
Top panel: change in the moment of inertia induced by the CS
modification as a function of the CS coupling for two different EOSs
and two different values of the NS compactness. The dashed lines show
the analytic result for a nonrelativistic constant-density
star~\cite{AliHaimoud:2011fw}.
Bottom panels: numerical (symbols) and fitted (curve) results for the
dimensionless scalar dipole charge $\bar{\mu}$ (top left), the
quadrupole correction $Q$ (top right), the mass shift $\delta M$
(bottom left), and the angular momentum shift $\delta J$ (bottom
right), as functions of the NS mass $M_\star$. The y-axes are rescaled as explained in~\cite{Yagi:2013mbt}.
[From~\cite{AliHaimoud:2011fw,Yagi:2013mbt}].
}
\end{center}
\end{figure}
\subsubsection{dCS theory}\label{sec:ns-dcs}
The CS term does not introduce any modification to GR for spherically
symmetric configuration: isolated NSs differ from their GR
counterparts only when they are rotating. So far, rotating solutions
have been computed only in the slow-rotation approximation. The
``baldness theorem'' discussed above implies that the scalar monopole
will vanish, but NSs can still support a nontrivial scalar field. In
fact, spinning NSs have a {\em scalar dipole} ``hair,'' which modifies
the geometry and the properties of the star. The CS correction to the
NS moment of inertia was calculated
in~\cite{AliHaimoud:2011fw,Yunes:2009ch} (at first order in slow
rotation). Calculations of the CS correction to the NS quadrupole
moment require going to second order in rotation, and they can be
found in~\cite{Yagi:2013mbt}.
A summary of the main results is presented in
Figure~\ref{fig:NSsCS}. The plot shows: (1) the CS corrections to the
moment of inertia, (2) the scalar dipole susceptibility (which is
related to the NS sensitivity, discussed in
Section~\ref{sec:sensitivities} below), (3) the CS corrections to the
quadrupole moment and (4) the mass and angular momentum shifts induced
by rotation at second order in the spin. The mass shift is always
negative, while the quadrupole moment deformation is always positive.
\subsection{Lorentz-violating theories}
\label{sec:NS_LV}
\input{sec-compobj/NS_LV}
\subsection{Massive gravity and Galileons}
There are very few phenomenological studies of NSs in massive
gravity. This is probably due to the technical difficulties related to
the Vainshtein mechanism: cf.~Section~\ref{subsec:massgrav}, and~\cite{Babichev:2013usa} for a review. In a nutshell, the helicity-0
graviton mode becomes strongly coupled at distances smaller than the
so-called ``Vainshtein radius,'' a characteristic length scale that
depends on both the theory parameters and the source. The Vainshtein
effect resolves the vDVZ discontinuity afflicting Fierz-Pauli theory,
but the presence of a new scale (which might be parametrically larger
than the stellar radius and is much smaller than the Compton
wavelength of the graviton) makes it difficult to obtain nonvacuum
solutions without resorting to approximations.
Before the recent developments in nonlinear, ghost-free massive
gravity~\cite{deRham:2010kj}, Damour et al.~\cite{Damour:2002gp}
reconsidered the vDVZ discontinuity problem and its possible
resolution through Vainshtein's nonlinear resummation of nonlinear
effects. As part of this study, the authors investigated the viability
of spherically symmetric stars in a nonlinear version of Fierz-Pauli
theory. They found that some solutions show physical singularities,
but also that there exist regular solutions interpolating between a
modified GR interior and a de Sitter exterior, with curvature
proportional to the square of the graviton mass.
Another relevant study in this context was performed by Babichev et
al., who considered the problem of recovering GR from the decoupling
limit of the theory in the case of static, spherically symmetric
sources~\cite{Babichev:2009jt,Babichev:2010jd}.
To the best of our knowledge, the only attempt to construct static
stellar configurations in full dRGT massive gravity was carried out by
Gruzinov and Mirbabayi~\cite{Gruzinov:2011mm}, but phenomenological
studies of observational constraints (including stellar rotation and
realistic EOSs) are still lacking.
In the context of the cubic Galileon model
(cf. Section~\ref{subsec:massgrav}), in which the Vainshtein mechanism
suppresses the scalar field interactions with matter,
Ref.~\cite{Chagoya:2014fza} studied nonrelativistic, slowly rotating
stars and static relativistic stars, finding that deviations from GR
are suppressed at high densities.
Spherically-symmetric gravitational collapse in Galileon theories was studied
in~\cite{Bellini:2012qn,Barreira:2013xea}.
\subsection{Gravity with auxiliary fields}\label{sec:NS_auxiliary}
Gravitational theories with auxiliary fields are equivalent to GR in
vacuum and only differ from GR in their coupling to matter. Therefore
these theories may look like the prototypical example of modified
theories of gravity whose phenomenology can be explored using NSs, but
not BHs. While this is true, it turns out that it is quite difficult
to put observational constraints on these theories, due to a severe
degeneracy between the nuclear matter EOS and beyond-GR
effects. Furthermore these theories do not violate the strong
equivalence principle, and NSs do not acquire extra charges that could
leave an imprint in binary dynamics. For this reason, the very concept
of ``sensitivity'' (discussed in Section~\ref{sec:sensitivities}
below) is meaningless in these theories. We will now briefly review
the literature on NSs in various theories with auxiliary fields.
\paragraph{Palatini $f({\cal R})$.}
The Palatini formulation of $f({\cal R})$ gravity has been
investigated in detail in other respects, but the literature on NS
solutions is not very extensive. Most studies dealt with the problem
of curvature singularities (a problem shared with EiBI gravity: see Refs.~\cite{Kainulainen:2006wz,Kainulainen:2007bt,Barausse:2007pn,Barausse:2007ys,Olmo:2008pv} and
below for a unified discussion), but there is no detailed analysis of
NS properties in Palatini gravity and of possible strong-field
tests. It is reasonable to expect that most of the phenomenology
should be at least qualitatively similar to EiBI theory, reviewed
below. This is confirmed by the findings of Pani et
al.~\cite{Pani:2013qfa}, who showed that Palatini $f({\cal R})$
gravity and EiBI gravity are only two representative examples of a
{\em single} class of theories that modify GR by adding nondynamical
fields.
\paragraph{Eddington-inspired Born-Infeld gravity.}
Static and slowly rotating NSs in EiBI gravity were constructed
in~\cite{Pani:2011mg}, and their phenomenology was studied
in~\cite{Pani:2012qb,Harko:2013wka,Sham:2013sya}. For a given EOS, the
maximum mass of equilibrium NS configurations can be twice as large as
the corresponding GR model for experimentally viable values of the
EiBI parameter. An example of NS properties for a piecewise polytropic
EOS~\cite{Read:2008iy} that reproduces the FPS EOS is shown in Figure~\ref{fig:NSsEiBI} (cf.~\cite{Pani:2012qb} for
details).
\begin{figure*}[tb]
\begin{center}
\begin{tabular}{cc}
\includegraphics[width=0.5\textwidth]{NSsEiBI1-eps-converted-to.pdf}
\includegraphics[width=0.5\textwidth]{NSsEiBI2-eps-converted-to.pdf}\\
\end{tabular}
\includegraphics[width=0.5\textwidth]{NSsEiBI3-eps-converted-to.pdf}
\caption{Compact stars in EiBI theory constructed using the FPS EOS
and different values of the EiBI parameter $\kappa$
[cf.~Eq.~\eqref{actionEiBI}]. Top-left panel: mass as a function of
the central baryonic density $\rho_b$ (inset: binding energy as a
function of $\rho_b$). Top-right panel: mass-radius
relation. Bottom panel: moment of inertia as a function of the
stellar mass. The central density and coupling constant are
normalized by the typical density of nuclear matter,
$\rho_0=8\times 10^{17}$~kg\ m$^{-3}$. Curves terminate when either
condition~\eqref{lim1} or \eqref{lim2} are not fulfilled, and
self-gravitating objects can not exist. In the top-right panel, a
horizontal band corresponds to the maximum observed NS mass
$M\simeq 2M_\odot$~\cite{Demorest:2010bx,Antoniadis:2013pzd},
whereas the shaded region ($R\gtrsim 2.9 M$) is excluded by
causality~\cite{Lattimer:2006xb}. [Adapted
from~\cite{Pani:2012qb}.]
\label{fig:NSsEiBI}}
\end{center}
\end{figure*}
Within GR, the FPS EOS is ruled out by observations of NSs with masses
$M=(1.97\pm0.04)M_\odot$~\cite{Demorest:2010bx} (horizontal band in
Figure~\ref{fig:NSsEiBI}) and
$M=(2.01\pm0.04)M_\odot$~\cite{Antoniadis:2013pzd}. In EiBI gravity
the maximum mass of a NS can be much larger than in GR, and
observations of high-mass NSs can be accommodated without invoking a
stiffer EOS.
An interesting property of EiBI gravity is that, for a given value of
$\kappa$, no self-gravitating objects can exist above some critical
central density $\rho_c$ (or pressure $P_c$). More specifically, one
must have
\begin{align}
P_c\kappa<1\,&\quad\text{for $\kappa>0$} \,,\label{lim1}\\
\rho_c|\kappa|<1\,&\quad\text{for $\kappa<0$}\,,\label{lim2}
\end{align}
where $P_c$ and $\rho_c$ are the central pressure and density,
respectively~\cite{Pani:2011mg}. Assuming that NSs can reach central
densities $\rho_c\sim10^{18}$~kg\ m$^{-3}$ and $P_c\sim
10^{34}$~N\ m$^{-2}$, these bounds would constrain the theory,
yielding $|\kappa|\lesssim 1 \mbox{ m}^5\mbox{kg}^{-1}\mbox{s}^{-2}$.
Avelino~\cite{Avelino:2012ge} derived even stronger constraints,
$|\kappa|<R^2$, from the existence of a self-gravitating astrophysical
objects of size $R$. For a typical NS this translates into the bound
\begin{equation}
|\kappa|\lesssim 10^{-2} \mbox{ m}^5\mbox{kg}^{-1}\mbox{s}^{-2}\,.
\end{equation}
However these constraints are only indicative, because they are based
on the (untestable) assumption that matter in the NS core reaches
nuclear densities in EiBI gravity (as it does in GR). Furthermore,
there is no constraint from causality (cf.~the shaded region in the
top-right panel of Figure~\ref{fig:NSsEiBI}) because causality
constraints are always satisfied, even for large values of
$\kappa$. This is due to the existence of a maximum compactness
($M/R\lesssim0.3$) and it can be understood by looking at the theory's
effective stress-energy tensor~\cite{Delsate:2012ky}.
Sham et al.~\cite{Sham:2012qi} studied the radial stability of
relativistic stars in EiBI. Sotani~\cite{Sotani:2014xoa} investigated
nonradial oscillations in the Cowling approximation, finding that the
observation of the fundamental mode frequency may be used to
distinguish EiBI gravity from GR if the coupling is sufficiently
large. Because of the peculiar nonlinear coupling to matter, exotic
star-like solutions (such as pressureless
stars~\cite{Pani:2011mg,Pani:2012qb}, wormholes~\cite{Harko:2013aya}
and geons~\cite{Olmo:2013gqa}) exist in this theory. Furthermore (for
$\kappa>0$) the Chandrasekhar limit $M\lesssim 1.4 M_\odot$ on the
mass of white dwarfs is replaced by a minimum radius
condition~\cite{Pani:2012qb}: $R_{\rm WD}>\sqrt{{3\kappa}/({16\pi})}$.
Pani et al.~\cite{Pani:2012qb} studied nonrelativistic stellar
collapse in EiBI, finding that for any $\kappa>0$ the final state of
the collapse is a pressureless star (instead of a singularity, as in
GR). No relativistic simulations of collapse in the full theory are
available to date.
\paragraph{Degeneracy between NS EOS and nonlinear matter coupling.}
The degeneracy between beyond-GR effects and uncertainties in the NS
EOS is an intrinsic limitation in our ability to carry out precision
tests of gravity with NS observations. Uncertainties in the EOS often
translate into uncertainties in macroscopic observables (such as
masses, radii or oscillation frequencies) which are larger than
putative deviations from GR. This is usually the case for theories
(such as scalar-tensor theory) that are well constrained in the
weak-field limit.
Because gravitational theories with auxiliary fields are essentially
unconstrained in the weak-field limit, and they only modify GR in
their coupling to matter, NSs may look like ideal laboratories to
constrain them. Unfortunately, in this case GR modifications are
``maximally degenerate'' with EOS uncertainties. The reason is that
these theories do not contain extra dynamical fields, so the
right-hand side of the gravitational field
equations~\eqref{eqAuxiliary} can be interpreted as the stress-energy
tensor of an ``effective'' fluid with a contrived EOS. In particular,
Delsate et al.~\cite{Delsate:2012ky} proved analytically that EiBI
gravity coupled to a perfect fluid is \emph{equivalent} to GR with an
``effective'' perfect fluid stress-energy tensor. If the original
$T_{\mu\nu}$ satisfies the energy conditions usually imposed in GR,
the same is not necessarily true for the effective stress-energy
tensor. For this reason the singularity theorems of GR do not
apply~\cite{Delsate:2012ky}, and theories with auxiliary fields have
singularity-avoidance properties. The intrinsic ambiguity between
modifications in the gravitational coupling and variations in the EOS
makes it difficult (if not conceptually impossible) to constrain these
theories.
\paragraph{Curvature singularities.}
Palatini $f({\cal R})$ gravity and EiBI gravity are experimentally
viable theories passing all weak-field tests.
However, it has been pointed out that their peculiar coupling with
matter might lead to the appearance of curvature singularities at the
surface of macroscopic objects, where large gradients are
present\footnote{As
discussed in Section~\ref{subsec:auxiliary}, Palatini $f({\cal R})$
gravity may be affected by other pathologies, including potential
conflicts with the Standard Model~\cite{Flanagan:2003rb} and issues
with averaging in cosmology~\cite{Flanagan:2003rb,Li:2008fa}. These
problems are still debated (cf.~\cite{Olmo:2011uz} for a review),
but they cast serious doubts on the viability of this class of
theories.}~\cite{Barausse:2007pn,Pani:2012qd,Pani:2013qfa}. The root of the
problem lies in the fact that
Eq.~\eqref{eqAuxiliary} contains third-order derivatives of the matter
fields. This is in contrast to GR, where at most first derivatives of
the matter fields appear on the right-hand side of Einstein's
equations.
This different structure is also evident in the Newtonian limit of the
theory. The solution of the modified Poisson
equation~\eqref{PoissonEiBI}, $\Phi=\Phi_{\rm N}+2\pi {\kappa}\rho$,
shows that the gravitational potential $\Phi$ is algebraically related
to~$\rho$.
Any matter configuration which is discontinuous or just not smooth
enough will produce discontinuities in the metric, as well as
singularities in the curvature invariants that depend on the second
derivatives of~$\Phi$, and ultimately lead to unacceptable
phenomenology~\cite{Barausse:2007pn,Barausse:2007ys,Barausse:2008nm,Pani:2012qd}. For
example, the Ricci curvature $R$ of a self-gravitating barotropic
perfect fluid would depend on the second derivatives of the pressure
field, whereas in GR it simply reads $R=8\pi(\rho-3P)$. If the
function $P(r)$ is continuous but not differentiable at the stellar
surface, then $P'(r)$ would be discontinuous at the radius and
$P''(r)$ would introduce an unacceptable Dirac-delta contribution to
the curvature.
This was shown explicitly in~\cite{Barausse:2007pn,Pani:2012qd} by
assuming a polytropic EOS near the stellar surface: for rather
standard values of the polytropic index, the scalar curvature is
actually \emph{divergent} near the NS radius due to the presence of
higher-order derivatives of the matter fields.
Strong deviations from GR are expected when the curvature becomes
unbound near the stellar surface. For example, surface singularities
would give rise to tidal forces which can be orders of magnitude
larger than in Einstein's theory. This and other consequences of
curvature singularities in theories with auxiliary fields are
discussed in~\cite{Barausse:2007pn,Barausse:2007ys,Barausse:2008nm,Pani:2012qd,Pani:2013qfa}.
The appearance of curvature singularities may look like a fatal blow
for these theories, but there is some controversy surrounding this
issue. Kainulainen et al.~\cite{Kainulainen:2007bt} criticized the
analysis of Palatini $f({\cal R})$ gravity carried out
in~\cite{Barausse:2007pn}, whereas
Refs.~\cite{Barausse:2007ys,Barausse:2008nm} argued that the original
analysis is essentially correct. Olmo~\cite{Olmo:2008pv} showed that
in a prototypical Palatini theory with $f({\cal R})={\cal R}+\lambda
{\cal R}^2$, where $\lambda$ is of the order of the Planck length
squared, the curvature invariant grows only at extremely small
densities, so that the theory is practically viable (but in this case
the macroscopic properties of NSs are indistinguishable from their GR
counterparts, and no interesting phenomenology can be probed with
astrophysical observations).
Finally, Kim~\cite{Kim:2013nna} pointed out that strong tidal forces
near the stellar surface can modify the effective EOS in such a way
that curvature singularities are removed. This result suggests that
the gravitational backreaction on the matter dynamics can modify the
effective description of the fluid. Further analysis is needed to test
the generality of this conclusion and the viability of these theories.
\subsection{Strong-field tests of gravity with universal relations in NSs and quark stars}\label{subsec:ILQ}
\input{sec-compobj/univ-rel}
\subsection{Neutron star sensitivities in modified gravity}
\label{sec:sensitivities}
In preparation for our discussion of the dynamics of compact binaries,
that will be the main topic of Chapter~\ref{sec:CB}, we conclude this
chapter with a brief introduction to the so-called ``sensitivities''
of extended self-gravitating objects in modified gravity.
In many extensions of GR the strong equivalence principle is violated
due to the presence of additional fields that mediate the
gravitational interaction. Self-gravitating objects (and in
particular compact objects) are not test particles, and violations of
the strong equivalence principle imply that the local value of the
gravitational constant depends on the additional dynamical
field(s). When self-gravitating bodies move in regions of spacetime
where the extra fields are not constant, their internal gravitational
energy, and therefore their total mass-energy, will change.
This effect can be described by endowing extended bodies in modified
gravity with a macroscopic property (the ``sensitivity'') measuring
how the body's internal energy depends on the additional field(s)~\cite{Eardley:1975}. The sensitivity can be expected to be larger (and
more relevant for, e.g., binary dynamics) for compact objects, such as
NSs, because their self-gravity is stronger.
\paragraph{Scalar-tensor theories.}
How can we represent the stress-energy tensor for an extended object
in scalar-tensor theories? Eardley~\cite{Eardley:1975} showed that it
can be modeled as a normal point-particle stress-energy tensor, as
long as we promote the (constant) mass of each body to a function of
$\phi$: $M_{\scriptscriptstyle A} = M_{\scriptscriptstyle A}(\phi)$.
More precisely, in order to describe the orbital dynamics of a widely
separated binary system perturbatively in the size-to-separation
ratio, it is useful to integrate out length scales smaller than the
size of the bodies. We obtain an EFT in which the bodies are
represented by point particles, whose dynamics are governed by an
effective matter action
\begin{equation}\label{sctens_pp_action}
S_{\rm m} = \sum_A
\int_{\Gamma_{\scriptscriptstyle A}} \mathcal{L}_{\scriptscriptstyle A} ds_{\scriptscriptstyle A} \,.
\end{equation}
Here $A$ is an index that labels the bodies,
$\Gamma_{\scriptscriptstyle A}$ is the world-line of body $A$, and
$ds_{\scriptscriptstyle A}$ is the differential arclength
along it. The effective Lagrangian $\mathcal{L}_{\scriptscriptstyle
A}$ encodes information about the internal structure of body $A$,
and admits a derivative expansion
\begin{equation}
\mathcal{L}_{\scriptscriptstyle A} = -M_{\scriptscriptstyle A}(\phi) + \dots \,,
\end{equation}
where the leading-order term describes the dependence of the
mass-energy on the local scalar value, and subleading terms are
discussed in the appendix of~\cite{Damour:1998jk}. For bodies with
negligible gravitational binding energy $M_{\scriptscriptstyle
A}(\phi)$ is independent of $\phi$, and so
Eq.~(\ref{sctens_pp_action}) reduces to the action of a collection of
test particles, which move along geodesics of the Jordan-frame metric.
As will be discussed in Chapter~\ref{sec:CB}, the theoretical
predictions for the orbital dynamics and emission of radiation in
compact binaries depend on $M_{\scriptscriptstyle A}(\phi_0)$ and
successive derivatives $M_{\scriptscriptstyle A}^{(k)}(\phi_0)$, where
$\phi_0$ is the asymptotic value of $\phi$ far away from the binary
system. In particular, the sensitivity of body $A$ is defined as
\begin{equation}
s_A \equiv \left(\frac{d\ln M_{\scriptscriptstyle A}(\phi)}{d\ln \phi}\right)_{\phi=\phi_0} \,. \label{sensitivity}
\end{equation}
Higher-order derivatives of $M_{\scriptscriptstyle A}(\phi)$ are used
to define higher-order sensitivities, e.g.~$s'_{\scriptscriptstyle A}$
and $s''_{\scriptscriptstyle A}$.
The first detailed calculation of the NS sensitivities in
scalar-tensor theories was carried out by
Zaglauer~\cite{Zaglauer:1992bp}. The sensitivities are related to the
gravitational binding energy, and therefore they depend on the
microphysics of the specific self-gravitating object we consider. For
ordinary stars and for white dwarfs the sensitivities are of order
$\approx 10^{-6}$ and $\approx 10^{-4}$, respectively, because the
gravitational binding energy of these objects is small. For compact
objects (such as NSs and QSs) the sensitivities can be of the order of
$0.1$, and nonlinear phenomena (like spontaneous scalarization) can
even produce arbitrarily large sensitivities. A special case are BHs:
their mass in the Einstein frame is
constant~\cite{Damour:1992we,Jacobson:1999vr}, so BH sensitivities are
constant for a given scalar-tensor theory ($s_{\rm BH}=1/2$ in
Bergmann-Wagoner theories).
\paragraph{Quadratic gravity.}
Just like scalar-tensor theories, also quadratic gravity theories
generically violate the strong equivalence principle, and this
violation can be described in terms of sensitivities. Due to an
effective coupling between the matter fields and the scalar field, the
observable NS properties depend on the local value of the scalar field
near the object.
Yagi et al.~\cite{Yagi:2011xp,Yagi:2013mbt} computed the sensitivities
of compact objects in dCS theory. An important qualitative difference
with scalar-tensor theories consists in the fact that in quadratic
gravity BHs can carry a nontrivial scalar charge. BHs in dCS gravity
have hair (and a nontrivial sensitivity) only if they are spinning; in
EdGB gravity, this happens even for static BHs. On the other hand, the
sensitivity of nonspinning NSs in EdGB gravity is vanishing, as argued
at the beginning of Section~\ref{sec:NSs/quadratic-gravity}
(cf.~Appendix A of~\cite{Yagi:2011xp} for details). This is another
example of how scalar-tensor theories and quadratic gravity are in
some sense ``orthogonal'' in the context of the structure and dynamics
of compact objects.
The sensitivity (more precisely, the rescaled scalar dipole
charge $\bar{\mu}$: cf.~\cite{Yagi:2013mbt} for definitions
and more details) of spinning NSs in dCS theory is shown in the top
inset of the bottom-left panel of Figure~\ref{fig:NSsCS}. To the best of
our knowledge, the analogous calculation for spinning NSs in EdGB
gravity is not available in the literature at the time of writing.
\paragraph{Lorentz-violating theories.}
In Lorentz-violating theories the sensitivities (also known as \AE
ther or khronon ``charges'') characterize the amount of violation of
the strong equivalence principle due to the effective coupling between
matter and the \AE ther or khronon field in the strong-gravity
regime. Due to this coupling, the compact object's structure, binding
energy and gravitational mass will be functions of the motion relative
to the \AE ther or khronon. The action describing the motion of a
strongly gravitating body with gravitational mass $\tilde m$ is given,
in the point-particle approximation, by~\cite{Foster:2007gr}
\begin{equation}
\label{eq:actionpp}
S_{\rm pp} = - \int d\tau \; \tilde{m} (\gamma) = - \tilde{m} \int d\tau \;
\left\{1 + \sigma \left(1- \gamma \right) + {\cal{O}}\left[\left(1-\gamma \right)^{2}\right] \right\}\,,
\end{equation}
where $\gamma \equiv u_{\mu} U^{\mu}$ represents the Lorentz factor of
the body relative to the \AE ther, and $d\tau$ is the proper time
along the particle's trajectory. In the second equality, we performed
a slow-motion PN expansion with $\gamma \ll 1$ and $\tilde{m} \equiv
\tilde m (1)$. The sensitivity parameter $\sigma$ is defined by
\begin{equation}
\sigma \equiv - \left.\frac{d \ln \tilde{m} (\gamma)}{d \ln \gamma}
\right|_{\gamma = 1}\,.\label{sensitiv_lv}
\end{equation}
For weakly gravitating objects, $\sigma \approx 0$. More in general,
Refs.~\cite{Yagi:2013qpa,Yagi:2013ava} showed that the sensitivity of
a star, whatever its compactness, can be extracted from the asymptotic
behavior of the metric describing the star to \textit{first} order in
a perturbative expansion in the velocity relative to the \AE ther or
khronon. These \textit{slowly moving} stellar solutions have been
discussed in Section~\ref{sec:NS_LV}, and it is possible to show that
the sensitivities can be mapped to the parameter $A$ appearing in
those solutions [cf.~Eq.~\eqref{eq:k1-asympt}]. Indeed, one can show
that the sensitivity in Einstein-\AE ther theory is
\begin{align}
\label{eq:sensitivity-AE}
\sigma_{\ifmmode\text{\AE}\else\AE\fi} &= \frac{2 c_{1} \left(2 A - 4 - \alpha_{1}^{\ifmmode\text{\AE}\else\AE\fi}\right)}{c_{-} \left(8 + \alpha_{1}^{\ifmmode\text{\AE}\else\AE\fi}\right)}\,,
\end{align}
while in khronometric theory one has
\begin{equation}
\sigma^{\mathrm{kh}} = \frac{2 A - 4 - \alpha_{1}^{\mathrm{kh}}}{8 + \alpha_{1}^{\mathrm{kh}}}\,.
\end{equation}
Here $\alpha^{\ifmmode\text{\AE}\else\AE\fi}_{1}$ and $\alpha^{\mathrm{kh}}_{1}$ are the
weak-field PPN parameters given in terms of the theory's coupling
constants (cf.~Section~\ref{lv-theories}). In the weak-field limit,
one can show that the sensitivity scales as~\cite{Foster:2007gr}
\begin{equation}
\label{eq:weak-field-AE-s}
s_{\ifmmode\text{\AE}\else\AE\fi}^{\mathrm{wf}} = \left( \alpha_1^{\ifmmode\text{\AE}\else\AE\fi} - \frac{2}{3} \alpha_2^{\ifmmode\text{\AE}\else\AE\fi} \right)
\frac{\Omega}{M} + \mathcal{O} \left( \frac{\Omega^2}{M^2}
\right)
\end{equation}
in Einstein-\AE ther theory, where $\Omega$ denotes the gravitational
binding energy. One obtains a similar expression in khronometric
theory by replacing $\alpha_{1,2}^{\ifmmode\text{\AE}\else\AE\fi}$ with
$\alpha_{1,2}^{\mathrm{kh}}$.
The top-left panel of Figure~\ref{fig:sens-alpha1} presents the
sensitivity as a function of the NS compactness for various EOSs~\cite{Yagi:2013qpa,Yagi:2013ava} . We
also show the weak-field sensitivity for the APR EOS, as defined in
Eq.~\eqref{eq:weak-field-AE-s}. The PPN parameters
$\alpha_{1,2}^{\ifmmode\text{\AE}\else\AE\fi}$ are chosen to saturate the Solar System bounds,
and the plot assumes $c_+=c_-=10^{-4}$. The bottom-left panel shows
the fractional difference between the sensitivity defined in
Eq.~\eqref{eq:sensitivity-AE} and the weak-field sensitivity for each
EOS: the two tend to the same limit for small compactness, and their
fractional difference is of $\sim 20\%$ at most in the
large-compactness limit. Observe also that the relation between the NS
sensitivity and compactness is insensitive to the EOS.
Similarly, the right panels of
Figure~\ref{fig:sens-alpha1} present the NS sensitivity (and the
fractional difference from the weak-field limit), with the PPN parameters $\alpha_{1,2}^{\mathrm{kh}}$ again
chosen to saturate the Solar System constraints, and $\beta =
10^{-4}$. As in Einstein-\AE ther theory, the weak-field and
strong-field sensitivities have a common limit as the compactness
decreases, but in the large-compactness regime their fractional
difference can exceed 100\%. The relation between the NS sensitivity
and compactness in khronometric theory is also insensitive to the EOS.
\begin{figure*}[h]
\capstart
\begin{center}
\begin{tabular}{r l}
\includegraphics[width=0.45\textwidth,clip=true]{sens-vs-C-new}
\includegraphics[width=0.55\textwidth,clip=true]{sens-C-HL}
\end{tabular}
\caption{\label{fig:sens-alpha1} (Left) The top panel shows the
absolute magnitude of the NS sensitivity in Einstein-\AE ther~theory
against the NS compactness for various EOSs, together with the
weak-field expression in Eq.~\eqref{eq:weak-field-AE-s} with the APR
EOS. The bottom panel presents the fractional difference between the
sensitivity and the weak-field expression in
Eq.~\eqref{eq:weak-field-AE-s}. The PPN parameters $\alpha_1^{\ifmmode\text{\AE}\else\AE\fi}$
and $\alpha_2^{\ifmmode\text{\AE}\else\AE\fi}$ saturate the Solar System constraint, and
$c_+=c_-=10^{-4}$. Observe that the weak-field result becomes
inaccurate for realistic NS compactnesses, and that the relation is
EOS-insensitive. (Right) Same as the left panels, but for
khronometric theory with $\beta =
10^{-4}$. [From~\cite{Yagi:2013ava}.]}
\end{center}
\end{figure*}
\subsection{Open problems}\label{op:NSs}
For the reader's convenience, here we list some important open
problems in the context of strong-field tests of gravity using NSs:
\begin{itemize}
\item Except for a few special cases, the properties of NSs in the
most general scalar-tensor theory with second-order equations of
motion (Horndeski gravity) have not been explored, even in the
static case.
\item As discussed in Section~\ref{NS_fofR}, the very existence of
compact stars in $f(R)$ gravity is still a matter of debate. The
generic consensus seems to be that while it is hard to construct NS
equilibrium configurations in $f(R)$ gravity from a numerical point
of view, there is no fundamental obstacle to their existence (but
see~\cite{Ganguly:2013taa} for a different point of view). Either
way, NS configurations with realistic values of the physical
parameters have never been constructed in viable $f(R)$ models.
\item There are no calculations of fast-rotating NSs in dCS gravity.
\item NS sensitivities have been computed only in scalar-tensor
theories, quadratic gravity and Lorentz-violating theories, but not
in other theories (such as massive gravity theories).
\item Despite the vast literature on the Vainshtein effect, there is
essentially no phenomenological study of NSs in massive gravity and
Galileon theories, even for static models.
\item Gravitational collapse in scalar-tensor theories has been
considered under very idealized assumptions for the
microphysics. Relativistic gravitational collapse in quadratic
gravity has not been studied yet. The analysis of relativistic
collapse in EiBI gravity and Palatini $f({\cal R})$ gravity is
crucial to clarify the issue of BH formation and singularity
avoidance of these theories.
\item The appearance of curvature singularities near the surface of
macroscopical
objects~\cite{Barausse:2007pn,Barausse:2007ys,Barausse:2008nm,Pani:2012qd,Pani:2013qfa}
in theories with auxiliary fields might be avoided using the
arguments put forward in~\cite{Kim:2013nna}, but it is important to
understand whether such effects are generic, and to devise tests to
discriminate these theories from GR.
\item Studies of universal relations for NSs and QSs are not
complete, even within GR. For example, there are no studies for
differentially rotating stars. The experimental interest of
differential rotation is probably limited, but differential
rotation may provide a way to understand universality breaking,
because it is expected to produce variations in the eccentricity of
isodensity contours~\cite{Stein:2013ofa,Yagi:2014qua}. Another
topic that deserves further investigation are magnetic
fields. Ref.~\cite{2014MNRAS.438L..71H} found that the universality
is lost for NSs with large magnetic fields, but it would be
interesting to see if it can be somehow restored for fixed
\emph{dimensionless} magnetic field strength (e.g.~by normalizing
the magnetic field strength with the NS mass or radius, as is the
case for rapidly rotating
NSs~\cite{Pappas:2013naa,Chakrabarti:2013tca,Yagi:2014bxa}).
\item Tests of GR with universal relations also deserve more study.
Theories that have not been considered in the context of universal
relations include quadratic gravity, Einstein-\AE ther and
Ho\v{r}ava gravity. Static and slowly rotating NS solutions at first
order in the slow-rotation parameter~\cite{Pani:2011xm}, as well as
fast rotating solutions~\cite{Kleihaus:2014lba}, are known for EdGB
gravity, whereas the properties of spinning NSs in dCS theory are
known only up at second order in
rotation~\cite{Yagi:2013bca}. Static NS solutions have been
constructed for Einstein-\AE ther gravity
in~\cite{Eling:2006df,Eling:2007xh,Greenwald:2009kp}. Finally,
slowly moving NS solutions were constructed for Einstein-\AE ther
and Ho\v{r}ava gravity in~\cite{Yagi:2013qpa,Yagi:2013ava}. These
works should be extended to at least quadratic order in rotation,
because the quadrupole moment is a second-order quantity. It would
also be interesting to consider massive scalar-tensor theories and
$f(R)$ gravity.
\item Universal relations among multipole moments are useful to
measure the NS mass and radius with future X-ray
observations~\cite{Baubock:2013gna,Psaltis:2013fha}. Using slowly
rotating NS solutions and ray-tracing algorithms, it will be
possible to compute X-ray pulse profiles from a rotating hot spot on
the NS surface in modified theories of
gravity~\cite{Psaltis:2010ww,Baubock:2011ke}. The number of
parameters describing the profile can be reduced using the modified
universal relations; then one could fit for the model parameters
(such as the stellar mass and radius) together with the coupling
constants in the modified theories. This analysis would reveal
whether future X-ray observations with
NICER~\cite{2012SPIE.8443E..13G} and
LOFT~\cite{2012AAS...21924906R,Feroci:2012qh} could constrain
deviations from GR (cf.~\cite{Lo:2013ava}).
\end{itemize}
\subsection{Tests of gravity from radio pulsars}
\label{subsec:pulsars}
\input{sec-gr-tests/subsec-pulsars}
\subsection{Testing general relativity with cosmology}
\label{subsec:cosmology}
\input{sec-gr-tests/subsec-cosmology}
\subsubsection{Open problems}\label{op:pulsars}
In the near future, radio astronomy will benefit from the operation of
new radio telescopes with significantly larger collecting area. By the
end of 2016 the Five-hundred-meter Aperture Spherical radio Telescope
(FAST)~\cite{Nan:2011um} should see ``first light,'' and in the early
2020s the SKA should reach its design sensitivity, greatly enhancing
timing precision of known pulsars (up to a factor of 100 for many of
them) and increasing the number of known pulsars by an order of
magnitude~\cite{Lazio:2013mea}. In terms of gravity tests with
pulsars, this means a leap in the precision of current tests, various
qualitatively new tests with already known systems, and the discovery
of many new ``pulsar laboratories.''
Concerning the known systems, for the first time we will be able to
test relativistic effects in binary pulsars beyond the leading
order. For instance, in the double pulsar we will be able to test the
mass-octupole and current-quadrupole corrections to the quadrupole
formula~\cite{Blanchet:1989cu,Kramer:2009zza}. In the double pulsar,
we should also be able to extract the Lense-Thirring drag from the
total $\dot\omega$, and by this measure the moment-of-inertia of
pulsar A: see Eq.~(5.23) of \cite{Damour:1988mr}. Since the moment of
inertia of a NS depends on its compactness and therefore on the
equation of state of NS matter, this measurement will test competing
NS matter models, and give insight into the properties of matter at
very high densities
($\sim 10^{15} \, {\rm
g\,cm^{-3}}$)~\cite{Lattimer:2004nj,Kramer:2009zza}.
Qualitatively, new tests could come from the discovery of a
pulsar-BH system, either a pulsar in orbit with a stellar mass
BH or in orbit around the supermassive BH in the
center of our Galaxy. If intermediate-mass BHs do exist in
some of the dense cores of globular clusters, this might be a third
option to find such a test system. The ranging capability that comes
with the timing of a pulsar would provide a unique probe for the
BH spacetime, and allow for tests of the frame dragging and
the no-hair theorem~\cite{Wex:1998wt,Liu:2011ae,Wex:2012au,Liu:2014uka}. But even for
theories that predict the same BHs as GR, a pulsar-BH
system could be a unique ``laboratory'' (see Figure~\ref{fig:stg}).
A completely different type of test could come from pulsar timing arrays, which
are presently used in the effort to detect nano-Hz GWs~\cite{IPTA):2013lea}.
With the timing capabilities of the SKA, one can hope to probe the polarization
and propagation properties of these long-wavelength GWs~\cite{Lee:2013sxl}, or even study the evolution of a single supermassive
BH binary~\cite{Corbin:2010kt,Lee:2011et}.
\subsubsection{Theory: the linear regime.}
Faced with the proliferation of ever-more exotic gravity theories, the conventional process of testing models on an individual basis is not the optimal way of testing GR on cosmological scales. Not only is the theory population too large to tackle (and growing still), but progress is slowed by the increasing mathematical complexity of the most popular ideas. It was for these reasons that, several years ago, a number of groups turned to the strategy of devising model-independent tests of gravity. Effectively, one attempts to build a template for what viable beyond-GR theories can look like, in terms of a minimal set of unknown parameters and functions. One then builds up a ``translation dictionary,'' that is, a correspondence of how the parameters involved in the general formalism relate to the parameters of specific theories. With this dictionary in hand, testing the unified framework provides an efficient way to test many theories simultaneously. However, the framework is more general than this, because it also covers regions of parameter space for which no corresponding theory has been established yet. Constraints on this part of the parameter space can be used to guide the direction of \textit{future} theoretical work.
While the PPN formalism~\cite{Will:2011nz} is an example of one such model-independent framework, it is limited to small-scale tests of gravity, say, in the Solar System or using compact object binaries. There is a need for an analogous formalism that can be applied to cosmological observables. This is a formidable task, so most work to date has focused on formalisms that capture the regime of linear cosmological perturbation theory only.
Let us consider, for a moment, the properties we would like a cosmological equivalent to PPN to have:
\begin{enumerate}
\item It should encapsulate a large portion of the existing theory space.
\item Existing theories should map onto it \textit{exactly}, rather than only as an approximation.
\item The parameters of the formalism should have physical meaning, rather than simply denoting mathematical terms.
\item The parameters (or combinations thereof) should be constrainable by near-future data.
\item The formalism should achieve all of the above with the minimum number of new parameters and/or free functions.
\end{enumerate}
There is no obvious, unique way to meet all of the above criteria. Instead, three species of parametrizations have been put forward in response to the challenge. One can think of these three formulations as corresponding to three manifestations of a gravity theory: its fundamental action, the field equations derived from that action, and the combinations of those field equations which directly influence observable quantities. We will now describe each of these formalisms briefly in turn.
\paragraph{Action-based approaches.}
To derive linearized gravitational field equations, one needs an action that is quadratic in perturbations. One way to parametrize gravity is to construct the most general quadratic action with a given field content that is consistent with some desirable symmetries. Placing a restriction on the derivative order of the field equations terminates the potentially infinite series of terms that could be constructed. A coefficient of appropriate mass dimensions is assigned to each term; these represent the ``dials'' of the formalism that can be tuned to match a specific theory.
This concept can, to a certain extent, be thought of as an EFT for modified gravity (cf.~Section~\ref{sec:EFT}), although the analogy with particle physics EFTs should not be taken too far.
To date, nearly all work has focused on actions constructed from the metric and
a single scalar degree of freedom~\cite{Battye:2012eu,Battye:2013aaa}, though a
bi-scalar case recently appeared in~\cite{Gergely:2014rna}.
Such parametrized actions can quickly grow to contain large numbers of terms. However, the authors of~\cite{Gubitosi:2012hu,Bloomfield:2012ff,Gleyzes:2013ooa, Gleyzes:2014rba,Gao:2014soa} have made use of a clever device to simplify the procedure. They first consider the situation as viewed from the \textit{unitary gauge}, in which the time coordinate is chosen such as to eliminate perturbations of the scalar field. As a result of having used up one gauge freedom in this way, the metric is left with three spin-0 perturbations instead of the usual two (after gauge fixing); one might say that the metric has ``eaten'' the scalar field.
Given the preferred space-time slicing, it is then natural to reformulate metric perturbations in the ADM formalism. An example of a resulting action is~\cite{Gleyzes:2013ooa}:
\begin{align}
S_{\mathrm{EFT}}={}&\int d^4x\sqrt{-g}\,\Bigg[\frac{M_P^2}{2}f(\eta)R-\Lambda(\eta)-c(\eta)g^{00}+\frac{M_2^4}{2}\left(\delta g^{00}\right)^2 \nonumber \\
&{}-\frac{m_1^3}{2}\delta K\delta g^{00}
-\frac{\bar{M}_2^2}{2}\delta K^2-\frac{\bar{M}_3^2}{2}\delta K^\mu_\nu\delta K_\mu^\nu+\frac{\mu_1^2}{2} {^{(3)}R}\delta g^{00} \nonumber \\
&{}+\frac{\bar{m}_5}{2} {^{(3)}R}\,\delta K+\frac{\lambda_1}{2} {^{(3)}R}^2+\frac{\lambda_2}{2} {^{(3)}R^\mu_\nu}\, {^{(3)}R}_\mu^\nu\Bigg]\,,
\end{align}
where $K_{\mu\nu}$ is the extrinsic curvature of the spatial hypersurfaces defined by the foliation, and $^{(3)}R$ is their three-dimensional intrinsic curvature. The metric component $g^{00}$ is related to the usual ADM lapse function by $g^{00}=-1/N^2$. Performing a Stuckelberg transformation~\cite{Cheung:2007st} on the time coordinate will exit the unitary gauge and cause the scalar field to reappear in the action. The action can then be varied in the usual manner.
The advantage of this EFT-inspired approach is that it directly parametrizes the
action, and therefore, constraints on the EFT coefficients would have direct
implications for which kinds of theories are allowed by the data. The
disadvantages are that i) it is necessary to fix the field content allowed by
the parametrization, and ii) the combinations of EFT parameters that filter
through to observable quantities (say, the modified growth rate or weak lensing
kernel) are cumbersome combinations of the action-level parameters. With regards
to i), we emphasize that nearly all work to date has focused on a single scalar
field. Clearly this covers only part of the broad space of gravity theories.
\paragraph{Field equation approaches.}
An alternative method is to directly parametrize the linearized gravitational field equations, the dynamical tools of a theory. After some consideration, one realizes that there are only three kinds of new terms that can appear in the field equations: perturbations of the metric, perturbations of the matter stress-energy tensor, and perturbations of any new degrees of freedom that a theory might add to GR.
An example in this category is the Parametrized Post-Friedmann formalism (PPF)~\cite{Baker:2012zs}. For simplicity, we consider here the parametrization of the spin-0 field equations, restricting them to be of second order in time derivatives. We will not show the full framework, but as a representative example the extended 00--component of the field equations in the conformal Newtonian gauge is:
\begin{align}
-a^2\delta G_0^0={}&8\pi G_N a^2 \rho_M\delta_M+A_0(k,a)k^2\Phi+F_0(k,a)k^2\Gamma \nonumber\\
&{}+\alpha_0(k,a)k^2\chi+\alpha_1(k,a)k^2\dot\chi\,,
\end{align}
where $\Gamma=(\dot\Phi+H\Psi)/k$, $\delta G^\mu_\nu$ is the usual linearized Einstein tensor and $\chi$ is a template variable representing a new spin-0 degree of freedom; only one new degree of freedom is shown above, but in principle more could be added. The new degree of freedom does not have to be a simple scalar field: for example, when the parametrization is used to describe Einstein-\AE ther theory, $\chi$ represents the spatial spin-0 perturbation of a vector.
$A_0(k,a),\,F_0(k,a),\,\alpha_0(k,a)$ and $\alpha_1(k,a)$ are the free functions of time and scale that act as the ``dials'' of the parametrization in this case. In fact, the scale-dependence of $A_0\ldots\alpha_1$ is fixed by Lorentz symmetry and the derivative order of the parametrization -- they can only contain powers like $k^{2n}$.
The other components of the spin-0 field equations follow an analogous pattern, see~\cite{Baker:2012zs} for details. Naively, it appears that 22 free functions are needed to map out all possible extensions to the field equations. In reality, this is too much freedom -- not all of these 22 PPF coefficients are independent. By connecting the PPF approach to the action-based parametrizations described above, one determines that only $\sim 5-9$ independent functions are needed to describe a simple scalar field theory (the exact number depends on assumptions made about Lorentz symmetry and whether one fixes the background expansion rate or not). The advantage of field equation approaches such as PPF is that they can encapsulate numerous gravity theories without requiring any assumptions about field content. The disadvantage is that, at face value, the parametrization contains redundant freedoms that could cause problematic degeneracies for a constraint analysis.
\paragraph{Quasistatic approaches.}
The final approach to parametrizing deviations from GR is the simplest, and therefore the easiest to constrain, but also the least comprehensive from a formal standpoint. It makes use of the \textit{quasistatic approximation}. To implement this, we focus on a restricted range of distance scales that are considered to be significantly smaller than the cosmological horizon, but sufficiently large that~\cite{Silvestri:2013ne}:
\begin{enumerate}
\item[(1)] Perturbations are in the linear regime.
\item[(2)] $H/k\ll 1$, so any term in the field equations containing this
prefactor can be dropped.
\item[(3)] The time derivatives of perturbations are negligible compared to
their spatial derivatives on these scales.
\end{enumerate}
In GR, it can be shown that (3) follows as a consequence of (2). In modified
theories this is no longer necessarily the case; instead it must be
\textit{assumed} that (3) applies to any new fields introduced by a theory, as
well as to the metric itself. The majority of theory-specific simulations to
date support this set of
assumptions~\cite{Oyaizu:2008sr,Noller:2013wca,Brax:2013mua,Brax:2012nk}, but
see~\cite{Llinares:2013qbh, Comelli:2014fg} for counter-examples.
By making the above approximations (and taking appropriate combinations of the field equations), one finds that in the quasistatic regime many theories of gravity can be reduced to the simple form:
\begin{align}
\nabla^2\Psi&=4\pi G_N a^2 \mu(k,a) \rho_M\Delta_M\,,\\
\gamma(k,a)&=\frac{\Phi}{\Psi}\,,
\end{align}
where $\Delta_M$ is a gauge-invariant matter density perturbation and, as in the previous subsection, $\mu(k,a)$ and $\gamma(k,a)$ are functions of time with fixed scale dependence of the form $k^{2n}$ (in the majority of cases). In particular, one can show that in the case of Horndeski gravity, which is the most general theory of a single scalar field with second-order field equations, $\mu$ and $\gamma$ have the form~\cite{DeFelice:2011hq, Amendola:2012ky}:
\begin{align}
\mu(k,a)&=h_1(a)\left(\frac{1+h_5(a)k^2}{1+h_3(a) k^2}\right)\,,\label{HDmu}\\
\gamma(k,a)&=h_2(a)\left(\frac{1+h_4(a)k^2}{1+h_5(a)k^2}\label{HDgamma}\right)\,,
\end{align}
where $h_i(a)$ are pure functions of time.
This parametrized Poisson equation and ``slip relation'' (the ratio of $\Phi$ and $\Psi$) are all that is needed to start calculating observable quantities, such as galaxy weak lensing and the growth rate of large-scale structure~\cite{Baker:2014tc,Leonard:2015hha}. Example forecasts for constraints on $\mu$ and $\gamma$ with the Large Synoptic Survey Telescope~\cite{LSST} can be found in~\cite{Hojjati:2013xqa}. Recent measurements using the 6dF peculiar velocity survey did not find any evidence for scale dependence in the growth rate of large-scale structure~\cite{Johnson:2014kaa}, which places a lower limit on any new mass scale involved in theories described by Eqs.~(\ref{HDmu}) and (\ref{HDgamma})~\cite{Baker:2014tc}.
The advantage of the quasistatic parametrization is clearly its simplicity and direct connection to observations. Its chief disadvantage is that it is an approximation with a limited regime of applicability, and does not exactly match the form of the underlying space of gravity theories. This obscures attempts to work out what constraints on $\mu$ and $\gamma$ really mean for a particular theory of interest.\newline
\subsubsection{Theory: the nonlinear regime.}\label{subsec-cosm-nonlinear}
Calculating the effects of modified gravity becomes significantly harder as we move into the nonlinear regime. Local (laboratory and Solar System) experiments place strong constraints on any deviation from GR: the results of such experiments require any new gravitational fifth force to be either very weakly coupled to matter or very short-ranged in the environments where the experiments have been performed. To avoid these strong experimental constraints, and at the same time give rise to interesting observable cosmological signatures, a screening mechanism is required. By screening we mean a way of hiding the modifications of gravity in our local (high-density) environments where high-precision gravity experiments have been performed, while at the same time allowing for potentially large deviations in regions of spacetime where the average density is much lower. We will briefly explain how screening works and which theories have some form of screening.
Most of the known screening mechanisms for scalar-tensor theories are encompassed by the general (Einstein-frame) Lagrangian:
\begin{equation}
\mathcal{L} = \frac{R}{2}M_{\rm Pl}^2 + \mathcal{L}(\phi,\partial\phi,\partial\partial\phi) + \mathcal{L}_m(A^2(\phi)g_{\mu\nu},\psi_m)\,.
\end{equation}
The matter fields are coupled to a metric $\tilde{g}_{\mu\nu} = g_{\mu\nu}A^2(\phi)$ that is conformally related to the space-time metric $g_{\mu\nu}$. In the nonrelativistic limit, such a theory gives rise to a fifth force given by:
\begin{equation}
\vec{F}_\phi = \frac{\beta(\phi)}{M_{\rm Pl}}\vec{\nabla}\phi,~~~~\beta(\phi) \equiv \frac{d\log A}{d\phi}M_{\rm Pl}\,.
\end{equation}
If the field equation for $\phi$ is linear, then the superposition principle is in play and screening cannot be achieved, so a fundamental requirement for a screening mechanism to work is nonlinear field equations. There are three ways of achieving this: with a self-interacting potential, in the coupling to matter or in the kinetic terms. To see the different ways screening can emerge we expand the Lagrangian about a field value $\phi_0$:
\begin{equation}
\mathcal{L} \simeq \frac{R}{2}M_{\rm Pl}^2 + Z^{\mu\nu}(\phi_0)\delta\phi_{,\mu}\delta\phi_{,\nu} +m^2(\phi_0)\delta\phi + \frac{\beta(\phi_0)\rho_m}{M_{\rm Pl}} + ...
\end{equation}
In a low-density environment, the field sits at some value $\phi_0 = \phi_A$ and the scalar field produces a fifth force on a test mass with strength $\alpha \propto \beta^2(\phi_A)$ relative to the gravitational force. Consider now a high-density region of space where $\phi_0 = \phi_B \not= \phi_A$. One way to reduce the effect of the fifth force is by having the field $\phi$ acquire a large local mass $m(\phi_B) \gg m(\phi_A)$, which implies a very short interaction range -- this is the chameleon mechanism~\cite{Khoury:2003rn,Mota:2006fz}. If the matter coupling is small, $\beta(\phi_B) \ll \beta(\phi_A)$, the fifth force will be suppressed -- the symmetron mechanism~\cite{Hinterbichler:2010es}. Lastly, the condition $|Z^{\mu\nu}(\phi_B)| \gg |Z^{\mu\nu}(\phi_A)|$ leads, after canonical normalization, to a weakened matter source and therefore also a weakened fifth force -- the Vainshtein mechanism~\cite{Vainshtein:1972sx,Deffayet:2009wt}.
In the nonlinear regime of structure formation the screening effect will be in operation and must be taken into account to obtain reliable theory predictions. Linear perturbation theory is unable to account for screening, as it is a purely nonlinear effect, and to understand structure formation in screened theories one is therefore led to N-body simulations. In such simulations one solves for the full evolution of the scalar field in the simulation box just as one does for the metric potential, though one can often apply the quasi-static approximation (see previous section) to simplify the field equation. The N-body equations of motion are given by: i) the particle displacement equation:
\begin{equation}
\ddot{\bf x} + \left(2H+\frac{\beta(\phi)\dot{\phi}}{M_{\rm Pl}}\right)\dot{\bf x} = -\frac{1}{a^2}\vec{\nabla}\Phi - \frac{1}{a^2}\frac{\beta(\phi)}{M_{\rm Pl}}\vec{\nabla}\phi\,,
\end{equation}
where the last term represents the scalar fifth force, and ii) the Poisson equation for the metric potential:
\begin{equation}
\nabla^2\Phi = 4\pi G_N a^2\delta\rho_{\rm eff}\,,
\end{equation}
where $\delta\rho_{\rm eff}$ is the perturbed total effective energy density, which contains contributions from matter and modifications to the Einstein tensor due to modified gravity. Lastly we have the field equation for $\phi$, which is model-dependent. Due to the nonlinearities in this equation, the method of choice for solving it is (Newton-Gauss-Seidel) relaxation. Accurately solving the field equation for the scalar field is by far the most challenging and time-consuming part of such simulations. To date several different models with different kinds of screening mechanism have been simulated, including
\begin{itemize}
\item Chameleon screening: Chameleon models~\cite{Brax:2013mua} and $f(R)$ gravity~\cite{Oyaizu:2008sr,Llinares:2013jza,Li:2011vk,Puchwein:2013lza}.
\item Symmetron screening: Symmetron models~\cite{Davis:2011pj,Brax:2012nk}.
\item Vainshtein screening: Dvali-Gabadadze-Porrati (DGP)~\cite{Schmidt:2009sv,Schmidt:2009sg,Li:2013nua} and Galileon models~\cite{Barreira:2013eea}.
\end{itemize}
The results from such simulations have so far mostly been used to map out potential signatures rather than computing explicit constraints. An exception is Ref.~\cite{Schmidt:2009am}, which constrains $f(R)$ gravity using cluster abundances, however the constraints found are not yet compatible with those found from taking local experiments into account.
The first key observable where the effects of modified gravity are seen is the matter power spectrum. When measured relative to $\Lambda$CDM, one typically finds an enhancement: a bump around $k\sim 1~h/$Mpc, as seen in Figure~\ref{fig:fofrpower}. N-body simulations have shown the importance of taking screening into account when making predictions: linear theory (or simulations with a linearized field equation for the scalar field) produces way too much clustering, and the power spectrum on nonlinear scales can be off by several tens of percent from the true result, as seen in Figure~\ref{fig:fofrpower}. Different screening mechanisms, different models and also different model parameters can give rise to very different results, making the construction of a model-independent parametrization, as presented above for the linear regime, hard to achieve. However, a parametrization valid for a certain sub-class of models (of the chameleon type) has been proposed~\cite{Brax:2012gr}.
\begin{figure}
\capstart
\includegraphics[scale=0.5,bb=0 130 0 700]{fofr_power_hu.pdf}
\caption{Relative power spectrum enhancement over $\Lambda$CDM at $a = 1$ for
the full $f(R)$ simulation compared with the no-chameleon (simulations with a
linearized field equation) simulations and linear perturbation theory. At high
$k$, linear theory overestimates the relative enhancement. Without the chameleon
screening mechanism, power is sharply enhanced on scales smaller than the
Compton scale in the background. For $|f_{R0}| = 10^{-6}$ the screening is very
effective and the fifth force in the simulation is strongly suppressed. For
$|f_{R0}| = 10^{-4}$, this suppression is nearly absent except for a residual
effect from the chameleon at high redshift. [From~\cite{Schmidt:2008tn}.]}
\label{fig:fofrpower}
\end{figure}
Another key observable is the halo mass function, which is also enhanced relative to $\Lambda$CDM, see Figure~\ref{fig:dndmfofr}. When the screening mechanism is working effectively, the enhancement is typically found for low to mid-size halos ($M \sim 10^{12}-10^{14}~M_{\rm sun}/h$). A key property of all known screening mechanisms is that they are more effective for more massive halos. This implies that the mass function converges to that of the underlying cosmological model in the high-mass end ($M \gtrsim 10^{15}~M_{\rm sun}/h$). If the screening mechanism is not very effective then this does not have to happen, and the mass function is enhanced even for the largest halo masses.
\begin{figure}
\capstart
\includegraphics[scale=0.5,bb=0 130 0 720]{dndm-reldev.pdf}
\caption{Relative mass-function enhancement over $\Lambda$CDM at $a = 1$ for the
full $f(R)$ simulation compared with the no-chameleon (simulations with a
linearized field equation) simulations and the spherical collapse model. For
the lowest value ($|f_{R0}|=10^{-4}$) the screening mechanism is not much in
play, and the mass function is enhanced in the high-mass end. As we go to
smaller values of $|f_{R0}|$ the screening mechanism becomes more and more
effective, and the mass function is only enhanced for mid-sized halo masses. We
also see the importance of solving the full field equation: the no-chameleon
simulation significantly overestimates the mass function in the high-mass end
for $|f_{R0}|=10^{-6}$. [From~\cite{Schmidt:2008tn}.]}
\label{fig:dndmfofr}
\end{figure}
Results from N-body simulations have also revealed several smoking-gun signatures of modified gravity. One particularly interesting signature is the difference between dynamical and lensing masses~\cite{Schmidt:2010jr}. The mass contained within a dark-matter halo can be found by either gravitational lensing measurements or by some measurement that relies on the dynamics of test masses. Gravitational lensing is determined by the sum of the two metric potentials $\Phi+\Psi$, and the inferred lensing mass in many theories with screening gives the same result as in GR. Dynamical masses, on the other hand, are affected by the fifth force, and will consequently be different from the GR versions. Combining lensing and dynamical mass measurements can therefore probe modified gravity. Additionally we have the effect that the amount of screening will depend on the density of the environment a certain massive object lies in. This will give rise to an environment dependence on dynamical mass estimates, and serves as a smoking gun signal for theories with screening. In principle, this effect could also distinguish between different types of screening~\cite{Zhao:2011cu}.
Quite often, the strongest effects of modified gravity (measured relative to $\Lambda$CDM) are found in the velocity field. Interesting signatures that have been found here include the low-order moments of the pairwise velocity distribution~\cite{Hellwing:2014nma} and the full phase space around galaxy clusters~\cite{Lam:2012by}.
There are issues related to the nonlinear regime that need to be better understood before we can fully exploit its potential as a probe of modified gravity. For example, when it comes to the matter power spectrum, the modified gravity signal found in simulations is degenerate with several other effects, such as massive neutrinos~\cite{Shim:2014uta} and baryonic feedback processes, that are not fully understood at the moment. One way around this issue is to look for observables that are not significantly affected by baryonic physics, or combining observables that allow us to break the degeneracies.
\subsubsection{Observations, current and future.}
There has been remarkable progress in constraining cosmological perturbations, and it is useful to summarize the data sets either in hand or expected over the next decade. The observables of choice are:
\begin{itemize}
\item Anisotropies in the CMB; the main statistic is the angular power spectrum of fluctuations, $C_\ell$. The current and future experiments are: WMAP, PLANCK, ACT, SPT, ACTPol, SPTPol, Spider, Polarbear, BICEP2, Keck Array.
\item Surveys cataloguing the angular positions and redshifts of individual galaxies leading to the power spectrum of fluctuations, $P(k)$, or the two-point correlation function, $\xi(r)$. The current and future experiments are: BOSS, DES, Weave, HETDEX, eBOSS, MS-DESI, LSST, Euclid, SKA, Chime, Baobab, MEERKAT, ASKAP.
\item Weak lensing; images of distant galaxies will be distorted and correlated by intervening gravitational potential wells, leading to statistics such as the convergence power spectrum, $C^{\kappa}_\ell$. The current and future experiments are: DES, RCS, KIDS, HSC, LSST, Euclid, SKA.
\item Peculiar velocities; by measuring redshifts {\it and} radial distances of galaxies and clusters it is possible to reconstruct a radially projected map of large-scale motions. Progress in this field will primarily come for the latter method with Planck.
\end{itemize}
Cosmological data has been used to place constraints on standard scalar-tensor theories~\cite{Avilez:2013dxa} and related theories (such as $f(R)$~\cite{Lombriser:2010mp}, Galileons and more general Horndeski theories~\cite{Barreira:2012kk}), Einstein-Aether theories~\cite{Zuntz:2008zz,Koivisto:2008xf}, braneworld models (specifically DGP) and specific massive gravity models (see e.g.~\cite{Akrami:2013ffa,Konnig:2014xva} and references therein). There have been preliminary attempts at placing constraints on more generalized parametrizations. If one restricts oneself to the quasi-static functions, $\mu$ and $\gamma$, constraints are found to be very dependent on assumptions about time- and scale-dependence~\cite{Bean:2010zq,Hojjati:2013xqa}. So, for example, if these functions are assumed to be constant, constraints are found at the percent level, while freeing up the time evolution (but assuming scale independence) gives constraints of order $50-100\%$.
\subsubsection{Open problems}\label{op:cosm}
Cosmological observations will constrain GR on length scales which are fifteen orders of magnitude greater than current constraints. Current and future data will give us a unique opportunity to do so with remarkable precision. We currently have an excellent understanding of what happens in the linear regime and how it maps onto an incredibly broad family of models. A few self-consistent formulations of how to parametrize GR currently exist which mirror the PPN approach, the workhorse for testing gravity on Solar System scales. Observations will not be able to constrain all the parameters in these approaches, but it should be possible to constrain some of them reasonably well in the quasi-static regime.
There have been attempts at developing such a general framework on nonlinear scales. The inclusion of screening mechanisms has been an essential aspect of this approach. Unlike with the linear regime, a number of theoretical issues remain: how to efficiently and accurately model the nonlinear effects, how to incorporate the uncertainties that arise from baryonic physics -- most notably from supernovae and active galactic nuclei feedback -- and the more general problem of bias (or how galaxies trace the density field). These are hard problems that need to be understood and solved if we are to use the (abundant) information on smaller length scales.
There have been a number of suggestions for how to test gravity on galactic and cluster scales which may be promising. These involve the transverse Doppler gravitational redshifts in galaxy clusters~\cite{Zhao:2013nka}, the motion of BHs embedded in galaxies~\cite{Hui:2012jb}, constraints from distance indicators~\cite{Jain:2012tn}, galactic brightness~\cite{Davis:2011qf}, cluster abundances~\cite{Ferraro:2010gh} and cold tidal streams in galaxies~\cite{Penarrubia:2012ab}. These would add to, and complement, the constraints arising from large-scale structure.
In conclusion of this chapter we mention the results of the BICEP2 collaboration~\cite{Ade:2014xna}. Measurements of the CMB B-mode polarization at 150 GHz by the BICEP2 experiment were initially found consistent with a $\Lambda$CDM cosmological model plus a spectrum of tensor modes, described by a tensor-to-scalar ratio $r\approx0.2$. If confirmed, this result would have provided an independent confirmation of the existence of GWs, and placed restrictions on viable inflationary potentials.
However, the interpretation of the measurement in terms of primordial GWs was
quickly challenged by several authors (see e.g.~\cite{Flauger:2014qra}).
Subsequent polarization data from the ESA Planck satellite~\cite{Adam:2015wua}
(in seven frequency bands between 30 GHz and 353 GHz) revealed the signal
contribution from galactic dust in the BICEP2 fields to be much more
significant than accounted for by the foreground subtraction performed
in~\cite{Ade:2014xna}.
A joint analysis of data from BICEP2, the Keck Array and the ESA Planck satellite reduced the initial detection of $r$ to an upper limit of $r<0.12$ at $95\%$ confidence (using the standard primordial spectrum pivot scale of $0.05$~Mpc$^{-1}$) \cite{Ade:2015tva}. This is consistent with the upper limit obtained from CMB data alone, $r<0.11$ at 95$\%$ confidence (pivot scale $0.002$~Mpc$^{-1}$ \cite{Ade:2015lrj}; the constraint relaxes to $r<0.15$ if the tensor and scalar spectral indices are allowed to be scale-dependent).
As such, primordial GWs remain undetected. The BICEP2 and Keck Array continue to take data, now observing at both 150 GHz and 100 GHz. Upgrades of the Atacama Cosmology Telescope (``Advanced ACTPol'' \cite{Calabrese:2014gwa}) and the South Pole Telescope (``SPT-3G'' \cite{Benson:2014qhw}) have been proposed to improve the constraint on $r$.
|
\section{Introduction}
The EM algorithm going back to the seminal paper \cite{dempster} is a very general method for iterative computation of maximum likelihood estimates in the setting of incomplete data. The algorithm consists of an expectation step (E-step) followed by a maximization step (M-step) which led to the name EM algorithm.
Due to its general applicability and relative simplicity it has nowadays found its way into a great number of applications. These include maximum likelihood estimates of hidden Markov models in \cite{macdonald}, non-linear time series models in \cite{chan} and full information item factor models in \cite{meng} to give just a very limited selection.
Despite the simplicity of the basic idea of the algorithm its implementation in more complex models can be rather challenging. The global maximization of the likelihood in the M-step has recently been addressed successfully (see e.g. \cite{meng1993} and \cite{liu}). On the other hand, when the expectation of the complete likelihood is not known in closed form only partial solutions have been given yet. One approach developed in \cite{wei} uses Monte Carlo approximations of the unknown expectation and was therefore named Monte Carlo EM (MCEM) algorithm. As an alternative procedure the stochastic approximation EM algorithm was suggested in \cite{delyon}.
In this paper we take a completely different route by using a forward-reverse algorithm (cf. \cite{BS}) to approximate the conditional expectation of the complete data likelihood. In this respect we extend the idea from \cite{BS} to a Markov chain setting, which is considered an interesting contribution on its own. Indeed, Markov chains are more general in a sense, since any
diffusion monitored at discrete times yields canonically a Markov chain, but not every chain can be embedded (straightforwardly) into some continuous time diffusion the other way around.
The central issue is the identification of a parametric Markov
chain model $(X_{n},$ $n=0,1,\ldots)$ based on incomplete data, i.e. realizations of the
model, given on a typically course grid of time points, let us say
$n_{1},n_{2},\ldots n_{N}.$ Let us assume that the chain runs through
$\mathbb{R}^{d}$ and that the transition densities $p_{n,m}^{\theta}(x,y),$
$n\geq m,$ of the chain exist (with $p_{n,n}^{\theta}(x,y):=\delta_{x}(y)$),
where the unknown parameter $\theta$ has to be determined. The log-likelihood function based on the incomplete observations $(X_{n_1},\ldots, X_{n_N})$ is then given by
\begin{equation}\label{lh}
l(\theta,x_{n_1},\ldots,x_{n_N})=\sum_{i=0}^{N-1} \ln p_{n_{i},n_{i+1}}^{\theta
}(x_{n_{i}},x_{n_{i+1}}),
\end{equation}
with $X_{n_{0}}=x_{0}$ being the
initial state of the chain. Then the standard
method of maximum likelihood estimation would suggest to evaluate%
\begin{equation}
\underset{\theta}{\arg\max}\; l(\theta,X)=\underset{\theta}{\arg\max} \sum_{i=0}^{N-1}\ln p_{n_{i},n_{i+1}}^{\theta
}(X_{n_{i}},X_{n_{i+1}}).\label{am}%
\end{equation}
The problem in this approach lies in the fact that usually only the one-step transition
densities $p_{n,n+1}^{\theta} (x,y)$ are explicitly known, while any multi-step
density $p_{n,m}^{\theta}(x,y)$ for $m>n$ can be expressed as an $m-n-1$ fold
integral of one-step densities. In particular for larger $m-n,$ these multiple
integrals are numerically intractable however.
In the EM approach, we therefore consider the
alternative problem%
\begin{equation}
\underset{\theta}{\arg\max}\sum_{i=0}^{N-1}\sum_{j=n_{i}}^{n_{i+1}%
-1}\mathbb{E}\ln p_{j,j+1}^{\theta}(X_{j},X_{j+1}),\label{am1}%
\end{equation}
in terms of the \textquotedblleft missing data\textquotedblright\ $X_{n_{i}%
+1},...,X_{n_{i+1}-1},$ $i=0,...,N-1.$ As such, between two such consecutive
time points, $n_{i}$ and $n_{i+1}$ say, the chain may be considered as a
bridge process starting in realization $X_{n_{i}}$ and ending up in
realization $X_{n_{i+1}}$ (under the unknown parameter $\theta$ though), and
so each term in (\ref{am1}) may be considered as an expected functional of the
\textquotedblleft bridged\textquotedblright\ Markov chain starting at time
$n_{i}$ in (data point) $X_{n_{i}},$ conditional on reaching (data point)
$X_{n_{i+1}}$ at time $n_{i+1}.$
We will therefore develop firstly an
algorithm for estimating the terms in (\ref{am1}) for a given parameter
$\theta.$ This algorithm will be of forward-reverse type in the spirit of the
one in \cite{BS} developed for diffusion bridges. It should be noted here
that in the last years the problem of simulating diffusion bridges has
attracted much attention. Without pretending to be complete, see for example, \cite{BlS,DH,MT,S,SVW,SMZ}.
Having the forward-reverse algorithm at
hand, we may construct an approximate solution to (\ref{am1}) in a sequential
way by the EM algorithm: Once a generic approximation $\theta_m$ is constructed after $m$ steps, one estimates
\[
\theta_{m+1}:=\underset{\theta}{\arg\max}\sum_{i=0}^{N-1}%
\sum_{j=n_{i}}^{n_{i+1}-1}\widehat{\mathbb{E}}\ln p_{j,j+1}^{\theta}%
(X_{j}^{\theta_m},X_{j+1}^{\theta_m}),
\]
where $X^{\theta_m}$ denotes the Markov bridge process under the
transition law due to parameter $\theta_m$ and each term
\[
\widehat{\mathbb{E}}\ln p_{j,j+1}^{\theta}(X_{j}^{\theta_m%
},X_{j+1}^{\theta_m})
\]
represents a forward-reverse approximation of%
\[
\mathbb{E}\ln p_{j,j+1}^{\theta}(X_{j}^{\theta_m},X_{j+1}%
^{\theta_m})
\]
as a (known) function of $\theta.$
Convergence properties of approximate EM algorithms have drawn considerable
recent attention in the literature mainly driven by it's success in earlier
intractable estimation problems. An overview of existing convergence results
for the MCEM algorithm can be fund in \cite{neath}. Starting from a
convergence result for the forward-reverse representation for Markov chains we
prove almost sure convergence of the FREM sequence in the setting of curved
exponential families based on techniques developed in \cite{BS} and
\cite{fort}. Essentially the only ingredient from the Markov chain model for
this convergence to hold are exponential tails of the transition densities and
their derivatives that are straightforward to check in many examples.
Since computational complexity is always an issue in estimation techniques
that involve a simulation step, we also include a complexity analysis for the
forward-reverse algorithm. We show that the algorithm achieves an expected
cost of the order $O(N \log N )$ for sample size of $N$ forward-reverse
trajectories.
We mention a recent application paper by Bayer, Moraes, Tempone, and Vilanova
\cite{BMTV}, which focuses on practical and implementation issues of the
forward-reverse EM algorithm in the setting of Stochastic Reaction Networks
(SRNs), i.e., of continuous time Markov chains with discrete, but generally
infinite state space.
The structure of the paper is as follows. In Section \ref{sec2} we
recapitulate and adapt the concept of reversed Markov chains, initially
developed in \cite{MSS2} using the ideas in \cite{MSS1} on reversed
diffusions. A general stochastic representation --- involving standard
(unconditional) expectations only --- for expected functionals of
conditional Markov chains is constructed in Section \ref{sec3}. This
representation allows for a forward reverse EM algorithm that is introduced
and analyzed in Section \ref{sec4}. In Section \ref{sec:conv} we proof almost
sure convergence of the forward-reverse EM algorithm in the setting of curved
exponential families. Implementation and Complexity of the FREM algorithm are
addressed in Section \ref{sec:impl-forrev}. The paper is concluded with two
application examples in Section \ref{sec:appl} that demonstrate the wide scope
of our method.
\section{Recap of forward and reverse representations for Markov chains}\label{sec2}
Consider a discrete-time Markov process $(X_{n},\mathcal{F}_{n}%
),\ n=0,1,2,...,$ on a probability space $(\Omega,\mathcal{F},$ $\mathbb{P})$
with phase space $(S,\mathcal{S})$, henceforth called Markov chain. In general
we assume that $S$ is locally compact and that $\mathcal{S}$ is the Borel
$\sigma$-algebra on $S.$ For example, $S=\mathbb{R}^{d}$ or a proper subset
of $\mathbb{R}^{d}.$ Let $P_{n},\ n\geq0,$ denote the one-step transition
probabilities defined by
\begin{equation}
P_{n}(x,B):=\mathbb{P}(X_{n+1}\in B\ |\ X_{n}=x),\ n=0,1,2,...,\ \ x\in
S,\ B\in\mathcal{S}. \label{Ch0}%
\end{equation}
In the case of an autonomous Markov chain all the one-step transition
probabilities coincide and are equal to $P:=P_{0}=P_{1}=\cdot\cdot\cdot.$
Let $X_{m}^{n,x},$\ $m\geq n,$ be a trajectory of the Markov chain which is at
step $n$ in the point $x,$ i.e., $X_{n}^{n,x}=x.$ The multi-step transition
probabilities $P_{n,m}$ are then defined by%
\[
P_{n,m}(x,B):=\mathbb{P}(X_{m}^{n,x}\in B),\ \ x\in S,\ \ B\in\mathcal{S}%
,\ \ m\geq n.
\]
Due to these definitions, $P_{n,n}(x,B)=\delta_{x}(B)=1_{B}(x)$ (Dirac
measure)$,$ $P_{n}=P_{n,n+1},$ and the Chapman - Kolmogorov equation has the
following form:
\begin{equation}
P_{n,m}(x,B)=\int P_{n,k}(x,dy)P_{k,m}(y,B),\ \ \ \ x\in S,\ B\in
\mathcal{S},\ n\leq k\leq m. \label{Ch1}%
\end{equation}
Let us fix $N>0$ and consider for $0\leq n\leq N$ the function
\begin{equation}
u_{n}(x):=\int P_{n,N}(x,dy)f(y)=\mathbb{E}\ f(X_{N}^{n,x}), \label{Ch2}%
\end{equation}
where $f$ is $\mathcal{S}$-measurable and such that the mathematical
expectation in (\ref{Ch2}) exists; for example, $f$ is bounded. By the Markov
property we have for $0\leq n<N:$%
\begin{align*}
u_{n}(x) & =\mathbb{E}\ f(X_{N}^{n,x})=\mathbb{E}\ f(X_{N}^{n+1,X_{n+1}^{n,x}})\\
& =\mathbb{E}\ \mathbb{E}^{\mathcal{F}_{n+1}}f(X_{N}^{n+1,X_{n+1}^{n,x}})=\mathbb{E}\ \mathbb{E}^{X_{n+1}^{n,x}%
}f(X_{N}^{n+1,X_{n+1}^{n,x}})\\
& =\mathbb{E}\ u_{n+1}(X_{n+1}^{n,x})=\int u_{n+1}(y)P_{n}(x,dy).
\end{align*}
Thus, $u_{n}(x)$ satisfies the following discrete integral Cauchy problem
\begin{align}
u_{n}(x) & =\int u_{n+1}(y)P_{n}(x,dy),\ n<N,\label{Ch3}\\
u_{N}(x) & =f(x), \label{Ch4}%
\end{align}
and (\ref{Ch2}) is a forward probabilistic representation of its solution. In
fact, the probabilistic representation (\ref{Ch2}) can be used for simulating
the solution of (\ref{Ch3})-(\ref{Ch4}) by Monte Carlo. For our purpose,
reverse probabilistic representations we need a somewhat more general version
of the above result.
\begin{theorem}[cf.~\cite{MSS2}]
\label{Theorem1} Let $P_{n}$ be the one-step transition density of a Markov
chain $X$ as in (\ref{Ch0}) and let the function $f:$ $S\rightarrow\mathbb{R}$
be measurable and bounded. Let further $\varphi_{n}:$ $S\times S\rightarrow
\mathbb{R}$ be a measurable and bounded functions for $n=0,1,2,...$ Then, the
solution of the problem%
\begin{align}
w_{n}(x) & =\int w_{n+1}(z)\varphi_{n}(x,z)P_{n}%
(x,dz),\ \ \ n<N,\label{SysT1}\\
w_{N}(x) & =f(x) \label{SysT1a}%
\end{align}
\ has the following probabilistic representation:%
\begin{equation}
w_{n}(x)=\mathbb{E}\left[ f(X_{N}^{n,x})\mathcal{X}_{N}^{n,x,1}\right] , \label{PU}%
\end{equation}
where $(X,\mathcal{X})$ is an extended Markov chain in which $\mathcal{X}$ is
governed by the equations%
\begin{equation}
\mathcal{X}_{k+1}^{n,x,\gamma}=\mathcal{X}_{k}^{n,x,\gamma}\varphi_{k}%
(X_{k}^{n,x},X_{k+1}^{n,x}),\qquad\mathcal{X}_{n}^{n,x,\gamma}=\gamma
,\nonumber
\end{equation}
where $n\leq k<N.$
\end{theorem}
\begin{proof}
Note that $\mathcal{X}_{k}^{n,x,\gamma}=\gamma\mathcal{X}_{k}^{n,x,1}.$ Thus,
for $n<N,$ (\ref{PU}) may be written as%
\begin{align*}
w_{n}(x) &= \mathbb{E}\left[ f(X_{N}^{n+1,X_{n+1}^{n,x}})\mathcal{X}_{N}^{n+1,X_{n+1}%
^{n,x},\mathcal{X}_{n+1}^{n,x,1}}\right] \\
&= \mathbb{E}\ \mathcal{X}_{n+1}^{n,x,1}\mathbb{E}^{(X_{n+1}^{n,x},\mathcal{X}_{n+1}^{n,x,1}%
)}\left[ f(X_{N}^{n+1,X_{n+1}^{n,x}})\mathcal{X}_{N}^{n+1,X_{n+1}^{n,x}%
,1}\right] \\
&= \mathbb{E}\ \left[ \mathcal{X}_{n+1}^{n,x,1}w_{n+1}(X_{n+1}^{n,x})\right] \\
&=\mathbb{E}\ \left[ \varphi_{n}(x,X_{n+1}^{n,x})w_{n+1}(X_{n+1}^{n,x})\right] \\
&=\int w_{n+1}(z)\varphi_{n}(x,z)P_{n}(x,dz),
\end{align*}
and (\ref{SysT1a}) is trivially fulfilled for $n=N.$
\end{proof}
\subsection{Reverse probabilistic representations}
We henceforth take $\left( S,\mathcal{S}\right) =\left( \mathbb{R}%
^{d},\mathcal{B}(\mathbb{R}^{d})\right) $ and assume that the transition
probabilities $P_{n,m}(x,dy)$ have densities $p_{n,m}(x,y)$ with respect to
the Lebesgue measure on $(S,\mathcal{S}).$ We note however that without any
problem one may consider more general state spaces equipped with some
reference measure, and transition probabilities absolutely continuous to with
respect to it. The representation (\ref{Ch2}) can thus be written in the form
\begin{equation}
I(f):=\mathbb{E}\ f(X_{N}^{n,x})=\int p_{n,N}(x,y)f(y)dy,\text{ \ \ }0\leq n\leq N.
\label{R1}%
\end{equation}
Let the initial value $\xi$ of the chain $X$ at moment $n$ be random with
density $g(x).$ Consider the functional
\begin{equation}
I(g,f)=\int\int g(x)p_{n,N}(x,y)f(y)dxdy=\mathbb{E} f(X_{N}^{n,\xi}). \label{R2}%
\end{equation}
Formally, by taking for $g$ a $\delta$-function we obtain (\ref{R1}) again,
and by taking $f$ to be a $\delta$-function we obtain the integral%
\begin{equation}
J(g):=\int g(x)p_{n,N}(x,y)dx. \label{R3}%
\end{equation}
We now propose suitable (reverse) probabilistic representations for $J(g),$
where $g$ is an arbitrary test function (not necessarily a density). For this
we are going to construct a class of reverse Markov chains that allow for a
probabilistic representation for the solution of (\ref{R3}).
Let us fix a number $N\in\mathbb{N}$ and consider for $0\leq m<N,$ functions
$\psi_{m}:$ $S\times S\rightarrow\mathbb{R}_{+}$ such that for each $m$ and
$y$ the function
\begin{equation}
q_{m}(y,\cdot):=\frac{p_{N-m-1}(\cdot,y)}{\psi_{m}(y,\cdot)},\text{ \ \ }0\leq
m<N, \label{varfi}%
\end{equation}
is a density on $S.$ For example, one could take $\psi_{m}$ independent of the
second argument, and then obviously%
\begin{equation}
\psi_{m}(y)=\int p_{N-m-1}(z,y)dz. \label{exa}%
\end{equation}
We now introduce a \textquotedblleft reverse\textquotedblright\ processes
$(Y_{m}^{y},\mathcal{Y}_{m}^{y})_{0\leq m\leq N}$ by the system%
\begin{gather}
\mathbb{P}(\left. Y_{m+1}^{y}\in dz^{\prime}\right\vert \text{ }Y_{m}%
^{y}=z)=q_{m}(z,z^{\prime})dz^{\prime},\nonumber\\
\mathcal{Y}_{m+1}^{y}=\mathcal{Y}_{m}^{y}\psi_{m}(Y_{m}^{y},Y_{m+1}%
^{y}),\label{Yrev}\\
Y_{0}^{y}:=Y_{0}^{0,y}:=y,\text{ \ \ }\mathcal{Y}_{0}^{y}:=\mathcal{Y}%
_{0}^{0,y,1}:=1,\text{ \ \ }0\leq m<N,\nonumber
\end{gather}
hence $Y^{y}$ is governed by the one-step transition probabilities
$Q_{m}(z,dz^{\prime}):=$ $q_{m}(z,z^{\prime})dz^{\prime}$ (i.e. $Q_{m}$
instead of $P_{m}$).
\begin{theorem}
\label{rev_th} For any $n,$ $0\leq n\leq N,$ (\ref{R3}) has the following
probabilistic representation.%
\[
\int g(x)p_{n,N}(x,y)dx=\mathbb{E}\left[ g(Y_{N-n}^{y})\mathcal{Y}_{N-n}^{y}\right]
,
\]
where $g$ is an arbitrary test function (a \textquotedblright
density\textquotedblright\ $p_{m,m}$ has to be interpreted as a Dirac
distribution or $\delta$-function).
\end{theorem}
\begin{proof}
From the Chapman - Kolmogorov equation (\ref{Ch1}) we obtain straightforwardly
the Chapman-Kolmogorov equation for densities,
\begin{equation}
p_{n,m}(x,y)=\int p_{n,k}(x,z)p_{k,m}(z,y)dz,\ \ x,y\in S,\ n\leq k\leq m.
\label{R4}%
\end{equation}
Let us now fix $n,$ $n<N$ (for $n=N$ the statement is trivial) also, and
introduce the functions%
\begin{equation}
v_{k}(y):=\int g(x)p_{n,k}(x,y)dx,\ n\leq k\leq N. \label{R45}%
\end{equation}
From (\ref{R4}) we get
\begin{align}
v_{k}(y) & =\int v_{k-1}(z)p_{k-1}(z,y)dz,\ n<k\leq N,\label{R5}\\
v_{n}(y) & =g(y),\nonumber
\end{align}
where $p_{k-1}:=p_{k-1,k}$ denote the one-step densities. For $\ n<k\leq N$ we
now consider a \textquotedblleft reversed\textquotedblright\ time variable
$m=N+n-k$ and write with $\widetilde{v}_{m}(y):=v_{N+n-m}(y)$ and
(\ref{varfi}) system (\ref{R5}) in the form%
\begin{align}
\widetilde{v}_{m}(y) & =\int\widetilde{v}_{m+1}(z)\psi_{m-n}(y,z)q_{m-n}%
(y,z)dz,\ \ \ n\leq m<N,\label{R6_1}\\
\widetilde{v}_{N}(y) & =g(y).\nonumber
\end{align}
Let us write (\ref{R6_1}) in a slightly different form,%
\begin{align*}
\widetilde{v}_{m}(y) & =\int\widetilde{v}_{m+1}(z)\psi_{m}^{(n)}%
(y,z)q_{m}^{(n)}(y,z)dz,\ \ \ n\leq m<N,\\
\widetilde{v}_{N}(y) & =g(y)
\end{align*}
with $\psi_{m}^{(n)}:=\psi_{m-n}$ and $q_{m}^{(n)}:=q_{m-n}.$ Via
Theorem~\ref{Theorem1} we next obtain a probabilistic representation of the
form (\ref{PU}) for the solution of problem (\ref{R6_1}), hence (\ref{R3}) or
$J(g).$ Indeed, by taking in Theorem~\ref{Theorem1} instead of $X$ a Markov
chain $\left( Y_{m}^{(n),y}\right) _{n\leq m\leq N},$ where $Y^{(n),y}$ is
governed by the one-step transition probabilities $Q_{m}^{(n)}(z,dz^{\prime
}):=$ $q_{m}^{(n)}(z,z^{\prime})dz^{\prime},$ $n\leq m<N,$ with initial
condition $Y_{n}^{(n),y}=y,$ and constructing $\left( \mathcal{Y}_{m}%
^{(n),y}\right) _{n\leq m\leq N}$ according to%
\begin{equation}
\mathcal{Y}_{m+1}^{(n),y}=\mathcal{Y}_{m}^{(n),y}\psi_{m}^{(n)}(Y_{m}%
^{(n),y},Y_{m+1}^{(n),y}),\qquad\mathcal{Y}_{n}^{(n),y}=1,\qquad n\leq m<N,
\label{EE}%
\end{equation}
it follows by Theorem ~\ref{Theorem1} that%
\begin{equation}
J(g)=\widetilde{v}_{n}(y)=v_{N}(y)=\mathbb{E}\left[ g(Y_{N}^{(n),y})\mathcal{Y}%
_{N}^{(n),y}\right] . \label{RevProb}%
\end{equation}
It remains to see that
\[
\mathbb{E}\left[ g(Y_{N}^{(n),y})\mathcal{Y}_{N}^{(n),y}\right] =\mathbb{E}\left[
g(Y_{N-n}^{y})\mathcal{Y}_{N-n}^{y}\right]
\]
which follows from the fact that initial values and the one step transition
probabilities of the processes
\[
\left( Y_{n+i}^{(n),y},\mathcal{Y}_{n+i}^{(n),y}\right) _{i=0,...,N-n}\text{
\ \ and \ \ }\left( Y_{i}^{y},\mathcal{Y}_{i}^{y}\right) _{i=0,...,N-n}%
\]
coincide.
\end{proof}
It should be stressed that, in contrast to a corresponding theorem in
\cite{MSS2}, Theorem~\ref{rev_th} provides a family of probabilistic
representations indexed by $n=1,\ldots,N,$ that involves only one common
reverse process $Y^{y}.$ In Theorem~\ref{rev_th} $N$ was fixed but, when
different $N$ are in play, we will denote them by $Y^{y;N}.$ It turns out that
this extension of the related result in \cite{MSS2} is crucial for deriving
probabilistic representations for conditional Markov chains below (cf.
\cite{BS})
\section{Simulation of conditional expectations via forward-reverse
representations} \label{sec3}
In this section we describe for a Markov Chain (\ref{Ch0}) an efficient
procedure for estimating the final distributions of a chain $X=(X_{n}%
)_{n=0,...,N}$ conditioned, or pinned, on a terminal state $X_{N}.$ More
specifically, for some given (unconditional) process $X$ we aim at
simulation of the functional
\begin{equation}
\mathbb{E}\left[ \left. g(X_{m_{1}},\ldots,X_{m_{r}})\right\vert \text{
}X_{N}=y,\,X_{0}=x\right] , \label{cp}%
\end{equation}
where $0\leq m_{1}<m_{2}<\cdot\cdot\cdot<m_{r}<N$ (hence $r<N$), $g$ is an arbitrarily given
suitable test function, and $x, y\in\mathbb{R}^{d}$ are given states. The
procedure proposed below is in fact an extension of the method
developed in \cite{BS} to discrete time Markov chains. We note that similar
techniques as in \cite{BS} also allow us to treat the more general problem
\begin{equation*}
\mathbb{E}\left[ \left. g(X_{m_{1}},\ldots,X_{m_{r}})\right\vert \text{
}X_{N}\in A,\,X_{0}=x\right]
\end{equation*}
for suitable sets $A \subset \mathbb{R}^d$.
\subsection{Forward-reverse representations of conditional expectations}
Let us consider the problem (\ref{cp}) for fixed $x,y\in\mathbb{R}^{d}$
(i.e. $A=\{y\}$). We firstly state the following central theorem.
\begin{theorem}
\label{key} Given a grid $\mathcal{D}_{l}:=\{0\leq n^{\ast}<n_{1}<\cdot
\cdot\cdot<n_{l} =: N\},$ it holds that
\begin{multline*}
\mathbb{E}\left[ f(Y_{n_{l}-n_{0}}^{y;n_{l}},Y_{n_{l}-n_{1}}^{y;n_{l}}%
,\ldots,Y_{n_{l}-n_{l-1}}^{y;n_{l}})\mathcal{Y}_{n_{l}-n_{0}}^{y;n_{l}}\right]
\\
=\int_{\mathbb{R}^{d\times L}}f(y_{0},y_{1},\ldots,y_{l-1})\prod_{i=1}%
^{l}p_{n_{i-1},n_{i}}(y_{i-1},y_{i})dy_{i-1}%
\end{multline*}
with $y_{l}:=y$ and $n_{0}:=n^{\ast}.$
\end{theorem}
\begin{proof}
Without loss of generality, we assume in this proof that the grid satisfies
$n_{i} - n_{i-1} = 1$, $i=1, \ldots, l$. Indeed, extend $f:
\mathbb{R}^{d\times l} \to \mathbb{R}$ to a function $\tilde{f}:
\mathbb{R}^{d \times (N-n^\ast)} \to \mathbb{R}$ such that
\begin{equation*}
\tilde{f}\left( Y_{N-n^\ast}^{y;N}, Y_{N-n^\ast-1}^{y;N}, \ldots,
Y_{2}^{y;N}, Y_{1}^{y;N} \right) = f\left(
Y_{n_l-n_0}^{y;N}, Y_{n_l-n_1}^{y;N}, \ldots, Y_{n_l-n_{l-1}}^{y;N} \right).
\end{equation*}
Then, re-expressing the transition densities $p_{n_{i-1}, n_i}$ in terms of
the one-step transition densities $p_i$ using Chapman-Kolmogorov, we see
that the statement of the theorem is equivalent to
\begin{multline}
\label{eq:1}
\mathbb{E}\left[ \tilde{f}\left( Y_{N-n^\ast}^{y;N},
Y_{N-n^\ast-1}^{y;N}, \ldots, Y_{1}^{y;N} \right)
\mathcal{Y}^{y;N}_{N-n^\ast} \right] \\=
\int_{\mathbb{R}^{d\times(N-n^\ast)}} \tilde{f}\left( y_{n^\ast}, \ldots,
y_{N-1} \right) \prod_{i=n^\ast+1}^{N} p_{i-1}(y_{i-1}, y_i) dy_{i-1}
\end{multline}
with $y_{N} \equiv y$.
In fact, we shall prove that
\begin{equation}
\label{eq:2}
\mathbb{E}\left[ f_p\left( Y^{y;N}_{p}, \ldots, Y^{y;N}_{1} \right)
\mathcal{Y}^{y;N}_{p} \right] =
\int f_p(y_{N-p}, \ldots, y_{N-1}) \prod_{i=N-p+1}^N p_{i-1}(y_{i-1},
y_i) dy_{i-1}
\end{equation}
for any $1 \le p \le N-n^\ast$ for any (e.g., bounded measurable) function
$f_p : \mathbb{R}^{d\times p} \to \mathbb{R}$. \eqref{eq:2} gives the
formula from the statement of the theorem for $p=N-n^\ast$ with
$f_{N-n^\ast}$ being the function $\tilde{f}$ from above. We
prove~\eqref{eq:2} by induction on $p$. For $p=1$, this boils down to
Theorem~\ref{rev_th} with $n=N-1$.
For the step from $p-1$ to $p$, we note that by definition
\begin{equation*}
\mathcal{Y}_p^{y;N} = \mathcal{Y}_{p-1}^{y;N} \psi_{p-1}\left(
Y_{p-1}^{y;N}, Y_p^{y;N} \right),
\end{equation*}
with $\psi_{p-1}(y, \cdot) q_{p-1}(y, \cdot) = p_{N-(p-1)-1}(\cdot, y) =
p_{N-p}(\cdot, y)$ by~\eqref{varfi}. Hence, we have
\begin{align*}
\mathbb{E}\left[ f_p(Y_{p}^{y;N}, Y_{p-1}^{y;N} ,\ldots, Y_{1}^{y;N})
\mathcal{Y}_{p}^{y;N} \right] &= \mathbb{E}\left[ f_p(Y_{p}^{y;N},
Y_{p-1}^{y;N} ,\ldots, Y_{1}^{y;N}) \mathcal{Y}_{p-1}^{y;N}
\psi_{p-1}(Y_{p-1}^{y;N}, Y_{p}^{y;N}) \right] \\
&= \mathbb{E}\left[ g\left( Y_{p-1}^{y;N}, \ldots, Y_1^{y;N} \right)
\mathcal{Y}^{y;N}_{p-1} \right],
\end{align*}
with
\begin{align*}
g(z_{p-1}, \ldots, z_1) &\equiv \mathbb{E}\left[ \left. f_p\left( Y_{p}^{y;N},
Y_{p-1}^{y;N}, \ldots, Y_1^{y;N} \right) \psi_{p-1}(Y_{p-1}^{y;N},
Y_{p}^{y;N}) \right| Y_{p-1}^{y;N} = z_{p-1}, \ldots, Y^{y;N}_1 = z_1 \right]\\
&= \int f_p(z, z_{p-1}, \ldots, z_1) p_{N-p}(z, z_{p-1}) dz.
\end{align*}
Applying the induction hypothesis for $f_{p-1} = g$, we obtain
\begin{align*}
\mathbb{E}\left[ f_p(Y_{p}^{y;N}, Y_{p-1}^{y;N} ,\ldots, Y_{1}^{y;N})
\mathcal{Y}_{p}^{y;N} \right] &= \mathbb{E}\left[ g\left( Y_{p-1}^{y;N},
\ldots, Y_1^{y;N} \right) \mathcal{Y}^{y;N}_{p-1} \right]\\
&= \int g(y_{N-p+1}, \ldots, y_{N-1})
\prod_{i=N-p+2}^N p_{i-1}(y_{i-1}, y_{i}) dy_{i-1} \\
&= \int f_p(y_{N-p}, \ldots, y_{N-1}) \prod_{i=N-p+1}^N p_{i-1}(y_{i-1},
y_i) dy_{i-1}.\qedhere
\end{align*}
\end{proof}
We now consider an extended integer sequence%
\[
0<m_{1}<\cdot\cdot\cdot<m_{k}=n^{\ast}=n_{0}<n_{1}<\cdot\cdot\cdot<n_{l}=N,
\]
and a kernel $K_{\epsilon}$ of the form
\[
K_{\epsilon}(u):=\epsilon^{-d}K(u/\epsilon),\quad y\in\mathbb{R}^{d},
\]
with $K$ being integrable on $\mathbb{R}^{d}$ and $\int_{\mathbb{R}^{d}%
}K(u)du=1.$ Formally $K_{\epsilon}$ converges to the delta function
$\delta_{0}$ on $\mathbb{R}^{d}$ (in distribution sense) as $\epsilon
\downarrow0.$ We then have the following stochastic representation (involving
standard expectations only) for (\ref{cp}) \ with $n_{i}=m_{k+i},$
$i=0,...,l=r-k+1.$
\begin{theorem}
\label{FRR} Let the chain $\left( Y,\mathcal{Y}\right) :=\left(
Y^{y;N},\mathcal{Y}^{y;N}\right) $ be given by (\ref{Yrev}), and the modified
integer sequence $\left( \widehat{n}_{\cdot}\right) $ be defined by%
\begin{equation}
\widehat{n}_{i}:=n_{l}-n_{l-i},\quad i=1,\ldots,l.\label{eq:hat-grid}%
\end{equation}
It then holds%
\begin{gather}
\mathbb{E}\left[ \left. g(X_{m_{1}},\ldots,X_{m_{r}})\right\vert \ X_{m_{0}%
}=x,\ X_{N}=y\right] \nonumber\\
=\mathbb{E}\left[ \left. g(X_{m_{1}},\ldots,X_{m_{k-1}},X_{n^{\ast}}%
^{0,x},X_{n_{1}},\ldots,X_{n_{l-1}})\right\vert \ X_{m_{0}}=x,\,X_{N}%
=y\right] \nonumber\\
=\lim_{\epsilon\rightarrow0}\frac{\mathbb{E}\left[ g\left( X_{m_{1}}%
^{m_{0},x},\ldots,X_{m_{k-1}}^{m_{0},x},X_{n^{\ast}}^{m_{0},x},Y_{\widehat
{n}_{l-1}}^{y;N},\ldots,Y_{\widehat{n}_{1}}^{y;N}\right) K_{\epsilon}\left(
Y_{\widehat{n}_{l}}^{y;N}-X_{n^{\ast}}^{m_{0},x}\right) \mathcal{Y}%
_{\widehat{n}_{l}}^{y;N}\right] }{\mathbb{E}\left[ K_{\epsilon}\left(
Y_{\widehat{n}_{l}}^{y;N}-X_{n^{\ast}}^{m_{0},x}\right) \mathcal{Y}%
_{\widehat{n}_{l}}^{y;N}\right] }.\label{eq:for-rev-intro}%
\end{gather}
\end{theorem}
\begin{proof}
The proof is analogous to the corresponding one in \cite{BS}. As a
rough sketch, apply Theorem \ref{key} to%
\[
f(X_{m_{1}}^{0,x},\ldots,X_{n^{\ast}}^{0,x},y_{0},y_{1},\ldots,y_{l-1}%
):=g(X_{m_{1}}^{0,x},\ldots,X_{n^{\ast}}^{0,x},y_{1},\ldots,y_{l-1}%
)K_{\epsilon}(y_{0}-X_{n^{\ast}}^{0,x}),
\]
conditional on $X_{m_{1}}^{0,x},\ldots,X_{n^{\ast}}^{0,x},$ send
$\epsilon\rightarrow0,$ and divide the result by
\[
p_{0,N}(x,y)=\lim_{\epsilon\rightarrow0}\mathbb{E}\left[ K_{\epsilon}\left(
Y_{\widehat{n}_{l}}-X_{n^{\ast}}\right) \mathcal{Y}_{\widehat{n}_{l}}\right].
\qedhere
\]
\end{proof}
\subsection*{Forward-Reverse algorithm}\label{algo}
Given Theorem~\ref{FRR} the corresponding forward-reverse Monte Carlo
estimator for (\ref{eq:for-rev-intro}) suggests itself: Sample i.i.d. copies
$X^{0,x,(1)},...,X^{0,x,(M)}$ of the process $X^{0,x}$ and, independently,
i.i.d. copies $\left( Y^{y;N,(1)},\mathcal{Y}^{y;N,(\widetilde{M})}\right)
,...,\left( Y^{y;N,(1)},\mathcal{Y}^{y;N,(\widetilde{M})}\right) $ of the
process $\left( Y^{y;N},\mathcal{Y}^{y;N}\right) .$ Take for $K$ a second
order kernel, take for simplicity $M=\widetilde{M},$ and choose a bandwidth
$\epsilon_{M}\sim M^{-1/d}$ if $d\leq4,$ or $\epsilon_{M}\sim M^{-2/(4+d)}$ if
$d\geq4.$ By next replacing the expectations in the numerator and denominator
of (\ref{eq:for-rev-intro}) by their respective Monte Carlo estimates involving double sums, one
ends up with an estimator with Root-Mean-Square error $O(M^{-1/2})$ in the
case $d\leq4$ and $O(M^{-4/(4+d)})$ in the case $d>4$ (cf. \cite{BS} for
details).
\section{The forward-reverse EM algorithm}\label{sec4}
Let us now formulate the forward-reverse EM (FREM) algorithm in the setting of the missing data problem. Suppose that the parameter $\theta \in \Theta \subset \mathbb{R}^s$ and that the Markov chain $X=(X_n,n\in \mathbb{N})$ has state space $\mathbb{R}^d$. Assuming that the transition densities $p_{n,k}$ of $X$ exist for $n,k \in \mathbb{N}$ the full-data $\log$-likelihood then reads
\begin{equation}
l_c(\theta,x)=\sum_{i=0}^{n_N-1}\log p_{i,i+1}^{\theta
}(x_{i},x_{i+1}), \quad x \in \mathbb{R}^{n_N+1}.
\end{equation}
In the missing data problem only partial observations $X_{n_0},\ldots, X_{n_N}$ are available for $0=n_0 <n_1< \ldots < n_N$ with log-likelihood function $l$ given in \eqref{lh} that is intractable in most cases. Instead, the maximization of $l$ in $\theta$ has to be replaced by a two step iterative procedure, the EM algorithm.
\begin{description}
\item[E-step] In the $m$-th step evaluate the conditional expectation of the complete data $\log$-likelihood
\[
Q(\theta, \theta_m,x):= \mathbb{E}_{\theta_m} [l_c(\theta,X_0,\ldots,X_{n_N}) | X_{n_0}= x_{n_0},\dots,X_{n_N}=x_{n_N}], \quad x \in \mathbb{R}^{N+1}.
\]
\item[M-step] Update the parameter by
\[
\theta_{m+1} = \arg \max_\theta Q(\theta,\theta_m,X).
\]
\end{description}
Since in many Markov chain models the E-step is intractable in this form, we propose a forward-reverse approximation for the expectation of the transition densities evaluated at the observations.
\begin{description}
\item[FR E-step] Evaluate
\begin{equation}
Q_m(\theta, \theta_m,X):= \mathbb{E}_{\theta_m}^{FR} [l_c (\theta,X) | X_{n_i}, i = 0, \ldots,N],
\end{equation}
where $\mathbb{E}_{\theta_m}^{FR}$ denotes a forward-reverse approximation of the conditional expectation under the parameter $\theta_m$.
\end{description}
After this FR E-step is computed the M-step remains unchanged. This FREM algorithm gives a random sequence $(\theta_n)_{n \geq 0}$ that under certain conditions given in the next section converges to stationary points of the likelihood function. To assure a.s. boundedness of this sequence we apply a stabilization technique introduce in \cite{chen1988}.
\subsubsection*{The stable FREM algorithm}
Let $K_m \subset \Theta$ for $m \in \mathbb{N}$ be a sequence of compact sets such that
\begin{equation}\label{def:Kn}
K_m \subsetneq K_{m+1} \quad \text{and} \quad \Theta = \bigcup_{m \in \mathbb{N}} K_m
\end{equation}
for all $m \in \mathbb{N}$. We define the stable FREM algorithm by checking if $\theta_m$ after the $m$-th maximization step lies in $K_m$ and reseting the algorithm otherwise. Choose a starting value $\theta_0 \in K_0$ and let $p_n$ for $n \in \mathbb{N}$, $p_0 :=0$, count the number of resets.
\begin{description}
\item[stable M-step]
\begin{align}
\theta_{m+1} &= \arg \max_\theta Q_m (\theta,\theta_m,X) \text{ and } p_{n+1} = p_n,& \quad \text{ if } \arg \max_\theta Q_m (\theta,\theta_m,X) \in K_m,\\
\theta_{m+1} &= \theta_0 \text{ and } p_{n+1} = p_n+ 1,& \quad \text{ if } \arg \max_\theta Q_m (\theta,\theta_m,X) \notin K_m.
\end{align}
\end{description}
We will show in the next section that under weak assumption the number of resets $p_n$ stays a.s. finite. Our stable FREM algorithm consists now of iteratively repeating the FR E-step and the stable M-step.
\section{Almost sure convergence of the FREM algorithm} \label{sec:conv}
In this section we prove almost sure convergence of the stable FREM algorithm under the assumption that the complete data likelihood is from a curved exponential family. Our proof is mainly based on results from \cite{BS}, \cite{fort} and the classical framework for the EM algorithm introduced in \cite{dempster} and \cite{lange}.
\subsection{Model setting}
Suppose that $\phi:\Theta \to \mathbb{R}$, $\psi: \Theta \to \mathbb{R}^q$ and $S:\mathbb{R}^{(N+1)d} \to \mathbb{R}^q$ are continuous functions. We make the structural assumption that the full data log-likelihood is of the form
\begin{equation}
l(\theta,x_0, \ldots, x_N) = \phi(\theta)+ \langle S(x_0,\ldots,x_N),\psi(\theta)\rangle,
\end{equation}
i.e. $l$ is from a curved exponential family. In order to proof convergence we need the following properties to be fulfilled that naturally hold in many popular models. In Section \ref{sec:impl-forrev} we give several practical examples that fall into this setting.
\begin{assumption}\label{ass1}
\begin{enumerate}
\item There exists a continuous function $\bar \theta: \mathbb{R}^q \to \Theta$ such that $l(\bar \theta(s),s) = \sup_{\theta \in \Theta} l(\theta,s)$ for all $s \in \mathbb{R}^{(N+1)d}$.
\item The incomplete data likelihood $L$ is continuous in $\theta$, and the level sets $\{ \theta \in \Theta | L(
\theta,x) \geq C \}$ are compact for any $C > 0$ and all $x$.
\item The conditional expectation $\mathbb{E}_\theta [S(X_0, \ldots, X_{n_N})| X_{n_0}= x_{n_0},\dots,X_{n_N}=x_{n_N}]$ exists for all $(x_{n_0},\ldots, x_{n_N}) \in \mathbb{R}^{N+1}$ and $\theta \in \Theta$ and is continuous on $\Theta$.
\end{enumerate}
\end{assumption}
To simplify our notation we will neglect in the following the dependence of
$l$ on $s$. Under these assumption we can separate the E- and M-step. In order to do so we define
\[
g(\theta):= \mathbb{E}_\theta [S(X_0, \ldots, X_{n_N})| X_{n_0}= x_{n_0},\dots,X_{n_N}=x_{n_N}].
\]
An iteration of the EM algorithm can now be written as $\theta_{m+1} = \bar{\theta} \circ g(\theta_m) =:T(\theta_m)$. Let us denote by $\Gamma$ the set of stationary points of the EM algorithm, i.e.
\[
\Gamma = \{\theta \in \Theta | \bar \theta \circ g(\theta) = \theta \}.
\]
It was shown in Theorem 2 in \cite{wu1983} that if $\Theta$ is open, $\phi$ and $\psi$ are differentiable and Assumption \ref{ass1} holds, then
\[
\Gamma = \{\theta \in \Theta | \partial_\theta l(\theta) =0 \},
\]
such that the fixed points of the EM algorithm coincide with the stationary
points of $l$. In \cite{wu1983} it was proved that the set $\Gamma$ contains
all limit points of $(\theta_n)$ and that $(l(\theta_n))$ converges to
$l(\theta_0)$ for some $\theta_0 \in \Gamma$. In the following theorem we
extend these results to the FREM algorithm. Let $d(x,A)$ be the distance
between a point $x$ and a set $A$.
For the convergence of our forward-reverse based EM algorithm, we naturally
also need to guarantee convergence of the corresponding forward-reverse
estimators. This can be guaranteed by the following assumption (cf. also \cite[Section 4]{BS}).
\begin{assumption}
\label{ass2}
\begin{enumerate}
\item For any multi-indices $\alpha,\beta \in \mathbb{N}_0^d$ with $|\alpha|
+ |\beta| \le 2$ and any index $i$ there are constants $C_1 =
C_1(i,\alpha,\beta), C_2 = C_2(i, \alpha,\beta) > 0$ such
that
\begin{equation*}
| \partial_x^\alpha \partial_y^\beta p_{i}(x,y) | \le C_1 \exp\left( -
C_2 |x-y|^2 \right).
\end{equation*}
\item $S$ is twice differentiable in its arguments and both $S$ and its
first and second derivatives are polynomially bounded.
\end{enumerate}
\end{assumption}
Now we are ready to state as the main result of this section a general convergence theorem for the FREM algorithm.
\begin{theorem}\label{thm:FREMconv}
Let $(K_n)_{n \in \mathbb{N}}$ satisfy \eqref{def:Kn} and choose $\theta_0
\in K_0$. Suppose that Assumptions \ref{ass1} and \ref{ass2} hold and that
$l(\Gamma)$ is compact, then the stable FREM random sequence $(\theta_n)_{n
\geq 0}$ has the following properties:
\begin{enumerate}
\item $\lim_n p_n < \infty$ a.s. and $(\theta_n)$ is almost surely bounded.
\item $\lim_n d(l(\theta_n),l(\Gamma))= 0$ almost surely.
\item If also $l$ is $s$-times differentiable, then $\lim_n d(\theta_n,\Gamma)= 0$ almost surely.
\end{enumerate}
\end{theorem}
\begin{proof}
(1) Set
\[
g^{FR}(\theta) := \mathbb{E}_\theta^{FR} [S(X_0, \ldots, X_{n_N})| X_{n_0}= x_{n_0},\dots,X_{n_N}=x_{n_N}].
\]
With the above notation an iteration of the FREM algorithm can be written as
\[
\theta_{m+1} = \bar \theta (g^{FR} (\theta_m)).
\]
It was shown in Lemma 2 in \cite{delyon} that the incomplete data $\log$-likelihood $l$ is a natural Lyapunov function relative to $T$ and to the set of fixed points $\Gamma$. If for any $\epsilon >0$ and compact $K \subset \Theta$ we have
\begin{equation}\label{Tconv}
\sum_m \mathbf{1}_{|l\circ \bar \theta (g^{FR} (\theta_m))- l \circ T(\theta_m)|\mathbf{1}_{\theta_m \in K}\geq \epsilon} < \infty \quad \text{ a.s.},
\end{equation}
then Proposition 11 in \cite{fort} implies in our setting that $\lim \sup_n p_n < \infty$ almost surely and that $\theta_n$ is a compact sequence such that (1) follows. To obtain \eqref{Tconv} it is sufficient by Borel-Cantelli to prove that
\[
\sum_m P \left( |l\circ \bar \theta (g^{FR} (\theta_m))- l \circ T(\theta_m)|\mathbf{1}_{\theta_m \in K}\geq \epsilon\right) < \infty.
\]
Define for any $\delta >0$ an $\epsilon$-neighborhood of $K$ by
\[
K_\delta := \{ x \in \mathbb{R}^q |\inf_{z \in K} |z-x| \leq \delta \}.
\]
By assumption $l$ and $\bar \theta$ are continuous, such that for any $\delta >0$ there exists $\eta >0$ such that for any $x,y \in K_\delta$ we have $|l\circ \bar \theta (x) - l \circ \bar \theta (y)| \leq \epsilon$ whenever $|x-y| \leq \eta$.
Choosing now $\bar \epsilon = \delta \wedge \eta$ yields
\begin{align*}
P \left( |l\circ \bar \theta (g^{FR} (\theta_m))- l \circ T(\theta_m)|\mathbf{1}_{\theta_m \in K}\geq \epsilon\right) = P \left( |l\circ \bar \theta (g^{FR} (\theta_m))- l \circ \bar \theta (g(\theta_m))|\mathbf{1}_{\theta_m \in K}\geq \epsilon\right)\\
= P \left( |l\circ \bar \theta (g^{FR} (\theta_m))- l \circ \bar \theta (g(\theta_m))|\mathbf{1}_{\theta_m \in K}\geq \epsilon ,|g^{FR} (\theta_m) - g(\theta_m)|\mathbf{1}_{\theta_m \in K} \leq \delta \right)\\
+ P \left( |l\circ \bar \theta (g^{FR} (\theta_m))- l \circ \bar \theta (g(\theta_m))|\mathbf{1}_{\theta_m \in K}\geq \epsilon ,|g^{FR} (\theta_m) - g(\theta_m)|\mathbf{1}_{\theta_m \in K} \geq \delta \right)\\
\leq 2 P\left( |g^{FR} (\theta_m) - g(\theta_m)|\mathbf{1}_{\theta_m \in K} \geq \bar \epsilon \right)
\end{align*}
Markov's inequality gives then
\[
P \left( |l\circ \bar \theta (g^{FR} (\theta_m))- l \circ T(\theta_m)|\mathbf{1}_{\theta_m \in K}\geq \epsilon\right) \leq 2 \epsilon^{-k} \mathbb{E} \left[
\left|g^{FR} (\theta_m) - g(\theta_m) \right|^k \mathbf{1}_{\theta_m \in K} \right]
\]
for some $k >0$.
By \cite[Theorem 4.18]{BS} (see also Remark~\ref{rem:FREM-convergence} below),
we can always choose a number $N$ of samples for the forward-reverse algorithm
and a corresponding bandwidth $\epsilon = \epsilon_N = N^{-\alpha}$ such that
\begin{equation*}
\mathbb{E} \left[
\left|g^{FR} (\theta_m) - g(\theta_m) \right|^2 \mathbf{1}_{\theta_m \in K}
\right] \le \f{C}{N}
\end{equation*}
for some constant $C$.
We note that the choice of $\alpha$ depends on the
dimension $d$ as well as on the \emph{order} of the kernel. For instance, for
$d \le 4$ and a standard first order accurate kernel $K$, we can choose any
$1/4 \le \alpha \le 1/d$.
In any case, if we choose $N > m$ then
\begin{equation}\label{FRconv}
\sum_m P \left( |l\circ \bar \theta (g^{FR} (\theta_m))- l \circ T(\theta_m)|\mathbf{1}_{\theta_m \in K}\geq \epsilon\right)< \infty,
\end{equation}
which proves (1).
To prove (2) and (3) observe that for every $K \subset \Theta$ we have
\[
\lim_m | l(\theta_{m+1}) - l \circ T( \theta_m) | \mathbf{1}_{\theta_m \in K} =0 \quad a.s.
\]
By Borel-Cantelli it is sufficient to prove that
\[
\sum_m P (| l(\theta_{m+1}) - l \circ T( \theta_m) | \mathbf{1}_{\theta_m \in K} \geq \epsilon ) < \infty.
\]
But since we have shown in (1) that $p_n$ is finite a.s., we have in the above sum that $\theta_{m+1} = \bar \theta (g^{FR} (\theta_m))$ in almost all summands. Hence, it is sufficient to show that
\[
\sum_m P (| l(\bar \theta (g^{FR} (\theta_m))) - l \circ T( \theta_m) | \mathbf{1}_{\theta_m \in K} \geq \epsilon ) < \infty,
\]
which is nothing else than \eqref{FRconv}. The statement of (2) and (3) follows now from Sard's theorem (cf. \cite{broeckner}) and Proposition 9 in \cite{fort}.
\end{proof}
\begin{remark}
\label{rem:FREM-convergence}
In the above convergence proof we need to rely on the convergence proof of
the forward-reverse estimator when the bandwidth tends to zero and the
number of simulated Monte Carlo samples tends to infinity. Such a proof is
carried out for the diffusion case in \cite[Theorem 4.18]{BS}, where also
rates of convergence are given. We note that the proof only relies on the
transition densities of (a discrete skeleton of) the underlying diffusion
process. Hence, it immediately carries over to the present setting.
\end{remark}
Theorem \ref{thm:FREMconv} is a general convergence statement that links the limiting points of the FREM sequence to the set of stationary points of $l$. In many concrete models the set $\Gamma$ of stationary points consists of isolated points only such that an analysis of the Hessian of $l$ gives conditions for local maxima. A more detailed discussion in this direction can be found in \cite{delyon} for example.
\section{Implementation and complexity of the FREM algorithm}
\label{sec:impl-forrev}
Before presenting two concrete numerical examples, we will first discuss
general aspects of the implementation of the forward-reverse EM algorithm. For
this purpose, let us, for simplicity, assume that the Markov chains $X$ and
$(Y, \mathcal{Y})$ are time-homogeneous, i.e., that $p \equiv p_k$ and $q
\equiv q_k$ do not depend on time $k$. We assume that we observe the Markov
process $X$ at times $0 = i_0 < \cdots < i_r = N$, i.e., our data consist of
the values $X_{i_k} = x_{i_k}$, $k=0, \ldots, r$. For later use, we introduce
the shortcut-notation $\mathbf{x} := (x_{i_j})_{j=0}^r$.
The law of $X$ depends on an $s$-dimensional parameter $\theta \in \mathbb{R}^s$,
which we are trying to estimate, i.e., $p = p^\theta$. To this end, let
\begin{equation*}
\ell(\theta; x_0, \ldots, x_N) := \sum_{i=1}^N \log p^\theta(x_{i-1}, x_i)
\end{equation*}
denote the log-likelihood function for the estimation problem assuming full
observation. As before, we make the structural assumption that
\begin{equation}
\label{eq:structural}
\ell(\theta; x_0, \ldots, x_N) = \phi(\theta) + \sum_{i=1}^n S_i(x_0,
\ldots, x_N) \psi_i(\theta).
\end{equation}
For simplicity, we further assume that there are functions $S_i^j$ such that
\begin{equation*}
S_i(x_0, \ldots, x_N) = \sum_{j=1}^r S^j_i(x_{i_{j-1}}, \ldots, x_{i_j}).
\end{equation*}
The structural assumption~(\ref{eq:structural}) allows us to effectively
evaluate the conditional expectation of the log-likelihood $\ell_c$ for
different parameters $\theta$, without having to re-compute the conditional
expectations. More precisely, recall that for a given guess
$\widetilde{\theta}$ the E step of the EM algorithm consists in calculating
the function
\begin{equation}
\label{eq:Q-function-again}
\theta \mapsto Q(\theta; \widetilde{\theta}, \mathbf{x}) := \mathbb{E}_{\widetilde{\theta}}\left[
\left. \ell_c\left( \theta; X_0, \ldots, X_N) \right) \right| X_{i_j} =
x_{i_j}, \ j=0, \ldots, r \right],
\end{equation}
with $\mathbb{E}_{\widetilde{\theta}}$ denoting (conditional) expectation under the parameter
$\widetilde{\theta}$. Inserting the structural assumption~(\ref{eq:structural}), we
immediately obtain
\begin{equation*}
Q(\theta; \widetilde{\theta}, \mathbf{x}) = \phi(\theta) + \sum_{i=1}^m \psi_i(\theta)
\mathbb{E}_{\widetilde{\theta}}\left[
\left. S_i(X_0, \ldots, X_N) \right| X_{i_j} =
x_{i_j}, \ j=0, \ldots, r \right] = \phi(\theta) + \sum_{i=1}^m
S_i(\theta) z_i^{\widetilde{\theta}}
\end{equation*}
with $z_i^{\widetilde{\theta}} := \mathbb{E}_{\widetilde{\theta}}\left[ \left. S_i(X_0, \ldots, X_N) \right|
X_{i_j} = x_{i_j}, \ j=0, \ldots, r \right]$, $i=1, \ldots, m$. Note that
the definition of $z_i^{\widetilde{\theta}}$ does not depend on the free parameter
$\theta$. Thus, only one (expensive) round of calculations of conditional
expectations is needed for a given $\widetilde{\theta}$, producing a cheap-to-evaluate
function in $\theta$, which can then be fed into any maximization algorithm.
For any given $\widetilde{\theta}$, the calculation of the numbers $z_1^{\widetilde{\theta}},
\ldots, z_m^{\widetilde{\theta}}$ requires running the forward-reverse algorithm for
conditional expectations. More precisely, using the Markov property we decompose
\begin{multline*}
z_i^{\widetilde{\theta}} := \mathbb{E}_{\widetilde{\theta}}\left[ \left. S_i(X_0, \ldots, X_N) \right|
X_{i_j} = x_{i_j}, \ j=0, \ldots, r \right] \\
= \sum_{j=1}^r \mathbb{E}_{\widetilde{\theta}}\left[ \left. S_i^j(X_{i_{j-1}}, \ldots, X_{i_j})
\right| X_{i_{j-1}} = x_{i_{j-1}}, X_{i_j} = x_{i_j} \right].
\end{multline*}
All these conditional expectations are of the Markov-bridge type for which the
forward-reverse algorithm is designed. Hence, for each iteration of the EM
algorithm, we apply the forward-reverse algorithm $r$ times, one for the
time-intervals $i_{j-1}, \ldots, i_j$, $j=1, \ldots, r$, evaluating all the
functionals $h_1^j, \ldots, h_m^j$ at one go.
\subsection{Choosing the reverse process}
\label{sec:choos-reverse-proc}
Recall the defining equation for the one-step transition density $q$ of the
reverse process given in~\eqref{varfi}. For simplicity, we shall again assume
that the forward and the reverse processes are time-homogeneous, implying
that~\eqref{varfi} can be re-expressed as
\begin{equation*}
q(y,z) = \frac{p(z,y)}{\psi(y,z)}.
\end{equation*}
Notice that in this equation only $p$ is given a-priori, i.e., the user is
free to choose any re-normalization $\psi$ provided that for any $y \in \mathbb{R}^d$
the resulting function $z \mapsto q(y,z)$ is non-negative and integrates to
$1$. In particular, we can turn the equation around, choose \emph{any}
transition density $q$ and \emph{define}
\begin{equation*}
\psi(y,z) := \frac{p(z,y)}{q(y,z)}.
\end{equation*}
Note, however, that for the resulting forward-reverse process square
integrability of the process $\mathcal{Y}$ is desirable. More precisely, only
square integrability of the (numerator of the) complete estimator
corresponding to (\ref{eq:for-rev-intro}) is required, but it seems
far-fetched to hope for any cancellations giving square integrable estimators
when $\mathcal{Y}$ itself is not square integrable. From a practical point of
view, it therefore seems reasonable to aim for functions $\psi$ satisfying
\begin{equation*}
\psi \approx 1
\end{equation*}
in the sense that $\psi$ is bounded from above by a number
slightly smaller than $1$ and bounded from below by a number slightly smaller
than $1$. Indeed, note that $\mathcal{Y}$ is obtained by multiplying terms of
the form $\psi(Y_n, Y_{n+1})$ along the whole trajectory of the reverse
process $Y$. Hence, if $\psi$ is bounded by a large constant, $\mathcal{Y}$
could easily take extremely large values, to the extent that buffer-overflow
might occur in the numerical implementation -- think of multiplying $100$
numbers of order $100$. On the other hand, if $\psi$ is considerably smaller
than $1$, $\mathcal{Y}$ might take very small values, which can cause problems
in particular taking into account the division by the forward-reverse
estimator for the transition density in the denominator of the forward-reverse
estimator.
Heuristically, the following procedure seems promising.
\begin{itemize}
\item If $y \mapsto \int_{\mathbb{R}^d} p(z,y) dz$ can be computed in closed form (or
so fast that one can think of a closed formula), then choose
\begin{equation*}
\psi(y) := \psi(y,z) = \int_{\mathbb{R}^d} p(z,y) dz.
\end{equation*}
\item Otherwise, assume that we can find a non-negative (measurable) function
$\widetilde{p}(z,y)$ with closed form expression for $\int_{\mathbb{R}^d}
\widetilde{p}(z,y) dz$ such that $p(z,y) \approx \widetilde{p}(z,y)$. Then
define
\begin{equation*}
q(y,z) := \frac{\widetilde{p}(z,y)}{\int_{\mathbb{R}^d} \widetilde{p}(z,y) dz},
\end{equation*}
which is a density in $z$. By construction, we have
\begin{equation*}
\psi(y,z) = \frac{p(z,y)}{q(y,z)} = \int_{\mathbb{R}^d} \widetilde{p}(z,y) dz
\, \frac{p(z,y)}{\widetilde{p}(z,y)},
\end{equation*}
implying that we are (almost) back in the first situation.
\end{itemize}
\begin{remark}
Even if we can, indeed, explicitly compute $\psi(y,z) = \int_{\mathbb{R}^d} p(z,y)
dz$, there is generally no guarantee that $\mathcal{Y}$ has (non-exploding)
finite second moments. However, in practice, this case seems to be much
easier to control and analyze.
\end{remark}
\subsection{Complexity of the forward-reverse algorithm}
\label{sec:compl-forw-reverse}
We end this general discussion of the forward-reverse EM algorithm by a
refined analysis of the complexity of the forward-reverse algorithm for
conditional expectations as compared to \cite{BS}. We start with an auxiliary
lemma concerning the maximum product of numbers of two species of balls in
bins, which is an easy consequence of a result by Gonnet~\cite{gon81}, see
also~\cite[Section 8.4]{sed}.
\begin{lemma}
\label{lem:ball-bin}
Let $X$ be a random variable supported in a compact set $D \subset \mathbb{R}^d$
with a uniformly bounded density $p$. For any $K \in \mathbb{N}$ construct a
partition $B_1^K, \ldots, B_K^K$ of $D$ in measurable sets of equal Lebesgue
measure $\lambda(B^K_i) = \lambda(D)/K$, $i = 1, \ldots, K$. Finally, for
given $N \in \mathbb{N}$ let $X_1, \ldots, X_N$ be a sequence of independent copies
of $X$ and define
\begin{equation*}
N_k \coloneqq \# \left\{\left.i \in \set{1, \ldots, N} \,\right| \, X_i
\in B^K_k \right\}, \quad k = 1, \ldots, K.
\end{equation*}
For $N, K \to \infty$ such that $N = \mathcal{O}(K)$ we have the asymptotic
relation
\begin{equation*}
\mathbb{E} \left[ \max_{k=1, \ldots, K} N_k \right] = \mathcal{O}\left(
\f{\log N}{\log \log N} \right).
\end{equation*}
\end{lemma}
\begin{proof}
Let
\begin{equation*}
p_k \coloneqq P\left( X \in B^K_k \right) \le \norm{p}_{\infty}/K
\end{equation*}
and observe that the random vector $\left( N_1, \ldots, N_K \right)$
satisfies a multi-nomial distribution with parameters $K, N$ and $(p_1,
\ldots, p_K)$.
The proof for the statement in the special case of $p_1 = \ldots = p_K =
1/K$ is given in Gonnet~\cite{gon81}, so we only need to argue that the
relation extends to the non-uniform case. To this end, let $K' \coloneqq
\floor{K/\norm{p}_\infty}$ and let $(M_1, \ldots, M_{K'})$ denote a
multi-nomial random variable with parameters $N$, $K'$ and $(1/K', \ldots,
1/K')$. As $N = \mathcal{O}(K')$ we have by Gonnet's result that
\begin{equation*}
\mathbb{E}\left[ \max_{k=1, \ldots, K'} M_k \right] = \mathcal{O}\left(
\f{\log N}{\log\log N} \right).
\end{equation*}
Moreover, it is clear that $\mathbb{E}\left[ \max_{k=1, \ldots, K} N_k
\right] \le \mathbb{E}\left[ \max_{k=1, \ldots, K'} M_k \right]$ and we have
proved the assertion.
\end{proof}
\begin{theorem}
\label{thr:complexity-fr}
Assume that the transition densities $p$ and $q$ have compact support in
$\mathbb{R}^d$.\footnote{Obviously, this assumption can be weakened.} Moreover,
assume that the kernel $K$ is supported in a ball of radius $R > 0$. Then
the forward-reverse algorithm for $N$ forward and reverse trajectory based
on a bandwidth proportional to $N^{-1/d}$ can be implemented in such a way
that its expected cost is $\mathcal{O}(N \log N)$ as $N \to \infty$.
\end{theorem}
\begin{proof}
In order to increase the clarity of the argument, we re-write the double sum
in the forward-reverse algorithm to a simpler form, which highlights the
computational issues. Indeed, we are trying to compute a double sum of the
form
\begin{equation}
\label{eq:double-sum-simplified}
\sum_{i=1}^N \sum_{j=1}^N F_{i,j} K_\epsilon\left(X^i_{n^\ast} -
Y^j_{\hat{n}_l} \right),
\end{equation}
where $F_{i,j}$ obviously depends on the whole $i$th sample of the forward
process $X$ and on the whole $j$th sample of the reverse process $(Y,
\mathcal{Y})$.
We may assume that the end points $X^i_{n^\ast}$ and $Y^j_{\hat{n}_l}$ of
the $N$ samples of the forward and reverse trajectories are contained in a
compact set $[-L,L]^d$. (Indeed, the necessary re-scaling operation can
obviously be done with $\mathcal{O}(N)$ operations.) In fact, for ease of
notation we shall assume that the points are actually contained in
$[0,1]^d$. We sub-divide $[0,1]^d$ in boxes with side-length $S \epsilon$,
where $S > R$ is chosen such that $1/(S \epsilon) \in \mathbb{N}$. Note that
there are $K := (S \epsilon)^{-d}$ boxes which we order lexicographically
and associate with the numbers $1, \ldots, K$ accordingly.
In the next step, we shall order the points $X^i_{n^\ast}$ and
$Y^j_{\hat{n}_l}$ into these boxes. First, let us define a function $f_1:
[0,1]^d \to \{1, \ldots, 1/(S\epsilon)\}^d$ by setting
\begin{equation*}
f_1(x) := \left( \lceil x_1/(S\epsilon) \rceil, \ldots, \lceil x_d/(S
\epsilon) \rceil \right),
\end{equation*}
with $\lceil \cdot \rceil$ denoting the smallest integer larger or equal
than a number. Moreover, define $f_2: \{1, \ldots, 1/(S\epsilon)\}^d \to
\{1, \ldots, K\}$ by
\begin{equation*}
f_2(i_1, \ldots, i_d) := (i_1-1) (S \epsilon)^{-d+1} + (i_2-1) (S
\epsilon)^{-d+2} + \cdots + (i_d-1) + 1.
\end{equation*}
Obviously, a point $x \in [0,1]^d$ is contained in the box number $k$ if and
only if $f_2(f_1(x)) = k$.\footnote{To make this construction fully
rigorous, we would have to make the boxes half-open and exclude the
boundary of $[0,1]^d$.}
Now we apply a sorting algorithm like quick-sort to both sets of points
$\left(X^1_{n^\ast}, \ldots, X^N_{n^\ast} \right)$ and $\left(
Y^1_{\hat{n}_l}, \ldots, Y^N_{\hat{n}_l} \right)$ using the ordering
relation defined on $[0,1]^d \times [0,1]^d$ by
\begin{equation*}
x < y : \iff f_2(f_1(x)) < f_2(f_1(y)).
\end{equation*}
Sorting both sets incurs a computational cost of $\mathcal{O}(N \log N)$, so
that we can now assume that the vectors $X^i_{n^\ast}$ and $Y^i_{\hat{n}_l}$
are ordered.
Notice that $K_{\epsilon}(x-y) \neq 0$ if and only if $x$ and $y$ are
situated in neighboring boxes, i.e., if $\abs{f_1(x) - f_1(y)}_{\infty} \le
1$, where we define $\abs{\alpha}_\infty := \max_{i=1, \ldots, d}
\abs{\alpha_i}$ for multi-indices $\alpha$. Moreover, there are $3^d$ such
neighboring boxes, whose indices can be easily identified, in the sense that
there is a simple set-valued function $f_3$ which maps an index $k$ to the
set of all the indices $f_3(k)$ of the at most $3^d$ neighboring
boxes. Moreover, for any $k \in \{1, \ldots, K\}$ let $X^{i(k)}_{n^\ast}$ be
the first element of the ordered sequence of $X^i_{n^\ast}$ lying in the box
$k$. Likewise, let $Y^{j(k)}_{\hat{n}_l}$ be the first element in the
ordered sequence $Y^j_{\hat{n}_l}$ lying in the box with index $k$. Note
that identifying these $2K$ indices $i(1), \ldots, i(K)$ and $j(1), \ldots,
j(K)$ can be achieved at computational costs of order $\mathcal{O}(K\log N)
= \mathcal{O}(N \log N)$.
After all these preparations, we can finally express the double
sum~\eqref{eq:double-sum-simplified} as
\begin{equation}
\label{eq:double-sum-computational}
\sum_{i=1}^N \sum_{j=1}^N F_{i,j} K_\epsilon\left(X^i_{n^\ast} -
Y^j_{\hat{n}_l} \right) = \sum_{k =1}^K \sum_{r \in f_3(k)}
\sum_{i=i(k)}^{i(k+1)-1} \sum_{j=j(r)}^{j(r+1)-1} F_{i,j}
K_\epsilon\left(X^i_{n^\ast} - Y^j_{\hat{n}_l} \right).
\end{equation}
Regarding the computational complexity of the right hand side, note that we
have the deterministic bounds
\begin{gather*}
K = \mathcal{O}(N),\\
\abs{f_3(k)} \le 3^d.
\end{gather*}
Moreover, regarding the stochastic contributions, the expected maximum
number of samples $X^i_{n^\ast}$ ($Y^j_{\hat{n}_l}$, respectively) contained
in any of the boxes is bounded by $\mathcal{O}(\log N)$ by
Lemma~\ref{lem:ball-bin}, i.e.,
\begin{equation*}
\mathbb{E}\left[ \max_{r=1, \ldots, K} (j(r+1) - j(r)) \right]
\mathcal{O}(\log N),
\end{equation*}
which then needs to be multiplied by the total number $N$ of points
$X^i_{n^\ast}$ to get the complexity of the double summation step.
\end{proof}
\begin{remark}
As becomes apparent in the proof of Theorem~\ref{thr:complexity-fr}, the
constant in front of the asymptotic complexity bound does depend
exponentially on the dimension $d$.
\end{remark}
\begin{remark}
Notice that the box-ordering step can be omitted by maintaining a list of
all the indices of trajectories whose end-points lie in every single
box. The asymptotic rate of complexity of the total algorithm does not
change by omitting the ordering, though.
\end{remark}
\section{Applications of the FREM algorithm} \label{sec:appl}
The forward reverse EM algorithm is a versatile tool for parameter estimation in dynamic stochastic models. It can be applied in discrete time Markov models, but also in the the setting of discrete observations of time-continuous Markov processes such as diffusions for example.
In this section we give examples from both worlds: we start by a discretized Ornstein-Uhlenbeck process that serves as a benchmark model, since the likelihood function can be treated analytically. Then we give an example of a partially hidden Markov model that is motivated be applications in system biology in \cite{langrock}. Finally, we remark on limitations of the EM algorithm when the log-likelihood is not integrable and demonstrate these limitations in the context of a Cox-Ingersoll-Ross process. For a complex real data application of our method we refer the interested reader to the forthcoming companion paper \cite{BMTV}.
\subsection{Ornstein-Uhlenbeck dynamics}
\label{sec:ou}
In this section we apply the forward-reverse EM algorithm to simulated data from a discretized Ornstein-Uhlenbeck process. The corresponding Markov chain is thus given by
\begin{equation} \label{eq:ou}
X_{n+1} = X_n + \lambda X_n \Delta t + \Delta W_{n+1},\quad n\geq 0
\end{equation}
where $W_n$ are independent random variables distributed according to $N(0,\Delta t)$. The drift parameter $\lambda \in \mathbb{R}$ is unknown and we will employ the forward reverse EM algorithm to estimate it from simulated data. The Ornstein-Uhlenbeck model has the advantage that the likelihood estimator is available in closed form and we can thus compare it to the results of the EM algorithm.
In each simulation run we suppose that we have known observations \[X_{0}, X_{10\Delta t}, \ldots,X_{40 \Delta t}\]
for varying step size $\Delta t$ and use the EM methodology to approximate the likelihood function in between. We perform six iteration of the algorithm with increasing number of data points $N$.
In table \ref{tableOUsim1} we summarize the results of two runs for the discrete Ornstein-Uhlenbeck chain. The mean and standard deviation are estimated from 1000 Monte Carlo iterations. We find that already after three steps the mean is very close to the corresponding estimate of the true MLE. This indicates a surprisingly fast convergence for this example. Note also that the approximated value of the likelihood function stabilizes extremely fast at the maximum.
\begin{table}
\begin{tabular}{ c c c c c c c}
$\Delta t$ &N &bandwidth & mean $\hat \lambda$ & std dev $\hat \lambda$ & likel. & std dev likel. \\
\toprule
0.1 & 2000&0.0005 & 0.972 & 0.0135 & -3.402 & 0.00290\\
&8000&0.000125&1.098&0.00841&-3.383&0.00062\\
&32000& 3.125e-05& 1.132 &0.00476 &-3.381 &0.000123\\
&128000&7.812e-06 & 1.151&0.00236& -3.381& 2.783e-05\\
&512000&1.953e-06& 1.157& 0.00117& -3.381& 4.745e-06\\
&2048000&4.882e-07 & 1.159 &0.000581& -3.381& 1.005e-06\\
\midrule
0.05&2000 &0.0005 &1.160 &0.0141 &-3.107 &0.000854 \\
&8000 &0.000125 &1.247 &0.00872 &-3.103 &9.867e-05 \\
&32000 &3.125e-05 &1.253 &0.00468 &-3.103 &1.329e-05 \\
&128000 &7.812e-06 &1.265 &0.00225 &-3.103 &3.772e-06 \\
&512000 &1.953e-06 &1.265 &0.00111 &-3.103& 6.005e-07 \\
\end{tabular}
\caption{Behavior of the forward-reverse EM algorithm for a discretized
Ornstein-Uhlenbeck model for different step sizes $\Delta t$, initial guess
$\lambda = 0.5$ and true MLE $\hat \lambda_{\text{MLE}} = 1.161$ and
$1.266$, respectively.}
\label{tableOUsim1}
\end{table}
Table \ref{tableOUsim2} gives results for the same setup as in Table \ref{tableOUsim1} but with initial guess $\lambda = 2$ such that the forward-reverse EM algorithm converges from above to the true maximum of the likelihood function. We observe that the smaller step size $\Delta t = 0.05$ results in a more accurate approximation of the likelihood and also of the true MLE. It seems that the step size has crucial influence on the convergence rate of the algorithm, since for $\Delta t = 0.05$ the likelihood stabilizes already from the second iteration.
\begin{table}
\begin{tabular}{ c c c c c c c}
$\Delta t$ &N &bandwidth & mean $\hat \lambda$ & std dev $\hat \lambda$ & likel. & std dev likel. \\
\toprule
0.1 & 2000&0.0005 & 1.554 &0.0353 &-3.457 &0.0134\\
&8000&0.000125&1.312& 0.0127& -3.393& 0.00221\\
&32000& 3.125e-05& 1.217& 0.00544& -3.382& 0.000351\\
&128000&7.812e-06 & 1.185 &0.00245& -3.381& 5.817e-05\\
&512000&1.953e-06& 1.168& 0.00121& -3.381& 1.227e-05 \\
\midrule
0.05&2000 &0.0005 &1.390& 0.0248& -3.108& 0.00238 \\
&8000 &0.000125 & 1.289&0.00925& -3.103& 0.000130 \\
&32000 &3.125e-05 &1.261& 0.00471& -3.103& 1.451e-05 \\
&128000 &7.812e-06 &1.266& 0.00221& -3.103& 2.538e-06 \\
&512000 &1.953e-06 &1.266& 0.00113& -3.103& 5.855e-07 \\
\end{tabular}
\caption{Behavior of the forward-reverse EM algorithm for a discretized
Ornstein-Uhlenbeck model for different step sizes $\Delta t$, initial guess
$\lambda = 2$ and true MLE $\hat \lambda_{\text{MLE}} = 1.161$ and $1.266$,
respectively.}
\label{tableOUsim2}
\end{table}
In Figure \ref{OUhist1} the empirical distribution of 1000 estimates for $\lambda$ is plotted. The initial value was $0.5$ and the true maximum of the likelihood function is at $1.161$. The step size between observations was chosen to be $\Delta t = 0.1$. The histogram on the left shows the estimates after only one iteration and on the right the estimates were obtained from five iterations of the forward-reverse EM algorithm.
Figure \ref{OUhist2} depicts the distribution of 1000 Monte Carlo samples of the likelihood values that led to the estimates in Figure \ref{OUhist1}. It is interesting to see that after one iteration of the algorithm the likelihood values are approximately bell shaped (left histogram) whereas after five iterations the distributions becomes more and more one-sided as would be expected, since the EM algorithm only increase the likelihood from step to step towards the maximum.
\begin{figure}[htb]
\centering
\includegraphics[scale=0.4]{hist_samples_ou_13-12-17_1_0}
\includegraphics[scale=0.4]{hist_samples_ou_13-12-17_1_5}
\caption{Empirical distribution of 1000 estimates after one iteration (right) and after five iteration (left) of the forward-reverse EM algorithm.} \label{OUhist1}
\end{figure}
\begin{figure}[htb]
\centering
\includegraphics[scale=0.4]{like_samples_ou_13-12-17_1_0}
\includegraphics[scale=0.4]{like_samples_ou_13-12-17_1_5}
\caption{Empirical distribution of the likelihood values of 1000 Monte Carlo samples after one iteration (right) and after five iteration (left) of the forward-reverse EM algorithm.} \label{OUhist2}
\end{figure}
Figure \ref{OUbox} shows the convergence of the forward reverse EM algorithm when the number of iterations increases. We find that already after 4 iterations the estimate is very close to the true MLE for $\lambda$. After six iterations the algorithm has almost perfectly stabilized at the value of the true MLE $\lambda = 1.16$.
\begin{figure}[htb]
\centering
\includegraphics[scale=0.7]{boxplot_samples_ou_13-12-17_1}
\caption{Convergence of the forward-reverse EM algorithm from one to six iterations for each 1000 estimates of $\lambda$. The value of the true MLE is $\hat \lambda=1.161$.} \label{OUbox}
\end{figure}
\subsection{Hidden Markov models}
Our forward reverse EM algorithm can also be applied in the context of hidden Markov models (HMM). A typical HMM consists of observed state process and a hidden (i.e. unobserved) Markov chain that models the internal regime switching of the state process (see for example \cite{macdonald} for a general introduction). When for the hidden process only some values can be observed we speak of a partially HMM.
Recently, partially HMMs have been applied with some success in the modeling of ecological systems to understand the relation between animal populations and environmental parameters in a dynamical setting. The example considered here is a partially hidden Markov model and is inspired by applications in modeling mark-recapture-recovery data as discussed in \cite{langrock} for example. In mark-recapture-recovery studies every individual of a population is marked by some tag or ring that uniquely identifies the individual and some properties of interest are measured (e.g. weight, age etc). In subsequent surveys each individual is either recaptured such that the measurements could be taken again or if it can not be recaptured this will be recorded as a missing data point. Hence, we obtain a sequence of data with missing values that leads naturally to a partially HMM.
Let us consider a two-dimensional Markov chain $X_n=(X_n^1, X_n^2)$. For the first component $N$ observations $X_0^1, \ldots, X_N^1$ are given, whereas the second component $X^2$ is only observed partially after each $k \in \{1, \ldots, N\}$ time steps, i.e.
\[
X_0^2, X_k^2, X_{2k}^2, \ldots, X_{\lfloor \frac{N}{k}\rfloor k}^2
\]
are known. Suppose that the one-step transition probabilities are normal distributions:
\[
\mathcal{L} (X_{n+1} | X_n) = N( \mu^\theta (X_n), \Sigma),
\]
with mean given by $\mu^\theta (x) = \theta + x^2$ for an unknown parameter $\theta \in \mathbb{R}^2$. The covariance matrix $\Sigma \in \mathbb{R}^{2 \times 2}$ is assumed to be constant.
In this setup the $\log$-likelihood function based on full observations $X$ is given by
\[
l_c (\theta, X_0, \ldots, X_N) = 1- 2 \pi \det (\Sigma)^{1/2} - \frac{1}{2} \sum_{l = 1}^{N} Y_l^\theta.
\]
where
\begin{align*}
Y_l^\theta = \left( X_l - \mu^\theta (X_{l-1})\right)^\top \Omega \left( X_l - \mu^\theta (X_{l-1})\right)
\end{align*}
with precision matrix $\Omega = (\omega_{ij}) := \Sigma^{-1}$. In coordinates we thus have
\[
Y_l^\theta = \omega_{11} (U_{l,1})^2 + (2\omega_{12}) U_{l,1} U_{l,2} + \omega{22} U_{l,2}^2,
\]
where $ U_{l,i} := X_l^i - \theta_i - (X_{l-1}^i)^2$. Calculating the score function gives therefore
\[
\frac{\partial l_c}{\partial \theta_1} = - \frac{1}{2} \sum_{l = 1}^{N}\omega_{11} \left( 2(X_l^1)^2 + \theta^1 + 2 (X_{l-1}^1)^2 \right) + 2 \omega_{12} \left( \theta^2 - X_l^2 + (X_{l-1}^2)^2 \right)
\]
and
\[
\frac{\partial l_c}{\partial \theta_2} = - \frac{1}{2} \sum_{l = 1}^{N}\omega_{22} \left( 2(X_l^2)^2 + \theta^2 + 2 (X_{l-1}^1)^2 \right) + 2 \omega_{12} \left( \theta^1 - X_l^1 + (X_{l-1}^1)^2 \right).
\]
Since not all values of $X^2$ are observed, this $\log$-likelihood cannot be maximized directly. Instead, the forward reverse algorithm approximates the expected $\log$-likelihood given the partial observations. The E-step in this model reads as follows.
\subsection*{FR E-step}
Evaluate the forward-reverse approximation
\[
Q_m (\theta, \theta_m, X_1, \ldots, X_N) = \mathbb{E}_{\theta_m^{FR}} \left[ l_c (\theta, X_0, \ldots, X_n) | X_0^1, \ldots, X_N^1; X_{i k}^2, i =0,\ldots, \lfloor N/k\rfloor \right].
\]
This can be rewritten as
\[
Q_m (\theta, \theta_m, X_1, \ldots, X_N) = \operatorname{const} - \frac{1}{2} \mathbb{E}_{\theta_m}^{FR} \left[ \sum_{l =1}^N Y_l^\theta | X_0^1, \ldots, X_N^1; X_{i k}^2, i =0,\ldots, \lfloor N/k\rfloor \right].
\]
By approximating directly the score function this can be further simplified to
\begin{align*}
Q_m^{\partial \theta_1} &(\theta, \theta_m, X_1, \ldots, X_N) = - \frac{1}{2}(\omega_{11} \theta^1 +2\omega_{12} \theta^2) \\
\mathbb{E}_{\theta_m}^{FR} &\left[ \sum_{l =1}^N 2\omega_{11}((X_l^2)^2 + (X_{l-1}^1)^2) 2\omega_{12} ((X_{l-1}^2)^2- X_l^2) | X_0^1, \ldots, X_N^1; X_{i k}^2, i =0,\ldots, \lfloor N/k\rfloor \right].
\end{align*}
and
\begin{align*}
Q_m^{\partial \theta_2} &(\theta, \theta_m, X_1, \ldots, X_N) = - \frac{1}{2}(\omega_{22} \theta^2 +2\omega_{12} \theta^1) \\
\mathbb{E}_{\theta_m}^{FR} &\left[ \sum_{l =1}^N 2\omega_{22}((X_l^2)^2 + (X_{l-1}^2)^2) 2\omega_{12} ((X_{l-1}^1)^2- X_l^1) | X_0^1, \ldots, X_N^1; X_{i k}^2, i =0,\ldots, \lfloor N/k\rfloor \right].
\end{align*}
Due to the linearity of the score function in $\theta$ the M-step is now straightforward.
\subsection*{M-step}
Solve the linear system
\begin{align}
Q_m^{\partial \theta_1} &(\theta, \theta_m, X_1, \ldots, X_N)=0 \\
Q_m^{\partial \theta_2} &(\theta, \theta_m, X_1, \ldots, X_N)=0
\end{align}
for $\theta_1$ and $\theta_2$ to update $\theta_{m+1} = (\theta_1, \theta_2)$.
\begin{remark}
The example given here can also be extended to HMMs with more involved likelihood structure. In particular, the case of a non linear score function in $\theta$ can easily be treated with standard numerical methods such that also in these examples the M-step remains feasible (see also \cite{liu} and \cite{meng1993}).
\end{remark}
\subsection{A final note of warning by a discrete Cox-Ingersoll-Ross example}
\label{sec:cir}
Consider the Markov chain given by
\begin{equation}
X_{n+1}=X_{n}+\lambda\left( \theta-X_{n}\right) \Delta t+\sigma\left\vert
X_{n}\right\vert ^{\gamma}\Delta W_{n+1},\label{eq:cir}%
\end{equation}
where $\Delta t$ is fixed and $\Delta W_{n}$ are independent random variables
distributed according to $\mathcal{N}(0,\Delta t)$. Moreover, we assume that
$0\leq\gamma$ is fixed and known. The other parameters $\sigma$, $\lambda$ and
$\theta$ are unknown and need to be estimated. In the case
$\gamma=1/2$ it corresponds a of Euler discretization of the
Cox-Ingersoll-Ross model from finance.
Up to constant terms (in the un-known parameters $\sigma$, $\lambda$ and
$\theta$), the log-likelihood function of a sequence of observations
$\mathbf{x}=(x_{0},\ldots,x_{N})$ of the full path of the process $X$ is given
by
\begin{align*}
\ell_{c}\left( \sigma,\lambda,\theta;\mathbf{x}\right) & =\log\left(
\prod_{i=1}^{N}p(x_{i-1},x_{i})\right) \\
& =-N\log\sigma-\frac{1}{2\sigma^{2}\Delta t}\sum_{i=1}^{N}\frac{\left(
x_{i}-(1-\lambda\Delta t)x_{i-1}-\lambda\theta\Delta t\right) ^{2}%
}{\left\vert x_{i-1}\right\vert ^{2\gamma}}\\
& =-N\log\sigma-\frac{1}{2\sigma^{2}\Delta t}\sum_{i=1}^{N}\Biggl[\frac
{x_{i}^{2}}{\left\vert x_{i-1}\right\vert ^{2\gamma}}-2(1-\lambda\Delta
t)\frac{x_{i}x_{i-1}}{\left\vert x_{i-1}\right\vert ^{2\gamma}}\\
& \quad-2\lambda\theta\Delta t\frac{x_{i}}{\left\vert x_{i-1}\right\vert
^{2\gamma}}+(1-\lambda\Delta t)^{2}\frac{x_{i-1}^{2}}{\left\vert
x_{i-1}\right\vert ^{2\gamma}}\\
& \quad+2\lambda\theta\Delta t(1-\lambda\Delta t)\frac{x_{i-1}}{\left\vert
x_{i-1}\right\vert ^{2\gamma}}+\lambda^{2}\theta^{2}\Delta t^{2}\frac
{1}{\left\vert x_{i-1}\right\vert ^{2\gamma}}\biggr].
\end{align*}
Assume that we have given partial observations $X_{i_{0}},\ldots,X_{i_{r}}$ with
$i_{0}=0<\cdots<i_{r}=N$, while the remaining points $X_{j}$, $j\notin
\{i_{0},\ldots,i_{r}\}$, are assumed to be unobserved. Define random variables
$Z_{0}:=N$ and
\begin{align*}
& Z_{1}:=\sum_{i=1}^{N}\frac{X_{i}^{2}}{\left\vert X_{i-1}\right\vert
^{2\gamma}}, & Z_{2} & :=\sum_{i=1}^{N}\frac{X_{i}}{\left\vert X_{i-1}%
\right\vert ^{2\gamma}},\\
& Z_{3}:=\sum_{i=1}^{N}\frac{X_{i-1}X_{i}}{\left\vert X_{i-1}\right\vert
^{2\gamma}}, & Z_{4} & :=\sum_{i=1}^{N}\frac{1}{\left\vert X_{i-1}\right\vert
^{2\gamma}},\\
& Z_{5}:=\sum_{i=1}^{N}\frac{X_{i-1}}{\left\vert X_{i-1}\right\vert
^{2\gamma}}, & Z_{6} & :=\sum_{i=1}^{N}\frac{X_{i-1}^{2}}{\left\vert
X_{i-1}\right\vert ^{2\gamma}}.
\end{align*}
Hence, we have with $\mathbf{X}=(X_{0},\ldots,X_{N})$
\begin{multline*}
\ell_{c}\left( \sigma,\lambda,\theta;\mathbf{X}\right) =-Z_{0}\log
\sigma-\frac{1}{2\sigma^{2}\Delta t}\bigl[Z_{1}-2\lambda\theta\Delta
tZ_{2}-2(1-\lambda\Delta t)Z_{3}\\
+\lambda^{2}\theta^{2}\Delta t^{2}Z_{4}+2\lambda\theta\Delta t(1-\lambda\Delta
t)Z_{5}(1-\lambda\Delta t)^{2}Z_{6}\bigr].
\end{multline*}
Then we do the E-step. Given guesses $\sigma^{n},\lambda^{n},\theta^{n}$ for
the parameters, let
\begin{equation}
z_{i}:=\mathbb{E}_{\sigma^{n},\lambda^{n},\theta^{n}}\left[ \left.
Z_{i}\right\vert X_{i_{0}}=x_{i_{0}},\ldots,X_{i_{r}}=x_{i_{r}}\right] ,\quad
i=1,\ldots,6,\label{infE}%
\end{equation}
and observe that
\begin{align*}
Q(\sigma,\lambda,\theta;\sigma^{n},\lambda^{n},\theta^{n};x_{i_{0}}%
,\ldots,x_{i_{r}}) & :=\mathbb{E}_{\sigma^{n},\lambda^{n},\theta^{n}}\left[
\left. \ell_{c}\left( \sigma,\lambda,\theta;\mathbf{X}\right) \right\vert
X_{i_{0}}=x_{i_{0}},\ldots,X_{i_{r}}=x_{i_{r}}\right] \\
& =-z_{0}\log\sigma-\frac{1}{2\sigma^{2}\Delta t}\bigl[z_{1}-2\lambda
\theta\Delta tz_{2}-2(1-\lambda\Delta t)z_{3}\\
& \quad+\lambda^{2}\theta^{2}\Delta t^{2}z_{4}+2\lambda\theta\Delta
t(1-\lambda\Delta t)z_{5} + (1-\lambda\Delta t)^{2}z_{6}\bigr].
\end{align*}
Now the trouble is that for $\gamma\geq1/2$ some of the expectations in
(\ref{infE}), in particular $z_{4},$ may fail to exist. As such, this example
shows that in certain cases the expectation of the log-likelihood statistic in
the EM algorithm does not exist and that, as a consequence, the EM algorithm
can not be applied. We underline that existence of the log-likelihood
expectation is a premises for the EM algorithm in general and is not related
to the particular approach presented in this paper.
In the case $\gamma<1/2$ the expectations in (\ref{infE}) do exist and we may
proceed with first order conditions for finding the maximum of
\[
(\sigma,\lambda,\theta)\mapsto Q(\sigma,\lambda,\theta;\sigma^{n},\lambda
^{n},\theta^{n};x_{i_{0}},\ldots,x_{i_{r}}).
\]
We have that
\begin{align*}
\partial_{\sigma}Q & =-\frac{z_{0}}{\sigma}+\frac{1}{\sigma^{3}\Delta
t}\bigl[z_{1}-2\lambda\theta\Delta tz_{2}-2(1-\lambda\Delta t)z_{3}\\
& \quad+\lambda^{2}\theta^{2}\Delta t^{2}z_{4}+2\lambda\theta\Delta
t(1-\lambda\Delta t)z_{5}(1-\lambda\Delta t)^{2}z_{6}\bigr],\\
\partial_{\lambda}Q & =-\frac{1}{2\sigma^{2}\Delta t}\bigl[-2\theta\Delta
tz_{2}+2\Delta tz_{3}+2\lambda\theta^{2}\Delta t^{2}z_{4}\\
& \quad+2\theta\Delta t(1-2\lambda\Delta t)z_{5}-2\Delta t(1-\lambda\Delta
t)z_{6}\bigr],\\
\partial_{\theta}Q & =-\frac{\lambda}{2\sigma^{2}\Delta t}\bigl[-2\Delta
tz_{2}+2\lambda\theta\Delta t^{2}z_{4}+2\Delta t(1-\lambda\Delta
t)z_{5}\bigr].
\end{align*}
and we so obtain the maximizers given by
\begin{align*}
\sigma^{2} & =\frac{z_{3}^{2}z_{4}-2z_{2}z_{3}z_{5}+z_{1}z_{5}^{2}+z_{3}%
^{2}z_{6}-z_{1}z_{4}z_{6}}{\Delta tz_{0}\left( z_{5}^{2}-z_{4}z_{6}\right)
},\\
\lambda & =\frac{z_{3}z_{4}-z_{2}z_{5}+z_{5}^{2}-z_{4}z_{6}}{\Delta t\left(
z_{5}^{2}-z_{4}z_{6}\right) },\\
\theta & =\frac{z_{3}z_{5}-z_{2}z_{6}}{z_{3}z_{4}-z_{2}z_{5}+z_{5}^{2}%
-z_{4}z_{6}}.
\end{align*}
For the forward-reverse algorithm, we finally need to specify the reverse
chain. In this case, we propose to take the reverse chain
\begin{equation}
Y_{n+1}=Y_{n}-\lambda\left( \theta-Y_{n}\right) \Delta t+\sigma\left\vert
Y_{n}\right\vert ^{\gamma}\Delta\widetilde{W}_{n+1}.\label{eq:cir-reverse}%
\end{equation}
In order to get the dynamics of $\mathcal{Y}$, we need to derive the
normalization function $\psi$ between the one-step transition densities $p$ of
the forward and $q$ of the reverse processes. (We suppress the indices as we
are in a time-homogeneous situation.) For~(\ref{eq:cir}) together
with~(\ref{eq:cir-reverse}) the one-step transition densities are normal
densities in the forward variables,
\begin{gather*}
p(x,y)=\frac{1}{\sqrt{2\pi\Delta t}\sigma\left\vert x\right\vert ^{\gamma}%
}\exp\left( -\frac{\left( y-x-\lambda(\theta-x)\Delta t\right) ^{2}%
}{2\sigma^{2}\left\vert x\right\vert ^{2\gamma}\Delta t}\right) ,\\
q(y,z)=\frac{1}{\sqrt{2\pi\Delta t}\sigma\left\vert y\right\vert ^{\gamma}%
}\exp\left( -\frac{\left( z-y+\lambda(\theta-y)\Delta t\right) ^{2}%
}{2\sigma^{2}\left\vert y\right\vert ^{2\gamma}\Delta t}\right) .
\end{gather*}
Hence, we get
\begin{equation}
\psi(y,z)=\frac{p(z,y)}{q(y,z)}=\left\vert \frac{y}{z}\right\vert ^{\gamma
}\exp\left( -\frac{1}{2\sigma^{2}\Delta t}\left[ \frac{\left(
y-z-\lambda(\theta-z)\Delta t\right) ^{2}}{\left\vert z\right\vert ^{2\gamma
}}-\frac{\left( z-y+\lambda(\theta-y)\Delta t\right) ^{2}}{\left\vert
y\right\vert ^{2\gamma}}\right] \right) .\label{eq:psi-cir}%
\end{equation}
\bibliographystyle{plainnat}
|
\section{Introduction}
It was firstly proposed by S. Hod that the scalar field can have real
bound states in the near-extremal Kerr black hole \cite{hodprd2012,hodepjc2013}. Soon later,
it was reported in \cite{herdeiroprl} that massive scalar fields can form bound states around Kerr black holes by using the numerical method to solve the scalar field equation in the background.
This bound states are the stationary scalar configurations in the black hole backgrounds,
which are regular at the horizon and outside. They are named as
scalar clouds. More importantly, it was shown that the backreaction
of clouds can generate a new family of Kerr black holes with scalar hair \cite{herdeiroprl,herdeiroprd}.
It is suggested that whenever clouds of a given matter
field can be found around a black hole, in a linear analysis, there
exists a fully non-linear solution of new hairy black hole correspondingly.
However, it requires that the field originating clouds yields a time
independent energy momentum tensor. Generally, the field should be complex, and
have a factor $e^{-i\omega_c t}$, where $\omega_c$ is the superradiance critical
frequency. For instance, real scalar fields
can give rise to clouds but not hairy black holes \cite{1501}.
So, it seems that the studies of scalar clouds in the linear level
are very important for us to find the
hairy black holes in the non-linear level.
This subject has attracted a lot of attention recently
\cite{hodprd2014,RNclouds,benone,sampaio,grahamprd,Degolladoclouds,
Brihaye,HRRPLB,hodplb1,hodplb2,acoustic}.
Generally speaking, the existence of stationary bound states
of matter fields in the black hole backgrounds requires two
necessary conditions.
The first is that the matter fields should undergo the classical
superradiant phenomenon \cite{bardeen,misner} in the black hole background. This
condition can be satisfied by the bosonic fields in the rotating
black holes or the charged scalar fields in the charged
black holes \cite{bekenstein}. When the frequencies of these matter fields $\omega$
are smaller than the superradiant critical frequency $\omega_c$,
there are time growing quasi-bound states. When $\omega>\omega_c$,
the fields are time decaying. So, the scalar clouds exists at the
boundary between these two regimes, i.e. the frequencies of the fields
are taken as the superradiant critical frequency $\omega_c$.
For the rotating black holes,
the critical frequency $\omega_c$ is
$m\Omega_H$, where $m$ is the azimuthal index
and $\Omega_H$ is the horizon angular velocity.
While for the charged black holes, $\omega=q\Phi_H$,
where $q$ is the charge
of scalar field, and $\Phi_H$ is the horizon electrostatic potential.
The second one is there should be a potential well
outside the black hole horizon in which the bound states can be trapped.
This potential well may be
provided by the mass term of the field, i.e. $\omega<\mu$,
where $\mu$ is the mass of the scalar field. However, sometimes
the artificial boundary conditions can also play the same role.
In this paper, we will study the scalar clouds in a spherically symmetric
and charged background. Specifically, we will consider the charged scalar field
in the backgrounds of the charged stringy black holes. At first sight,
it seems that the massive scalar field can form the clouds in this background.
However, it is proved that the the massive charged scalar field
is stable in this background and there is no superradiant instability \cite{liprd}.
To generate the superradiant instability \cite{detweiler}, the mirror-like boundary condition
should be imposed according to the black hole bomb mechanism \cite{press,cardoso2004bomb}.
The analytical and the numerical studies on this subject can be found in
\cite{liepjc2014} and \cite{liplb2015}.
Correspondingly, the scalar clouds are only possible
with the mirror-like boundary condition. Using the numerical method,
we will study the dynamics of the massless charged scalar field satisfying
the frequency condition $\omega=q\Phi_H$ and the mirror-like boundary condition.
We will show that, for the specific set of
black hole and scalar field parameters, the clouds are only possible for
the specific mirror locations $r_m$. It will be shown that
the analytical results of mirror location $r_m$ for the clouds
are perfectly coincide with the numerical results. In addition, we will show that
the scalar clouds are also possible when the mirror locations are
close to the horizon. At last, we will provide an analytical calculation
of the specific mirror locations $r_m$ for the scalar clouds in the $qQ\gg 1$ regime.
This paper is organized as follows. In Sec.II,
we will present the background geometry of charged string black hole
and the dynamic equation of the scalar field. In particularly,
we will give the superradiant condition and the boundary
condition of this black hole-mirror system.
In Sec.III, we describe the numerical procedure to solve the
radial equation under the certain boundary condition.
In this section, the numerical results are also
illustrated. Some general discussion on the numerical
results are also followed. In Sec. IV, an analytical calculation
of the mirror radius $r_m$ for scalar clouds in $qQ\gg 1$ regime
is present. The conclusion is appeared in Sec. IV.
\section{Description of the system}
We shall consider a massless charged scalar field minimally coupled to
the charged stringy black hole with the mirror-like boundary condition.
The black hole is a static spherical symmetric charged black holes
in low energy effective theory
of heterotic string theory in four dimensions, which is firstly found by
Gibbons and Maeda in \cite{GM} and independently
found by Garfinkle, Horowitz, and Strominger in \cite{GHS} a few years later.
The metric is given by
\begin{eqnarray}
ds^2&=&-\left(1-\frac{2M}{r}\right)dt^2+\left(1-\frac{2M}{r}\right)^{-1}
dr^2\nonumber\\
&&+r\left(r-\frac{Q^2}{M}\right)(d\theta^2+\sin^2\theta d\phi^2)\;,
\end{eqnarray}
and the electric potential and the dilaton field
\begin{eqnarray}
&&A_t=-\frac{Q}{r}\;,\;\;\\
&&e^{2\Phi}=1-\frac{Q^2}{Mr}\;.
\end{eqnarray}
The parameters $M$ and $Q$ are the mass and the electric charge
of the charged stringy black hole, respectively.
The event horizon of black hole is located at $r=2M$.
The area of the sphere approaches to zero when $r=Q^2/M$. Therefore, the
sphere surface of the radius $r=Q^2/M$ is singular. When $Q^2\leq 2M^2$,
this singular surface is surrounded by the event horizon.
In this paper, we will always assume
the cosmic censorship hypothesis, i.e.
we will only consider the black hole with the parameters
satisfying the condition $Q^2\leq 2M^2$.
The dynamics of the charged scalar field
is then governed by the Klein-Gordon equation
\begin{eqnarray}
(\nabla_\nu-iqA_\nu)(\nabla^\nu-iqA^\nu)\Psi=0\;,
\end{eqnarray}
where $q$ denotes the charge of the scalar field.
By taking the ansatz of the scalar field
$\Psi=e^{-i\omega t}R(r)Y_{lm}(\theta,\phi)$,
where $\omega$ is the conserved energy of the mode,
$l$ is the spherical harmonic index, and $m$ is the
azimuthal harmonic index with $-l\leq m\leq l$,
one can deduce the radial wave equation in the form of
\begin{eqnarray}
\Delta\frac{d}{dr}\left(\Delta \frac{dR}{dr}\right)+UR=0\;,
\end{eqnarray}
where we have introduced a new function $\Delta=\left(r-r_+\right)\left(r-r_-\right)$
with $r_+=2M$ and $r_-=Q^2/M$,
and the potential function is given by
\begin{eqnarray}
U=\left(r-\frac{Q^2}{M}\right)^2(\omega r-qQ)^2-\Delta l(l+1)\;.
\end{eqnarray}
The superradiant condition of the charged scalar field is given by
\begin{eqnarray}
\omega<q\Phi_H\;,
\end{eqnarray}
where $\Phi_H=\frac{Q}{2M}$ is the electric potential at the
horizon\cite{dilatonsr,liprd}. It is proved in \cite{liprd} that the massive charged
scalar field is stable in this black hole background. To have superradiant
instability, we should impose the mirror-like boundary condition \cite{liepjc2014, liplb2015}.
In order to study the bound states, we shall focus on the critical case that
the scalar frequency equals to the superradiant critical frequency, i.e.
\begin{eqnarray}
\omega=q\Phi_H\;.
\end{eqnarray}
To solve the radial equation (5), we should impose the following boundary conditions,
which are given by
\begin{eqnarray}
R(r)=\left\{
\begin{array}{ll}
R_0\left(1+\sum_{k\geq 1}R_k(r-r_+)^k\right), & r\rightarrow r_+ \\
0, & r=r_m
\end{array}
\right.
\end{eqnarray}
The first line indicates that the scalar field is regular near the horizon and the second line
implies that the system is placed in a perfectly reflecting cavity.
\section{Numerical procedure and results}
The numerical methods employed in this problem are based on the shooting method,
which is also called the direct integration (DI) method \cite{Degolladoprd, Dolanprd2010,
Cardosoprd2014,Uchikata}. It is shown that the DI method is specially suited to find
stationary field configuration with the mirror-like boundary condition.
Firstly, near the event horizon $r=2M$, we require the radial function is regular
and expand the radial function $R(r)$ as
a generalized power series in terms of $(r-r_+)$
as have done in the first line of Eq.(9). Because the radial equation
is linear, we can take $R_0=1$ without loss of generality. Substituting expansion of the radial
wave function into the radial equation (5), we can solve the coefficient $R_k$ order by order
in terms of $(r-r_+)$. We have only considered six terms in the expansion. The $R_k$s
can be expressed in terms of the parameters $(M, Q, q, l)$, which are not exhibited here.
Then, we can integrate the radial equation (5) from $r=r_+(1+\epsilon)$
and stop the integration at the radius of the mirror. In this procedure,
we have taken the small $\epsilon$ as $10^{-6}$. The procedure can be repeated
by varying the input parameters $(M, Q, q, l)$ until the mirror-like boundary condition $R(r_m)=0$ is reached with the desired precision.
We can use a numerical root finder to search the
location of the mirror that support the stationary scalar configuration.
We have found that, for the given input parameters $(M, Q, q, l)$,
scalar clouds exist for a discrete set of $r_m$, which is labeled by the quantum number
$n$ of nodes of the radial function $R(r)$.
\begin{figure}
\subfigure{\includegraphics{anavsnumq1.eps}}
\subfigure{\includegraphics{anavsnumq2.eps}}
\caption{Mirror location $r_m$
plotted versus the black hole charge $Q$ for $M=1, l=1, n=0$ and for various scalar charge $q$.
For the first panel, $q=0.2$, while for the second panel, $q=0.8$.
The solid line and the dashed line represent the analytical and the numerical
results respectively.}
\end{figure}
Firstly, we make a comparison of the numerical and analytical
results. From the analytical result Eq.(35) in Ref.\cite{liepjc2014},
one can obtain the mirror radius that supports scalar cloud can
be approximately given by
\begin{eqnarray}
r_m=\frac{j_{l+1/2,n'}}{q\Phi_H}\;.
\end{eqnarray}
We have labeled the $n'$-th positive zero of the Bessel function $J_{l+1/2}$
as $j_{l+1/2, n'}$. The numerical results show that this "quantum number"
is closely connected with the nodes number $n$ of the radial function $R(r)$
towards the simple relation $n=n'-1$.
It should be noted that this analytical expression for the mirror radius
is only valid for the case of $qQ\ll 1$. With the condition $qQ\ll 1$,
the asymptotic expansion matched method can be employed to solve the radial equation
approximately \cite{liepjc2014}.
In Fig.(1), we have displayed the analytical results and the numerical results
of the mirror location $r_m$ in terms of the black hole charge $Q$.
Here, we do not consider the naked singularity spacetime, so that the
value range of black hole charge $Q$ is $(0, \sqrt{2}]$,
where we have fixed the black hole mass as $M=1$.
It is shown that the analytical results of mirror location $r_m$ for the clouds
are perfectly coincide with the numerical results, even in the region
where the analytical approximation is unapplicable.
When $q=0.2$, the analytical approximation is always precise
in all range of $Q$. When $q=0.8$, the analytical results have obvious
difference with the numerical results only for large $Q$.
\begin{figure}
\subfigure{\includegraphics{diffn.eps}}
\caption{Mirror location $r_m$
plotted versus the black hole charge $Q$ for $M=1, l=1, q=0.6$ and for various node number $n$.
The dotted, dashed, and solid lines represent $n=0, 1$, and $2$ respectively.}
\end{figure}
In Fig.(2), we have drawn the mirror location $r_m$ that support the scalar cloud
as a function of the black hole charge $Q$ for various values of node number $n$ of the radial
function. It is observed that, when the black hole charge $Q$ increases, we need to
place the reflecting mirror more closer to the horizon in order to
have a scalar cloud. When the node number $n$ of radial function increases,
the plotted lines become away from the axis. This observation is coincide with the
analytical result (10) in the regime of $qQ\ll 1$.
\begin{figure}
\subfigure{\includegraphics{diffL.eps}}
\caption{Mirror location $r_m$
plotted versus the black hole charge $Q$ for $M=1, n=0, q=0.6$ and for various $l$.
The dotted, dashed, and solid lines represent $l=1, 2$, and $3$ respectively.}
\end{figure}
In Fig.(3) and (4), we display the mirror location $r_m$ as a function of
the black hole charge $Q$ for various different $l$ and $q$.
We can observe that, the lines become far away from the axis when increasing $l$,
while the lines become more closer to the axis when increasing the scalar charge $q$.
This is also expected from the analytical result (10). In addition,
Fig.(3) and (4) together with Fig.(2) show that,
when $Q\rightarrow 0$, $r_m\rightarrow \infty$. This indicates that
there is no massless scalar cloud for Schwarzschild black hole
with the mirror-like boundary condition \cite{acoustic},
even thought it is possible for massive scalar fields in Schwarzschild
black hole to have arbitrarily long-lived quasi-bound states \cite{barranco}.
\begin{figure}
\subfigure{\includegraphics{diffq.eps}}
\caption{Mirror location $r_m$
plotted versus the black hole charge $Q$ for $M=1, l=1, n=0$ and for various scalar charge $q$. The dotted, dashed, and solid lines represent $q=0.4, 0.6$, and $0.8$ respectively.}
\end{figure}
\begin{figure}
\subfigure{\includegraphics{PlotRL1.eps}}
\subfigure{\includegraphics{PlotRL2.eps}}
\caption{Radial functions $R(r)$ of scalar clouds
for $M=1, q=0.6, r_m=40$ with different harmonic index $l$ and
node number $n$. The first and the second panels correspond
$l=1$ and $2$ respectively. The dotted, dashed, and solid
lines represent $n=1, 2$, and $3$ respectively.}
\end{figure}
\begin{figure}
\subfigure{\includegraphics{PlotRdiffq.eps}}
\caption{Radial functions $R(r)$ of scalar clouds
for $M=1, q=0.8, l=1, r_m=40$ with different
node number $n$. The dotted, dashed, and solid
lines represent $n=1, 2$, and $3$, respectively,
and the corresponding black hole charge $Q$ are
$0.219882, 0.583819$, and $0.956562$.}
\end{figure}
We also consider the radial dependence of the massless
scalar clouds. In Fig.(5) and (6),
we have fixed the mirror radius as $r_m=40$.
We can solve the radial equation numerically and
obtain a discrete set of black hole charge $Q$ which is labeled
by the node number $n$ of the radial wave equation.
Then we can integrate the radial equation for the fixed node numbers
and obtain the corresponding numerical solutions of the radial wave functions.
It is shown that the radial profile
have the typical forms of standing waves with the
fixed boundary conditions. We have also calculated the case that $l=3$.
The results is not present here. The general form of the radial wave
function is similar to the profiles in Fig.(5).
In Fig.(7), we consider the case that the mirror location
is very close to the horizon. We take the mirror radius as $r_m=3$.
From our previous analytical and numerical work
on the superradiant instability of scalar field in the background
of the charged stringy black hole plus mirror system,
we need a large scalar field charge $q$. Here, we set $q=20$.
We can see that, the scalar field can be bounded by the reflecting mirror
very near the horizon to form the clouds. The radial wave function
in this case have the similar profiles as Fig.(5) and (6).
\begin{figure}
\subfigure{\includegraphics{PlotRforsmallrm.eps}}
\caption{Radial functions $R(r)$ of scalar clouds
for the small mirror radius $r_m=3$. The parameters of
black hole and scalar field are taken as $M=1, q=20$, and $l=1$.
The dotted, dashed, and solid
lines represent $n=1, 2$, and $3$, respectively, and the
corresponding black hole charge $Q$ are $0.306384, 0.600699$,
$0.913741$.}
\end{figure}
\section{Scalar clouds in $qQ\gg 1$ regime}
In the above numerical calculations, we find the radial equation
becomes hard to integrate when the scalar charge $q$ is large.
So it is important to make an analytical study of the stationary
charged scalar clouds in the $qQ\gg 1$ regime. In this section,
we will give an analytical expression of special mirror radius $r_m$
in $qQ\gg 1$ limit, for which the charged scalar field can be confined
to form stationary cloud configuration.
Following \cite{hodhigh}, it is convenient to introduce new
dimensionless variables
\begin{eqnarray}
x=\frac{r-r_+}{r_+}\;,\;\;\tau=\frac{r_+-r_-}{r_+}\;,
\end{eqnarray}
in terms of which the radial equation (5) becomes
\begin{eqnarray}
x(x+\tau)\frac{d^2 R}{dx^2}
+(2x+\tau)\frac{dR}{dx}
\nonumber\\+\left[q^2 Q^2 x(x+\tau)-l(l+1))\right]R=0\;,
\end{eqnarray}
where we have submit the superradiance critical frequency $\omega=q\Phi_H$
in the above equation.
This equation can be solved by Bessel function in the double limit
\begin{eqnarray}
qQ\gg1\;,\;\;x\ll\tau\;.
\end{eqnarray}
In this asymptotic regime, the radial equation can be reduced to
\begin{eqnarray}
x \frac{d^2 R}{dx^2} +\frac{dR}{dx}+ q^2 Q^2 x R=0\;.
\end{eqnarray}
The solution is then given by the Bessel function of the first
kind
\begin{eqnarray}
R(x)=J_0(qQx)\;,
\end{eqnarray}
i.e., the stationary scalar field is then described by the above function.
By taking account to the mirror-like boundary condition $R(x_m)=0$, we can obtain
the special mirror radius $r_m$ as
\begin{eqnarray}
r_m=2M+\frac{j_{0,n}}{q\Phi_H}\;, \;\;\;n=1,2,3,\cdots
\end{eqnarray}
where $j_{0,n}$ is the $n$th positive zero of the Bessel function $J_0(x)$.
From this expression, we can see that, when $qQ\gg1$,
the reflecting mirror should be placed very near the horizon
to form the cloud configuration. This is consistent with the
near horizon condition $x\ll\tau$.
\section{Conclusion}
In summary, in this paper, we have studied
the massless scalar clouds in the charged stringy black holes
with the mirror-like boundary conditions.
The scalar clouds are stationary bound states satisfying
the superradiant critical frequency $\omega=q\Phi_H$.
The scalar clouds in rotating black holes \cite{herdeiroprl,benone}
can be heuristically interpreted
in terms of a mechanical equilibrium between the
Black hole-cloud gravitational attraction and angular momentum
driven repulsion. For the charged black hole cases,
the charged clouds can not be formed because gravitational attraction
and electromagnetic repulsion can not reach equilibrium \cite{RNclouds}.
Additional mirror should be placed at special location
to reflect the charged scalar wave.
We show that, for the specific set of
black hole and scalar field parameters, the clouds are only possible for
the specific mirror location $r_m$. For example,
for the fixed parameters of black hole and scalar field $M, Q, q, l$,
the discrete set of the mirror location $r_m$ is characterized by the
node number $n$ of the radial wave function. It is shown that,
the analytical results of mirror location $r_m$ for the clouds
are perfectly coincide with the numerical results in the region of $qQ\ll 1$.
However, the agreement becomes less impressive for $qQ=O(1)$ values.
In addition, we also show that
the massless scalar clouds are also possible when the mirror locations are
very close to the horizon. At last, we present an analytical calculation
of the specific mirror locations $r_m$ for the scalar clouds in the $qQ\gg 1$ regime.
\section*{ACKNOWLEDGEMENT}
The authors would like to thank Dr. Hongbao Zhang
for useful discussion on the numerical methods.
This work was supported by NSFC, China (Grant No. 11205048).
|
\section{Introduction}
Quantum key distribution (QKD) promises unconditional secure communication through insecure
communication channels \cite{QKD}.
In real world implementations of QKD, however, the achievable secret-key rates are still
relatively low compared with standard telecommunication rates.
The rates of secret-key generation are not only constrained by experimental imperfections,
which can be amended in principle, but are also limited by the fundamental features of quantum physics.
As recently shown in \cite{TGW}, the entanglement between the two
ends of the communication channel ultimately bounds the
maximum rate of secret-key generation:
\begin{equation}\label{TGW_1}
R \leq E_\mathrm{sq}(\mathcal{N}) \, ,
\end{equation}
where $E_\mathrm{sq}(\mathcal{N})$ is an entropic quantity called the squashed entanglement
of the channel \cite{squash}, which is function of the quantum communication channel $\mathcal{N}$
linking the legitimate sender Alice to the legitimate receiver Bob.
In this paper we consider the case where the communication channel $\mathcal{N}$
is a lossy (and noisy) bosonic channel. This means that information is encoded in a
collection of bosonic modes whose corresponding canonical operators are denoted
$a_j$, $a_j^\dag$ and satisfy the commutation relations $[ a_{j'} , a_j^\dag ] = \delta_{jj'}$.
In the Heisenberg picture the quantum channel maps the canonical operators $a_j$, $a_j^\dag$
to $a_j \to \sqrt{\eta} \, a_j + \sqrt{1-\eta} \, v_j$,
$a_j^\dag \to \sqrt{\eta} \, a_j^\dag + \sqrt{1-\eta} \, v_j^\dag$, where $\eta \in [0,1]$ is
the attenuation factor (also called transmissivity) and $v_j$, $v_j^\dag$ are the
canonical ladder operators of an environment bosonic mode.
The lossy channel is obtained if the environment mode is initially in the vacuum state,
while the lossy and noisy channel corresponds to the environment mode being in a thermal
state with $N_T$ mean photon number.
These channels attenuate the input power by a
factor $\eta$ and model the ubiquitous processes of linear absorption and scattering
of light.
When applied to the case of the lossy bosonic channel, the squashed entanglement bound in (\ref{TGW_1}) yields \cite{TGW}:
\begin{equation}\label{TGW_2}
R \leq \log{\left( \frac{1+\eta}{1-\eta}\right) } \, ,
\end{equation}
where the rate is measured in bits (throughout this paper $\log \equiv \log_2$)
per bosonic mode (given the bandwidth of the channel, this can be easily translated
in bits per second).
For both free space and fiber optics communication, the attenuation factor $\eta = e^{-\ell/\ell_0}$
scales exponentially with the distance $\ell$ between sender and receiver, where the characteristic length
$\ell_0$ depends on experimental conditions.
For long distances, $R \leq 2 \,\eta = 2 \, e^{-\ell/\ell_0}$, and
the key rate decays at least exponentially with increasing communication distance.
This result marks a striking difference between quantum-secured communication and (insecure)
classical communication. In the latter case, one can in principle achieve a finite
communication rate over arbitrarily long distances, just by sufficiently
increasing the signal power \cite{Gconj0}.
Unfortunately, this is not the case for quantum communication where a fundamental
rate-distance tradeoff exists, requiring the use of quantum repeaters to perform QKD
on long distances.
It is thus clear that to go around the fundamental rate-distance tradeoff in
(\ref{TGW_2}) one should renounce unconditionally security.
Here we discuss QKD conditioned on the assumption
that technological limitations allow an eavesdropper Eve to store quantum information reliably
only for a known and finite -- but otherwise arbitrarily long -- time.
Such an eavesdropper may also have unlimited computational power, including
a quantum computer.
Indeed, any physical realization of a quantum memory can reliably store
quantum information only for a time of the order of its coherence time.
We stress that we do not require the legitimate receiver to have better quantum
storage technologies than the eavesdropper.
As will be shown, the legitimate parties could have a much shorter memory time than
the eavesdropper and the communication will still be secure.
\section{Security definitions}
According to the state of the art, one requires a quantum cryptography protocol to be
unconditionally and composably secure.
Unconditional security means that one does not rely on unproven statements (e.g, about the complexity
of factorizing large numbers, or in general about the computational power of the eavesdropper).
Composable security means that the given protocol is secure also when used as a subroutine
within an overarching protocol \cite{compo}.
Suppose that a given communication protocol aims at establishing
a secret message described as a random variable $X$. The information about $X$ in the
hands of the eavesdropper Eve is described, without loss of generality, by a bipartite quantum state
of the form
\begin{equation}
\rho_{XE} = \sum_x p_X(x) \, |x\rangle \langle x| \otimes \rho_E(x) \, .
\end{equation}
Ideally, one would like Eve's state to be completely uncorrelated with
the message $X$, that is, $\rho_{XE} = \rho_X \otimes \rho_E$ \cite{NOTA_prod}.
To quantify the deviation from such an ideal setting one considers the trace distance \cite{NOTA_tdist}
\begin{equation}
D(\rho_{XE},\rho_X \otimes \rho_E) := \frac{1}{2} \, \| \rho_{XE} - \rho_X \otimes \rho_E \|_1 \, .
\end{equation}
Therefore, the security of the communication protocol is assessed by the condition
\begin{equation}
D(\rho_{XE},\rho_X \otimes \rho_E) \leq \epsilon \, ,
\end{equation}
which implies that the state $\rho_{XE}$ is indistinguishable, up to a probability
smaller than $\epsilon$, from the state $\rho_{X} \otimes \rho_E$, that is,
the given communication protocol is secure up to a probability smaller than $\epsilon$ \cite{Renner}.
As a matter of fact this criterion guarantees unconditional and composable security \cite{Renner}.
In this paper we renounce unconditional security and seek security conditioned on the
fact that the eavesdropper can store quantum information only for a finite and known time $\tau$.
This means that Eve is forced to make a measurement within a time $\tau$ after obtaining the
quantum state.
Suppose that Eve has made a measurement $\Lambda$ described by the POVM (positive operator valued measurement)
elements $\{ \Lambda_y \}_y$ \cite{NOTA_POVM}.
After the measurement has been made, the state has `collapsed' to
\begin{align}
\rho'_{XE} & = \sum_y \mathrm{Tr}_E( \rho_{XE} \, \mathbb{I} \otimes \Lambda_y ) \, |y\rangle_E \langle y | \\
& = \sum_{x,y} p_X(x) \, \mathrm{Tr} \left( \rho_{E}(x) \, \Lambda_y \right) \, |x\rangle \langle x | \otimes |y\rangle_E \langle y | \, .
\end{align}
Since $\rho'_{XE}$ is diagonal in the basis $\{ |x\rangle \otimes |y\rangle \}$, we have
\begin{align}
D(\rho'_{XE} , \rho'_X \otimes \rho'_E) & = \sum_{x,y} \left| p_{XY}(x,y) - p_X(x) p_Y(y) \right| \\
& =: D(p_{XY}, p_X p_Y) \, ,
\end{align}
where $p_{XY}(x,y) = p_X(x) p_{Y|x}(y)$ with $p_{Y|x}(y) = \mathrm{Tr} \left( \rho_{E}(x) \, \Lambda_y \right)$
and $p_Y(y) = \sum_x p_X(x) p_{Y|x}(y)$, that is, the trace distance equals
the distance between classical probabilities.
Finally, optimizing over Eve's choice of her measurement, we obtain the following
security condition:
\begin{equation}\label{cond1}
\sup_{\Lambda} D(p_{XY}, p_X p_Y) \leq \epsilon \, .
\end{equation}
In this paper, instead of working directly with condition (\ref{cond1}), we require
\begin{equation}
I_\mathrm{acc}(X;E)_{\rho} \leq \epsilon' \, ,
\end{equation}
where $I_\mathrm{acc}(X;E)_{\rho}$ denotes the accessible information of Eve about $X$
given the state $\rho_{XE}$ \cite{NOTA_acc}.
The latter implies condition (\ref{cond1}), for $\epsilon = \sqrt{ 2\ln{(2)} \, \epsilon' }$,
via Pinsker inequality \cite{NOTA_Pinsker}
\begin{equation}
\max_\Lambda D(p_{XY}, p_X p_Y) \leq \sqrt{ 2\ln{(2)} \, I_\mathrm{acc}(X;E)_{\rho} } \, .
\end{equation}
It is worth recalling that accessible information was used as a security quantifier
during the first years of quantum cryptography, since it was found that a security criterion
based on the accessible information does not in general guarantee composable security in an unconditional
manner \cite{Renner}.
Here instead we have shown that composability holds under condition (\ref{cond1})
if we give up full unconditional security and seek security under the assumption that the
eavesdropper can store quantum information only for a finite and known time ---
i.e, she has a quantum memory with limited storage time.
\section{Summary of the results}
We present two novel key-generation protocols for continuous-variable quantum optical communication
through a lossy bosonic channel with transmissivity $\eta$, modeling linear attenuation
and scattering.
These protocols are composably secure under the condition that Eve's as a quantum memory
with finite, and known, but otherwise arbitrarily long, storage time.
The first protocol is a direct-reconciliation protocol (in which we allow information reconciliation
by forward public communication from the sender Alice to the receiver Bob).
We obtain a simple formula for the asymptotic key rate (see Fig.\ \ref{fig:direct}):
\begin{equation}\label{rate-dr-infty}
r_\mathrm{dr} = 1 + \log{\left(\frac{\eta}{1-\eta}\right)} \, .
\end{equation}
This protocol can generate a nonzero key rate for any $\eta > 1/3$.
By comparison, the maximum unconditionally secure key rate from direct reconciliation
is given by the quantum capacity formula $\log{\left( \frac{\eta}{1-\eta} \right)}$ \cite{Wolf}
and is positive only for $\eta > 1/2$ \cite{sext}.
\begin{figure}
\centering
\includegraphics[width=0.35\textwidth]{plot_direct}
\caption{
Achievable key rate for the pure loss channel ($N_T=0$) vs the channel transmissivity $\eta$, in bits per mode,
for direct reconciliation protocols.
Blue solid line: Achievable locked-key rate as given by the expression in (\ref{rate-dr-infty}).
Red dashed line: Maximum fully unconditional secret-key rate,
given by the expression $\max\{ 0 , \log{\left( \frac{\eta}{1-\eta} \right)} \}$ \cite{Wolf}.}
\label{fig:direct}
\end{figure}
The second protocol is a reverse-reconciliation protocol (we allow information reconciliation
by backward public communication from Bob to Alice). In this setting we show that Alice and
Bob can in principle generate key at an asymptotic rate of more than $1$ bit per bosonic mode sent
through the channel. This is true for any nonzero value of the transmissivity $\eta$, provided sufficient
input energy is provided --- hence reproducing the feature of insecure classical communication in a
quantum-secured communication framework. The achievable asymptotic key rate is (see Fig.\ \ref{fig:reverse})
\begin{equation}\label{lkrr}
r_\mathrm{rr} = 1 + \log{\left(\frac{1}{1-\eta}\right)} \, .
\end{equation}
By comparison, the maximum fully unconditional key rate is upper bounded by the expression in
(\ref{TGW_2}) and vanishes as $2\eta$ for small values of $\eta$.
\begin{figure}
\centering
\includegraphics[width=0.35\textwidth]{plot_reverse}
\caption{
Achievable key rate for the pure loss channel ($N_T=0$) vs the channel transmissivity $\eta$, in bits per mode,
for reverse reconciliation protocols.
Blue solid line: Achievable locked-key rate as from the expression in (\ref{lkrr}).
Red dashed line: Upper bound for the secret-key rate (assisted by two-way public communication),
given by the expression in (\ref{TGW_2}).
Yellow dash-dotted line: Achievable asymptotic secret-key rate
according to the standard security definition as given by the reverse
coherence information $\log{\left( \frac{1}{1-\eta} \right)}$ \cite{Pirs}.}
\label{fig:reverse}
\end{figure}
We also consider the case of lossy and noisy bosonic channel, which models the presence of experimental imperfection
or a thermal-like background with $N_T$ mean photons per mode.
The lossy and noisy channel is also used to model an `active attack' from the eavesdropper, who
injects noise in the channel.
In this case we obtain an asymptotic rate equal to
\begin{equation}\label{thermal}
r_\mathrm{rr} = 1 + \log{\left( \frac{1}{1-\eta} \right) } - g(N_T) \, ,
\end{equation}
which is nonzero at arbitrary distances provided $N_T \lesssim 0.3$ (see Fig.\ \ref{fig:eccess})
\begin{figure}
\centering
\includegraphics[width=0.35\textwidth]{plot_aattack}
\caption{
Tolerable excess noise $N_T$ vs the transmissivity $\eta$
for the reverse-reconciliation quantum data locking protocol, from Eq.\ (\ref{thermal}).
The asymptotic locked-key generation rate is nonzero for values of
$(\eta,N_T)$ below the curve.}
\label{fig:eccess}
\end{figure}
These protocols are instances of quantum data locking protocols (see Sec.\ \ref{sec:QDL}).
We henceforth call {\it locked key} a key which is generated by a quantum data locking
protocol, just to remind us that this key is not unconditionally secure, but secure conditioned on
the assumption of finite memory storage time.
\section{Comparison with other models}
It is known that high rates of secret-key generation can be attained
against an eavesdropper endowed with an imperfect quantum memory, as for
example in the Bounded Storage Model, where Eve can store only a constrained
number of qubits (see e.g.\ \cite{BSM}).
Even under bounded storage, no known protocol attains a constant rate as a function of distance.
Elsewhere we have shown that quantum data locking allows for a substantial
enhancement of the key rate \cite{PRL,NJP}.
Here we show for the first time that such an assumption allows us to
generate key at a constant rate across virtually any distance.
It is an open question whether the quantum data locking could be applied
in the bounded storage model to attain rates of key generation independent
on the distance.
Our results must be compared with the bounds on the optimal secret-key
rate obtained requiring fully unconditional security.
In the asymptotic setting, the security is usually quantified by the
quantum mutual information (see e.g. \cite{PrivateD}).
The gain in key generation rate that we achieve
follows from the existence of a large gap between the quantum mutual information and
the accessible information of the adversary. This gap is well known in quantum
information theory: it is the {\it quantum discord} \cite{QD},
which quantifies the quantum correlations that the adversary cannot access by local measurements
on her share of the quantum system.
\section{Quantum data locking and quantum enigma machines}\label{sec:QDL}
In a typical quantum data locking protocol \cite{QDL,CMP,Buhrman,Leung}, the two legitimate parties,
say Alice and Bob, publicly agree on a set of $MK$ quantum codewords.
They then use a preshared secret key of $\log{K}$ bits, labeled by $s=1,2,\dots,K$, to secretly agree on a set of $M$
(equally probable) codewords, labeled by $x=1,2,\dots,M$, used to
encode $\log{M}$ bits of classical information.
These quantum codewords are sent through $n$ uses of a quantum channel from Alice to Bob.
Suppose an eavesdropper Eve tampers with the communication line and obtains one of the states
$\rho^n_E(x,s)$.
The correlations between Eve's quantum system and the input message $x$ are described
by the state
\begin{equation}
\rho^n_{XE} = \frac{1}{M} \sum_{x=1}^M |x\rangle\langle x| \otimes \frac{1}{K}\sum_{s=1}^K \rho^n_E(x,s) \, ,
\end{equation}
where $\{ |x\rangle \}_{x=1,\dots,M}$ is an orthonormal basis for an auxiliary quantum
system encoding the messages $x$ --- notice that the summation over $s$ comes from the fact that
Eve does not know the value of the secret key.
One can prove that, if the states $\rho^n_E(x,s)$ have a suitable form and for $K$ large enough, Eve
can only obtain a negligible amount of the classical information
--- as quantified by the accessible information --- carried by the label $x$.
In the most powerful quantum data locking schemes known up to now, a constant-size
preshared secret seed of about $\log{K} = \log{1/\epsilon}$ bits
allows Alice and Bob to encrypt $\log{M}$ bits (with $M$ arbitrarily large), with the guarantee
that Eve's accessible information is of the order of $\epsilon \log{M}$ bits \cite{Fawzi,Dupuis,phase}.
It is worth remarking that quantum data locking provides a strongest violations of
classical information theory in the quantum setting. Indeed, according to a famous theorem
of Shannon's, which assesses the security of one-time pad encryption,
to encrypt $m$ bits of classical information Alice and Bob need at least $m$ bits
of preshared secret key \cite{Shannon}.
Quantum data locking violates this Shannon's result by an exponential amount.
A quantum data locking protocol can be seen as a quantum counterpart of the twentieth century
Enigma machine \cite{QEM}. Following \cite{QEM,PRX} we call `quantum enigma machine' an
optical cipher that harnesses the quantum data locking effect.
\subsection{Quantum bootstrapping}\label{sec:boot}
The first works on quantum data locking only considered the ideal case of a noiseless communication scenario.
Only recently the quantum data locking effect has been considered in a noisy setting \cite{QEM,PRX,AW} (see also \cite{Boixo}).
Here we combine quantum data locking with a key-recycling technique
that has been successfully applied to quantum data locking in a noisy communication scenario \cite{PRL,QDL,Entropy}.
We assume that eavesdropper Eve and the legitimate receiver can store quantum information
for a time $\tau_E$ and $\tau_B$, respectively.
Suppose then that Alice and Bob, using the quantum channel $n$ times, run a
quantum data locking protocol to communicate $\log{M} = n \chi$ bits of classical information,
and consume $\log{K} = n k$ bits of preshared secret key.
Bob may need to perform a collective measurement over $n$ quantum systems in order to decode.
Since, as from our assumption, Bob's quantum memory can store quantum information
only for times shorter than $\tau_B$, this requires that the $n$ quantum signals should be sent
within this time interval (this is always possible for $\tau_B$ large enough or by
increasing the repetition rate).
On the other hand, if Eve has a quantum memory with finite coherence time $\tau_E$, this implies that
she is forced to measure within a time $\tau_E$ after receiving the signals,
otherwise her memory will decohere anyway.
Therefore, what the legitimate parties Alice and Bob can do is to wait for a time longer than $\tau_E$
before sending more information through the channel.
After waiting such a time, Alice and Bob can safely recycle part of the obtained key
as a fresh key to run another round of quantum data locking.
Thus, for $\chi > k$, Alice and Bob can recycle part of the newly
established key and use it as a seed for another round of quantum data locking.
By repeating this procedure many times they will asymptotically obtain a overall locked-key rate of
$r = \chi - k$ bits per channel use, with a negligible amount of initially shared secret key.
While $r = \chi - k$ is the rate of bits per channel use, one could expect a lower
rate in terms of bits per second, due to the waiting times between quantum data locking subroutines.
There is a simple strategy to solve this problem: Alice and Bob can use the dead times to
run two (or more) independent quantum data locking protocols.
In this way they can in principle
achieve a rate of bits per second as high as $r \nu = (\chi - k) \nu$, where $\nu$ is the number of channel
uses per second.
Notice that this holds for any value of $\tau_E$, as long as it is known to Alice and Bob,
and independently of $\tau_B$ (for instance we can take $\tau_B = \tau_E$ or even $\tau_B < \tau_E$).
\section{The direct reconciliation protocol}\label{sec:dir}
Alice prepares multimode coherent states that encode
both the input message $x \in \{ 1,\dots, M\}$ and the value
of the secret seeds $s \in \{ 1, \dots, K\}$ she shares with Bob.
The encoding is by a random code (whose codebook is public) that assigns to each pair $(x,s)$
an $n$-mode coherent states
\begin{equation}
|\alpha^n(x,s)\rangle = \bigotimes_{j=1}^n |\alpha_j(x,s)\rangle \, ,
\end{equation}
where $\alpha_j(x,s)$ is the amplitude of the coherent state of the $j$-th bosonic mode
sent through the channel.
This is schematically depicted in Fig.\ \ref{fig:ddirect}, where the lossty channel is represented
as a beam-plitter.
To construct the random code, the amplitudes of the coherent states
are independently drawn from a circularly symmetric Gaussian distribution,
denoted $G_{(0,N)}$, with zero mean and mean photon number $\int d^2\alpha \, |\alpha|^2 \, G_{(0,N)} = N$.
The receiver Bob obtains the attenuated coherent states
\begin{equation}
|\sqrt{\eta} \, \alpha^n(x,s)\rangle = \bigotimes_{j=1}^n |\sqrt{\eta} \, \alpha_j(x,s)\rangle \, .
\end{equation}
The goal of Bob, who knows the value of $s$, is to decode $x$.
It is known that he can do that (with asymptotically negligible error)
with an asymptotic bit-rate for $x$ given by \cite{Gconj0}
\begin{equation}
\chi_\mathrm{dr} := \lim_{n\to\infty} \frac{\log{M}}{n} = g(\eta N) \, ,
\end{equation}
where
\begin{equation}
g(N) = (N+1)\log{(N+1)} - N \log{N} \, .
\end{equation}
To guarantee the security of the communication protocol, we have to bound Eve's
accessible information.
For any $x$ and $s$, Eve obtains the attenuated coherent states
\begin{equation}
|\sqrt{1-\eta} \, \alpha^n(x,s)\rangle = \bigotimes_{j=1}^n |\sqrt{1-\eta} \, \alpha_j(x,s)\rangle \, .
\end{equation}
We can show (see Sec.\ \ref{sec:proofs}) that Eve's accessible information about Alice's input message $x$
is negligibly small, provided Alice and Bob
initially share enough bits of secret key.
For $N$ large enough, and asymptotically in $n$, this is achieved for
\begin{equation}
k_\mathrm{dr} := \lim_{n\to\infty} \frac{ \log{K} }{n} = 2 g[(1-\eta)N] - g[2(1-\eta)N] \, .
\end{equation}
Applying the key-bootstrapping routine (see Sec.\ \ref{sec:boot}),
this yields a net asymptotic locked-key generation rate of
\begin{equation}
r_\mathrm{dr} = \chi_\mathrm{dr} - k_\mathrm{dr}
= g(\eta N) - 2 g[(1-\eta)N] + g[2(1-\eta)N] \, ,
\end{equation}
which in the limit of $N \to \infty$ becomes
\begin{equation}
r_\mathrm{dr} = 1 + \log{ \left( \frac{\eta}{1-\eta} \right) } \, .
\end{equation}
\begin{figure}
\centering
\includegraphics[width=0.4\textwidth]{fig_direct}
\caption{The lossy bosonic channel can be modeled as a beam-splitter with
transmissivity $\eta$ and the environment mode initially in the vacuum state.
In the direct reconciliation protocol, Alice sends coherent state down the channel.}
\label{fig:ddirect}
\end{figure}
\section{The reverse reconciliation protocol}\label{sec:rev}
In the first phase of the protocol Alice prepares $n$ instances of a two-mode squeezed vacuum state,
with $N$ mean photons per mode, that is, $\rho^n_{AA'} = \rho^{\otimes n}_{AA'}$
with
\begin{equation}
\rho_{AA'} = |\zeta_N\rangle_{AA'}\langle\zeta_N|
\end{equation}
and
\begin{equation}
|\zeta_N\rangle_{AA'} = \frac{1}{\sqrt{N+1}} \sum_{\ell=0}^\infty \left( \frac{N}{N+1} \right)^{\ell/2} |\ell\rangle_{A}|\ell\rangle_{A'} \, ,
\end{equation}
where $|\ell\rangle$ denotes the photon-number state with $\ell$ photons.
Alice keeps the modes labeled with `$A$' and sends through $n$ uses of a lossy bosonic channel those labeled with `$A'$',
see Fig.\ \ref{fig:dreverse}.
At the end of this first phase of the communication protocol, Alice, Bob and Eve share the $3n$-mode
state $\rho^n_{ABE} = \rho^{\otimes n}_{ABE}$, where $\rho_{ABE}$ is a $3$-mode Gaussian state
with zero mean and covariance matrix $V_{ABE}$ (whose explicit form is given in the Appendix).
In the second phase of the communication protocol, Bob makes a collective measurement
on his share of $n$ bosonic modes, described by the state $\rho^n_B = \rho_B^{\otimes n}$,
where $\rho_B$ is a Gaussian state with zero mean and variance $V_B$ (see Appendix for details).
Indeed, Bob applies a measurement $\Gamma(s)$ chosen from a set of measurements
parameterized by the label $s = 1, \dots, K$.
The value of $s$ is determined by the secret key he shares with Alice.
That is, while the list of possible $K$ measurement is public and hence known to
Eve, the specific choice of $\Gamma(s)$ is known only by Alice and Bob.
Bob's measurement is defined as follows.
First, Alice and Bob publicly agree on a set of $MK$ $n$-mode coherent states
\begin{equation}
|\beta^n(x,s)\rangle = \bigotimes_{j=1}^n |\beta_j(x,s)\rangle \, ,
\end{equation}
for $x = 1, \dots, M$ and $s = 1, \dots, K$.
These coherent states are defined by sampling the
amplitudes $\beta_j(x,s)$ i.i.d.\ from a circularly symmetric Gaussian distribution with zero mean and
variance $\eta N$.
For any given $s$, we consider the sliced operator
\begin{equation}
\Sigma(s) = \sum_{x=1}^{M} \mathbb{P}^n_B \, |\beta^n(x,s)\rangle \langle \beta^n(x,s)| \, \mathbb{P}^n_B \, ,
\end{equation}
where $\mathbb{P}^n_B$ is the projector on the strongly $\delta$-typical subspace defined by $\rho_B^{\otimes n}$ (see, e.g.\ \cite{Wilde}).
Applying the operator Chernoff bound (see Appendix for details) we obtain that the bounds
\begin{equation}
(1-\epsilon) M 2^{-n g(\eta N)} \mathbb{P}^n_B \leq \Sigma(s) \leq (1+\epsilon) M 2^{-n g(\eta N)} \mathbb{P}^n_B
\end{equation}
hold true with arbitrarily high probability provided $M \gg 2^{n g(\eta N)}$.
It follows that for any given $s$ the operators
\begin{equation}\label{gamma}
\Gamma_{x}(s) = \frac{ \mathbb{P}^n_B \, |\beta^n(x,s)\rangle \langle \beta^n(x,s)| \, \mathbb{P}^n_B }{(1+\epsilon) M 2^{-n g(\eta N)} }
\end{equation}
define a subnormalized POVM in Bob's typical subspace, which can be completed by
introducing the operator $\Gamma_{0}(s) = \mathbb{P}^n_B - \sum_{x} \Gamma_{x}(s)$.
In this way we have defined Bob's measurement $\Gamma(s)$ for all values of $s$.
After performing the measurement, Bob declares an error if he obtains the
measurement output corresponding to $\Gamma_{0}(s)$. This event, however, happens
with a negligible probability (see Appendix for details).
In the third phase of the protocol, Alice makes a measurement on her share of bosonic modes.
For a given value of $s$ (which is known to Alice and Bob) and $x$, we consider Alice's
conditional state $\rho^n_A(x,s)$.
As a matter of fact, Bob's measurement induces a virtual backward communication channel
from Bob to Alice.
As a result, for given $s$, Alice obtains an ensemble of states $\{ \rho^n_A(x,s) , p(x,s) \}_{x=1,\dots,M}$,
where $p(x,s) = \mathrm{Tr}(\Gamma_{x}(s) \rho^n_B(s))$.
The maximum amount of classical information (per mode) about $x$ that Alice can extract from this ensemble of states
is given, in the asymptotic setting, by the associated Holevo information \cite{Holevo} \cite{NOTA_superb}:
\begin{equation}
\chi_\mathrm{rr} = \frac{1}{n} \left[ S(\rho^n_A) - \sum_{x} p(x,s) S(\rho^n_A(x,s)) \right] \, ,
\end{equation}
where $S(\rho) = - \mathrm{Tr}\left( \rho \log{\rho}\right)$ denotes the von Neumann entropy.
From the explicit expressions for $p(x,s)$, $\rho^n_A(x,s)$ and $\rho^n_A$ (given in the Appendix)
we obtain
\begin{equation}
\chi_\mathrm{rr} = g(N) - g[(1-\eta)N']
\end{equation}
where $N' = N/(1+\eta N)$.
$\chi_\mathrm{rr}$ also quantifies the rate (in bits per mode) of shared randomness
that can be established, with the assistance of public communication, by Alice and Bob \cite{errcorr}.
Finally, to show the security of the communication protocol, we need to bound Eve's
accessible information about $x$.
Bob's measurement also induces a virtual quantum channel to Eve.
For any given $s$, the ensemble of states obtained by Eve is $\{ \rho^n_E(x,s) , p(x,s) \}_{x=1,\dots,M}$,
where $\rho^n_E(x,s)$ is Eve's state conditioned on Bob's measurement result $x$.
Given the explicit form of $\rho^n_E(x,s)$ we show (see Sec.\ \ref{sec:proofs} and the Appendix) that Eve's accessible
information about $x$ is negligibly small for $K$ such that
\begin{align}
k_\mathrm{rr} & := \lim_n \frac{\log{K}}{n } \\
& = 2 g[(1-\eta)N] - g[(1-\eta)N'] - g[(1-\eta)N''] \, ,
\end{align}
with $N'' = N(1+2\eta N)/(1+\eta N)$.
In conclusion, applying the bootstrapping routine, we obtain a net rate of locked-key generation of (in bits per mode)
\begin{equation}
r_\mathrm{rr} = \chi_\mathrm{rr} - k_\mathrm{rr}
= g(N) - 2 g[(1-\eta)N] + g[(1-\eta)N''] \, ,
\end{equation}
which in the limit of $N \to \infty$ reads
\begin{equation}
r_\mathrm{rr} = 1 + \log{\left( \frac{1}{1-\eta} \right) } \, .
\end{equation}
\begin{figure}
\centering
\includegraphics[width=0.4\textwidth]{fig_reverse}
\caption{The lossy bosonic channel can be modeled as a beam-splitter with
transmissivity $\eta$ and the environment mode initially in the vacuum state.
In the first phase of the reverse reconciliation protocol, Alice sends one mode of a two-mode
entangled state (denoted by the symbol `$\infty$') down the channel.}
\label{fig:dreverse}
\end{figure}
Similar results are obtained if the channel from Alice to Bob is
lossy and noisy. In this case the reverse reconciliation protocol
achieves an asymptotic locked-key rate of
\begin{equation}
r_\mathrm{rr} = 1 + \log{\left( \frac{1}{1-\eta} \right) } - g(N_T) \, ,
\end{equation}
where $N_T$ is the mean number of thermal photons per mode in the channel.
\section{Security proofs}\label{sec:proofs}
We discuss in details the case of the lossy channel.
The proof for the lossy and noisy channel can be obtained in a similar way.
The starting point of the proof are some mathematical tools presented in \cite{NJP}.
There we assumed that Eve's states $\rho^n_E(x,s)$ belongs to a finite-dimensional
space of dimension $d^n$.
Given the bipartite state
\begin{equation}
\rho^n_{XE} = \frac{1}{M} \sum_{x=1}^M |x\rangle\langle x| \otimes \frac{1}{K}\sum_{s=1}^K \rho^n_E(x,s) \, ,
\end{equation}
the following bound hold for the associated accessible information (see~\cite{NJP}):
\begin{equation}\label{Iaccn1}
I_\mathrm{acc
\leq \log{M} - \frac{d^n}{M} \, \min_{|\phi\rangle} \left\{ H[Q(\phi)] - \eta\left[ \sum_{x=1}^M Q_x(\phi)\right] \right\} \, ,
\end{equation}
where
\begin{equation}
Q_x(\phi) = \frac{1}{K} \sum_{s=1}^K \langle \phi | \rho^n_E(x,s) | \phi \rangle \, ,
\end{equation}
$H[Q(\phi)] = \sum_{x=1}^M \eta( Q_x(\phi) )$,
with $\eta( \, \cdot \, ) = - ( \, \cdot \, ) \log{( \, \cdot \, )}$.
The minimum is over all vectors $\phi$ in Eve's $d^n$-dimensional Hilbert space.
As shown in \cite{NJP}, if the ensemble of states from which the codewords
are sampled is such that for any unit vector $|\phi\rangle$,
\begin{equation}\label{1moment}
\mu := \mathbb{E}_s[ \langle \phi | \rho^n_E(x,s) | \phi \rangle ] = \frac{1}{d^n}
\end{equation}
($\mathbb{E}_s$ denotes the expectation value over $s$),
and
\begin{align}
\Sigma & := \mathbb{E}_s[ \langle \phi | \rho^n_E(x,s) | \phi \rangle^2] \nonumber \\
& = \mathbb{E}_s[ \langle \phi,\phi | \rho^n_E(x,s) \otimes \rho^n_E(x,s) | \phi,\phi \rangle ] \label{2moment} \, ,
\end{align}
(here $|\phi,\phi\rangle \equiv |\phi\rangle \otimes |\phi\rangle$)
then the right hand side of (\ref{Iaccn1}) is smaller than $\epsilon \log{M}$
provided that
\begin{equation}\label{Kqdl}
K > \max\left\{ 2 \gamma^n \left( \frac{1}{\epsilon^2} \ln{M} + \frac{2}{\epsilon^3} \ln{\frac{5}{\epsilon}}\right) ,
\frac{d^n}{M} \, \frac{4 \ln{2} \ln{d^n}}{\epsilon^2 }
\right\} \, ,
\end{equation}
with
\begin{equation}
\gamma^n = \frac{\Sigma}{\mu^2} \, .
\end{equation}
In our setting $n$ counts the number of modes employed in one quantum data locking routine.
Putting $M = 2^{n \chi}$ and $\epsilon = e^{-n^c}$ with $c \in (0,1)$, condition
(\ref{Kqdl}) yields an asymptotic rate of secret-key consumption (in bits per mode)
\begin{equation}
k = \lim_{n\to\infty} \frac{1}{n} \log{K} = \max\left\{ \log{\gamma} , \log{d}-\chi \right\} \, .
\end{equation}
In our continuous-variable setting, Eve's space is infinite-dimensional.
Therefore, to apply the result of \cite{NJP} we need to map Eve's space into
a finite dimensional one.
In both the direct and reverse reconciliation protocol, the expectation value over $s$
of the state of Eve has the form (see details in the Appendix)
\begin{equation}
\rho^n_E = \mathbb{E}_s [ \rho_E^n(x,s) ] = \rho_E^{\otimes n} \, ,
\end{equation}
that is, the average state is a direct product.
In particular, $\rho_E$ is a Gaussian state with zero mean, variance $V_E$,
and mean photon number $(1-\eta) N$.
We can hence consider the $\delta$-typical subspace projector $\mathbb{P}^n_\rho$
associated with $\rho_E^{\otimes n}$.
We use this projector to define an auxiliary bipartite state of the form
\begin{equation}
\sigma^n_{XE} = \frac{1}{M} \sum_{x=1}^M |x\rangle\langle x| \otimes \frac{1}{K}\sum_{s=1}^K \sigma^n_E(x,s) \, ,
\end{equation}
where
\begin{equation}
\sigma^n_E(x,s) = \mathbb{P}^n_\rho \, \rho^n_E(x,s) \, \mathbb{P}^n_\rho
\end{equation}
is obtained by slicing with the $\delta$-typical subspace projector.
From the properties of the typical projector we have
\begin{equation}
\| \sigma^n_{XE} - \rho^n_{XE} \|_1 \leq \delta \, .
\end{equation}
Since the two states are $\delta$-close in trace-norm, the security of the
state $\rho^n_{XE}$ follows, up to a probability $\delta$, from that
of $\sigma^n_{XE}$.
In such a way we have reduced the problem to a finite dimensional one,
where the dimension is that of the $\delta$-typical subspace, i.e.,
\begin{equation}
d^n := \mathrm{Tr}\left( \mathbb{P}^n_\rho \right) \in [ 2^{n [S(\rho_E) - c \delta]} , 2^{n [S(\rho_E) + c \delta]}]
\end{equation}
(for some constant $c$).
We use a notion of typical subspace that is a slightly different
from the one usually considered (see for instance \cite{Wilde}).
Given a hermitian operator $\xi$ we consider its spectral decomposition
\begin{equation}
\xi = \sum_\ell p_\ell \, P_\ell \, ,
\end{equation}
where the sum is over the eigenvalues $p_\ell$ and the corresponding eigenprojectors $P_\ell$,
in such a way that $p_\ell \neq p_{\ell'}$ for $\ell \neq \ell'$ (that is, $\mathrm{Tr} \left( P_\ell \right)$
equals the degeneracy of $p_\ell$).
We look at each projector $P_\ell$ as an event whose probability is $\pi_\ell = p_\ell \, \mathrm{Tr} \left( P_\ell \right)$.
Given $\xi^{\otimes n}$, we then define the $\delta$-typical projector $\mathbb{P}^n_\xi$ as
(we omit the subscript $\delta$ to simplify the notation)
\begin{equation}
\mathbb{P}^n_\xi = \sum_{p_{\ell_1}p_{\ell_2}\cdots p_{\ell_n} \in T_\delta^n} P_{\ell_1} \otimes P_{\ell_2} \otimes \cdots \otimes P_{\ell_n}
\end{equation}
where the sum is over the sequences $p_{\ell_1}p_{\ell_2}\cdots p_{\ell_n}$ which are
$\delta$-typical with respect to the probability distribution $\pi_\ell$.
Notice that this construction of the typical projector coincides with the
usual one when all the eigenvalues of $\xi$ are non degenerate.
First we compute (\ref{1moment}):
\begin{align}
\mu & = \mathbb{E}_s[ \langle \phi | \, \sigma^n_E(x,s) \, | \phi \rangle ] \\
& = \mathbb{E}_s[ \langle \phi | \, \mathbb{P}^n_\rho \, \rho^n_E(x,s) \, \mathbb{P}^n_\rho \, | \phi \rangle ] \\
& = \langle \phi | \, \mathbb{P}^n_\rho \, \rho_E^{\otimes n} \, \mathbb{P}^n_\rho \, | \phi \rangle \, .
\end{align}
Then, from the equipartition properties of the $\delta$-typical subspace we have
(for some constant $c$)
\begin{equation}
2^{-n [ S(\rho_E) + c\delta ]} \leq \mu \leq 2^{-n [ S(\rho_E) - c\delta]} \, .
\end{equation}
To compute (\ref{2moment}) we need to introduce another typical subspace projector.
We consider the state $( \rho_E \otimes \rho_E)^{\otimes n}$ and its associated
$(2\delta)$-typical subspace projector, denoted as $\mathbb{P}^n_{\rho \otimes \rho}$.
Notice that $[ \mathbb{P}_\rho^n \otimes \mathbb{P}_\rho^n , \mathbb{P}^n_{\rho \otimes \rho} ] = 0$, and that
$\mathbb{P}_\rho^n \otimes \mathbb{P}_\rho^n \leq \mathbb{P}^n_{\rho \otimes \rho}$.
We also consider the state
\begin{equation}
\rho_{2E}^n := \mathbb{E}_s[ \rho^n_E(x,s) \otimes \rho^n_E(x,s) ] = \rho_{2E}^{\otimes n} \, .
\end{equation}
By explicit computation (see Appendix) we can show that, in
both the direct and reverse reconciliation protocols, $\rho_{2E}$ is a Gaussian state
with zero mean and covariance matrix $V_{2E}$.
Moreover, $\rho_{2E}$ commutes with $\rho_E \otimes \rho_E$
since they are both diagonal in the photon-number basis (see Appendix).
It follows that $\rho_{2E}$ also commutes with $\mathbb{P}^n_{\rho \otimes \rho}$.
We also have that, given that $\rho_E$ has mean photon number $(1-\eta)N$, then both
$\rho_{2E}$ and $\rho_E \otimes \rho_E$ have $2(1-\eta)N$ mean photons.
We can now compute (\ref{2moment}):
\begin{align}
\Sigma & = \mathbb{E}_s[ \langle \phi, \phi | \, \sigma^n_E(x,s) \otimes \sigma^n_E(x,s) \, | \phi, \phi \rangle ] \\
& = \mathbb{E}_s[
\langle \phi, \phi | \, {\mathbb{P}_\rho^n}^{\otimes 2} \, \rho^n_E(x,s) \otimes \rho^n_E(x,s) \, {\mathbb{P}_\rho^n}^{\otimes 2} \, | \phi, \phi \rangle
] \\
& = \langle \phi, \phi | \, {\mathbb{P}_\rho^n}^{\otimes 2} \, \rho_{2E}^{\otimes n} \, {\mathbb{P}_\rho^n}^{\otimes 2} \, | \phi, \phi \rangle \, .
\end{align}
Since ${\mathbb{P}_\rho^n}^{\otimes 2}$ commutes with $\mathbb{P}_{\rho\otimes\rho}^n$ and
${\mathbb{P}_\rho^n}^{\otimes 2} \leq \mathbb{P}_{\rho\otimes\rho}^n$, we have
\begin{align}
\Sigma & \leq \langle \phi, \phi | \,
\mathbb{P}_{\rho\otimes\rho}^n \, \rho_{2E}^{\otimes n} \, \mathbb{P}_{\rho\otimes\rho}^n
\, | \phi, \phi \rangle \, .
\end{align}
To conclude, let us consider the sliced operator
$\mathbb{P}_{\rho\otimes\rho}^n \, \rho_{2E}^{\otimes n} \, \mathbb{P}_{\rho\otimes\rho}^n$.
Since $[ \mathbb{P}_{\rho\otimes\rho}^n , \rho_{2E}^{\otimes n} ] = 0$ we can apply
a classical argument concerning typical type classes (see, e.g., \cite{CT}).
Let us denote as $q_\ell$ the eigenvalues of $\rho_{2E}$.
We notice that the eigenvectors of $\mathbb{P}_{\rho\otimes\rho}^n \, \rho_{2E}^{\otimes n} \, \mathbb{P}_{\rho\otimes\rho}^n$
are those of $\rho_{2E}^{\otimes n}$ which are in the range of $\mathbb{P}_{\rho\otimes\rho}^n$
(that is, they are $\delta$-typical for $(\rho_E \otimes \rho_E)^{\otimes n}$).
Consider then an eigenvector whose $\delta$-typical type is $\tilde\pi$, the corresponding
eigenvalue of $\mathbb{P}_{\rho\otimes\rho}^n \, \rho_{2E}^{\otimes n} \, \mathbb{P}_{\rho\otimes\rho}^n$ is
\begin{equation}
w = \prod_\ell q_\ell^{n \tilde\pi_\ell} = 2^{n \sum_\ell \tilde\pi_\ell \log{q_\ell}} \, .
\end{equation}
Being $\rho_{2E}$ a zero-mean, thermal-like, Gaussian state, $q_\ell = Z^{-1} 2^{-\beta \ell}$, where $\ell$
is the photon number. This yields
\begin{align}
w & = 2^{- n \left( \beta \langle \ell \rangle_{\tilde\pi} + \log{Z} \right)} \\
& = 2^{- n \left( \beta 2(1-\eta)N + \log{Z} + \beta \Delta \langle \ell \rangle \right)} \\
& = 2^{- n \left( S(\rho_{2E}) + \beta \Delta \langle \ell \rangle \right)} \, .
\end{align}
Here $\langle \ell \rangle_{\tilde\pi} = \sum_\ell \tilde\pi_\ell \ell$ is the mean photon
number given by the $\delta$-typical distribution $\tilde\pi$.
Since $\tilde\pi$ is $\delta$-typical for $(\rho_E \otimes \rho_E)^{\otimes n}$,
we expect $\langle \ell \rangle_{\tilde\pi} = 2(1-\eta)N$,
$\Delta \langle \ell \rangle = \langle \ell \rangle_{\tilde\pi} - 2(1-\eta)N$
being the fluctuation about the expectation value.
Finally we have used
$S(\rho_{2E}) = \beta 2(1-\eta)N + \log{Z}$.
In the Appendix we show that, for a $\delta$-typical type $\tilde\pi$,
\begin{equation}
| \beta \Delta \langle \ell \rangle | \leq 2 c \delta [ (1-\eta)N + 1]
\end{equation}
(for some constant $c$), from which we obtain
\begin{equation}
w \leq 2^{- n \left( S(\rho_{2E}) + 2 c \delta [(1-\eta)N +1] \right)} \, ,
\end{equation}
and hence
\begin{equation}
\Sigma \leq 2^{- n \left( S(\rho_{2E}) + 2 c \delta [(1-\eta)N +1] \right)} \, .
\end{equation}
From these results, in the limits that $n \to \infty$ and $\delta \to 0$,
we obtain the following bound on the key consumption rate:
\begin{equation}
k = \max\left\{
2 S(\rho_E) - S(\rho_{2E}) ,
S(\rho_E) - \chi
\right\} \, .
\end{equation}
For the direct reconciliation protocol we have (see derivation in Appendix):
$\chi = g(\eta N)$,
$S(\rho_E) = g((1-\eta) N)$, and
$S(\rho_{2E}) = g(2(1-\eta) N)$.
For any given $\eta > 0$ and $N$ large enough we then obtain
\begin{equation}
k = 2 S(\rho_E) - S(\rho_{2E}) = 2 g[(1-\eta)N] - g[2(1-\eta)N] \, .
\end{equation}
For the reverse reconciliation protocol we have (see Appendix)
$\chi = g(N) - g((1-\eta) N')$, with $N' = N/(1+\eta N)$,
$S(\rho_E) = g((1-\eta) N)$, and
$S(\rho_{2E}) = g((1-\eta) N') + g((1-\eta) N'')$, with $N'' = N ( 1 + 2 \eta N )/( 1 + \eta N)$.
For any given $\eta > 0$ and $N$ large enough we then obtain
\begin{align}
k & = 2 S(\rho_E) - S(\rho_{2E}) \\
& = 2 g[(1-\eta)N] - g[(1-\eta)N'] - g[(1-\eta)N''] \, .
\end{align}
\section{Conclusion}
Quantum cryptography promises unconditionally secure communication through insecure communication channels.
However, fundamental properties of quantum entanglement bound the ultimate
secret-key generation rates that can be achieved through a communication channel \cite{TGW}.
For the relevant case of a lossy communication line, as e.g. free-space of fiber optics
communication, the bound of \cite{TGW} implies that the secret-key generation rate must
decrease at least exponentially with increasing communication distance.
Here we have analyzed the rate-distance tradeoff under the realistic assumption that
one can store quantum information reliably only for a finite time. Clearly, any quantum
memory device can store quantum information only for a time of the order of its coherence
time. We have shown that for any given finite, yet arbitrarily long, storage time, the
quantum data locking effect can be applied to generate key at a constant rate over
arbitrarily long distances through an optical channel with linear loss.
Moreover, we have shown that this result holds also in the presence of moderate noise or experimental
imperfections modeled as a thermal background.
It remains an open problem to show that these high rates of key generation can be achieved in practice.
One major problem is to find a decoding measurement that can be experimentally
realized with current technologies and still allows us to achieve a constant key
rate over long communication distances.
If this question will find a positive answer, our results could pave the way
to a new family of QKD protocols that yield a constant key
rate that does {\it not} decay with increasing communication distance.
This would also imply that long distance quantum communication can be in principle
realized without employing quantum repeaters.
\acknowledgments
We are grateful to Bhaskar Roy, Mark M. Wilde, Saikat Guha, and Hari Krovi for valuable discussions and suggestions.
This research was supported by the DARPA Quiness Program through U.S. Army Research Office Grant No. W31P4Q-12-1-0019.
CL was supported by the SUTD--MIT Graduate Fellows Program.
|
\section{Introduction}
Symmetry-protected topological (SPT) phases are gapped phases of matter
which are not adiabatically deformable, under a given set of symmetry conditions,
to a topologically trivial phase.
While SPT phases do not have an intrinsic topological order
(i.e., do not support (deconfined) fractional excitations),
they are sharply (topologically) distinct from topologically trivial states,
such as an atomic insulator,
which respect the same set of symmetries.
In other words,
the distinction between SPT and trivial phases cannot be made within Landau's theory;
SPT phases are beyond the classification of phases of matter based on broken symmetries.
A flurry of recent theoretical works includes,
among others,
the breakdown (or collapse) of the non-interacting classifications of fermionic SPT phases upon inclusion of interactions,
non-trivial bosonic SPT phases and the possibility of symmetry-respecting surface topological order,
classification of interacting electronic topological insulators in three dimensions,
and
proposals for a possible complete classification of SPT phases.
(For a partial and incomplete list of recent works on SPT phases, see Refs.
\onlinecite{
KaneRev, QiRev, Classification2, Kitaev2009, Turner2013, SenthilReview,
Fidkowski2010, Fidkowski2011,
Ryu2012, Qi2012, yao2013interaction,
CWang2014, Fidkowski2013, Metlitski2014,
Vishwanath2013, Geraedts2014, Lu2012a, Cho2014, Cappelli2013,
Chen2013, Kapustin2014a, Kapustin2014b, Kapustin2014c}.)
One of the most efficient and powerful methods to study SPT phases
is to \textit{twist} or \textit{gauge} symmetries protecting SPT phases.
\cite{Levin2012, Ryu2012, Sule2013}
Quite generically, global symmetries in quantum field theories can be twisted,
i.e., can be used to define twisted boundary conditions.
It was proposed that the twisted theory can be used to diagnose the original SPT phases,
i.e., to judge whether or not the original theory is symmetry-protected, and distinct from topologically trivial phases.
More specifically,
once twisted, the edge theory of an SPT phase
suffers from various kinds of \textit{quantum anomalies},
such as a global anomaly under large $U(1)$ gauge transformations,
or a global gravitational anomaly.
\cite{Ryu2012, Sule2013, Hsieh2014}
On the other hand,
gauging (non-spatial) symmetries effectively deconfines a set of quasiparticles (anyons).
The fractional statistics of the braiding in the gauged theory can be used to diagnose
the original SPT phases.
\cite{Levin2012, Lu2012b}
Twisting non-spatial unitary symmetries of SPT phases is by now reasonably well-understood.
Toward further developments of methodologies to SPT phases,
it is necessary to extend the twisting or gauging procedure to spatial and/or antiunitary symmetries,
such as parity symmetry ($P$-symmetry) and time-reversal ($T$-symmetry).
\cite{Hsieh2014, Kapustin2014a,Kapustin2014b, Kapustin2014c, Chen_Gauging_Tsymmetry}
The purpose of this paper is to provide an efficient and intuitive method
to diagnose SPT phases protected by a spatial symmetry such as parity (reflection).
We will study (1+1)-dimensional [(1+1)d] non-chiral gapless theories with spatial symmetries,
which are the edge theories of the corresponding (2+1)d bulk SPT phases protected by
parity ($P$) symmetry or parity combined with a unitary non-spatial symmetry
such as $CP$-symmetry (parity symmetry combined with charge conjugation).
In typical situations,
in addition to $P$- or $CP$-symmetry,
there are other non-spatial symmetries
such as $U(1)$ symmetry or time-reversal ($T$) symmetry.
Our starting point is Ref. \onlinecite{Hsieh2014},
where twisting parity within edge theories of SPT phases (protected by parity) has been used to diagnose the edge theories.
This twisting procedure leads to theories that are effectively defined on an unoriented manifold, e.g., the Klein bottle,
and thus suggests an interesting link
between SPT phases and so-called “orientifold field theories” - type of theories discussed
in unoriented superstring theory.
\cite{Hori2003, Blumenhagen2009,Callan1987,Sagnotti1988, PolchinskiCai1988, Horava89,Angelantonj02,Brunner02, Brunner03, Yamaguchi}
We make one further step in this paper
by making use of the so-called \textit{cross-cap states}
in formulating orientifold field theories.
Cross-cap states are quantum states in field theories (conformal field theories) obtained by twisting parity symmetry
and encode, quantum mechanically, the unoriented topology;
the fact that the theory is put on an unoriented surface is entirely encoded in the cross-cap states.
They are akin to boundary states in boundary conformal field theories, which are obtained by exchanging
the role of space and time coordinates.
One promising aspect of cross-cap states is that they
are formulated and constructed fully in terms of many-body physics, and hence are expected to
capture the effects of interactions.
We will identify quantum anomalies in the edge theories of SPT phases
as the non-invariance of a cross-cap state
under the action of the other symmetry than parity.
Correspondingly, a bulk theory which supports an anomalous edge theory
is diagnosed as a non-trivial SPT phase.
Our procedure to diagnose an edge theory with parity symmetry can be summarized as follows:
\begin{itemize}
\item[(i):]
The edge theory of a given bulk theory is put on an unoriented spacetime, the Klein bottle.
In practice, this can be achieved by twisting boundary conditions (orientifold procedure).
\item[(ii):]
The edge theory on the Klein bottle is then quantized.
The cross-cap state corresponding to the twisting is constructed.
\item[(iii):]
In the quantized theory,
the effects of other symmetry, such as time-reversal, fermion number parity, etc.
are studied.
When these symmetries are anomalous, i.e., when the cross-cap state
is non-invariant under these symmetries,
the edge theory cannot be gapped while preserving the symmetries.
The corresponding bulk phase is symmetry-protected and
topologically distinct from a trivial phase.
\end{itemize}
In the previous work,
\cite{Hsieh2014}
the edge theories of SPT phases with
$P \rtimes U(1)_A$ symmetry and those with $CP \rtimes U(1)_V$ symmetry have been studied.
The partition function on the Klein bottle was shown to be anomalous (non-invariant)
upon threading a unit flux of the $U(1)$ symmetry.
In fact, the partition function acquires the anomalous $(-1)$-sign,
implying the ${\mathbb Z}_{2}$ classification of these SPT phases.
The proposed formulation in this paper in terms of cross-cap states
reproduces these results for SPT phase with $P\rtimes U(1)_A$ and $CP\rtimes U(1)_V$.
In addition,
this formulation in terms of cross-cap states offers
a few technical advantages.
First,
the new formulation allows
us to discuss SPT phases protected by a broader set of symmetries than
$P\rtimes U(1)_A$ or $CP\rtimes U(1)_V$.
That is, we do not need a continuous $U(1)$ symmetry and a large $U(1)$ gauge transformation.
Second, it is not necessary to compute the full (symmetry-twisted) partition function; all information
necessary to diagnose edge theories are encoded in cross-cap states.
The rest of the paper is organized as follows.
In Sec.\ \ref{Crosscap states and SPT phases},
some generalities on twisting/gauging parity symmetries and finding cross-cap states are presented.
We further demonstrate that non-invariance of the cross-cap states under symmetry operations
is related to the non-invariance of the partition function, i.e., the quantum anomalies.
We also draw some parallel between
twisting non-spatial symmetries (so-called orbifold procedure)
and twisting parity symmetries (orientifold).
In Sec.\ \ref{Bosonic SPT phases},
we apply our strategy to the bosonic SPT phases protected by $P$- or $CP$-symmetry together with other symmetries,
and find that the non-invariance of the cross-cap states reproduces
all the non-trivial SPT phases found in Ref. \onlinecite{Hsieh_CPT}.
In Sec.\ \ref{Real fermionic SPTs},
we discuss real fermionic SPT phases (topological superconductors)
protected by parity symmetry.
By identifying quantum anomalies (anomalous phases) in the cross-cap states,
the results consistent with
known microscopic analysis of gapping potentials are obtained.
\section{Crosscap states and SPT phases}
\label{Crosscap states and SPT phases}
In this section, some generalities of our approach to SPT phases are presented.
We start by briefly reviewing twisting and gauging non-spatial unitary symmetries of SPT phases,
in particular within (1+1)d edge theories of (2+1)d SPT phases.
When the edge theories are realized at the boundary of non-trivial SPT phases,
this twisting procedure (orbifolding) reveals a conflict between different symmetries at the quantum level,
even when they are mutually consistent at the classical level.
This is an example of quantum anomalies, signaling the impossibility of realizing the edge theory on its own
once symmetry conditions are imposed.
Hence, it is also indicative of the presence of non-trivial bulk states.
Twisting unitary spatial symmetry, such as parity and $CP$ symmetry, can be discussed
in a similar way, resulting in an unoriented spacetime manifold.
Cross-cap states, which are quantum states fully encoding the unoriented nature of the theory,
are introduced.
A potential conflict of other symmetries with parity/$CP$ symmetry can be studied in the
language of cross-cap states.
\subsection{Edge theories of SPT phases}
By definition, when
going from a SPT phase to a trivial phase
in a phase diagram by changing parameters in the system's Hamiltonian,
one inevitably encounters a quantum phase transition, if the symmetry conditions are strictly enforced.
This in turn implies that if an SPT phase is spatially proximate to a trivial phase, there should be a gapless
state localized at the boundary between the two phases; this critical state can be thought of as
a “phase transition” occurring locally in space, instead of the parameter space of the Hamiltonian.
As implied by this construction, the edge state of a non-trivial SPT phase
should never be removable (completely gapped) if the symmetries are strictly imposed.
(It should be noted, however, that there are other interesting possibilities for symmetric
surface states for (3+1)d SPT phases.
\cite{Vishwanath2013, xuludwig, xu3dspt,Geraedts2014})
Hence, this critical boundary state signals the topological distinction between the SPT and trivial phases,
and many properties of SPT phases can be extracted from their boundary physics.
For example, by inspecting under which symmetry conditions a given edge theory is stable/unstable, one can
predict under which symmetry conditions a given phase can be a SPT phase.
Although studying the boundary instead of the bulk reduces the dimensionality of the problem,
it is still not straightforward to judge if a given state is topological or not.
In principle, one could enumerate all possible symmetry-allowed perturbations within the edge theory,
which can potentially gap out the edge.
Without any guiding principle, however, such “brute force” approach is quite cumbersome,
and also, more fundamentally, does not provide any intuition on the physics of SPT phases.
Hence it is necessary to have an efficient and illuminating
guiding principle for diagnosing topological properties of phases of matter with symmetries.
\subsection{Symmetry twist and conflicting symmetries}
Our approach to diagnose non-trivial SPT phases is to identify quantum anomalies within edge theories.
We illustrate our strategy by considering a fermionic SPT phase protected by
$U_c(1) \times U_s (1)$ symmetry, where $U_{c/s}(1)$ refers to $U(1)$ symmetry associated with the conservation
of electromagnetic charge and $z$-component of spin $S_z$, respectively.
This phase is akin to (2+1)d time-reversal symmetric topological insulators (the quantum spin Hall effect),
although because of the imposed conservation of $S_z$,
states with $U_c(1)\times U_s(1)$ symmetry are classified by two integral topological invariants
(i.e., ``charge'' and ``spin'' Chern numbers).
We will focus on the case of vanishing charge Chern number,
and hence the allowed phases are classified by the integer-valued spin Chern number.
Consider the following Hamiltonian describing the edge of the $S_z$-conserving
quantum spin Hall phase with $U_c(1) \times U_s(1)$ symmetry defined on a circle of circumference $\ell$
\begin{align}
H = \int^{\ell}_0 dx\,
\left[
\psi^{\dagger}_{\uparrow} (- v i\partial_{x})\psi_{\uparrow} + \psi^{\dagger}_{\downarrow} (+ vi \partial_{x})\psi_{\downarrow}
\right],
\label{intro3}
\end{align}
where
$x\in [0, \ell]$ is the spatial coordinate of the edge,
$\psi_{\uparrow/\downarrow}$ represents the fermion creation operator for up/down spin,
and
$v$ is the Fermi velocity.
Any global symmetry in quantum field theories can be twisted
(i.e., can be used to twist boundary conditions).
We choose $U_c(1)$ symmetry to twist the boundary condition as
\begin{align}
\psi_{s}(x) = e^{2\pi i\alpha}
\psi_{s}(x+\ell),
\quad
s \in \{ \uparrow, \downarrow \}.
\label{charge bc}
\end{align}
Let the ground state with the twisted boundary condition be $ |\mathrm{GS} \rangle_{\alpha}$,
which satisfies
\begin{align}
\left[ \psi_{s}(x) - e^{2\pi i\alpha}\psi_{s}(x+\ell) \right] |\mathrm{GS} \rangle_{\alpha} = 0,
\quad s \in \{ \uparrow, \downarrow \}.
\end{align}
Observe that the boundary condition (\ref{charge bc}) is invariant under global $U(1)_c$ transformations
\begin{align}
0&=
\psi_{s}(x) - e^{2\pi i\alpha} \psi_{s}(x+\ell)
\nonumber \\
&=
e^{i \phi Q} \left[ \psi_{s}(x) - e^{2\pi i \alpha} \psi_{s}(x+\ell) \right]e^{-i\phi Q},
\end{align}
where $Q$ is the total charge operator associated to $U(1)_c$ symmetry,
and $\phi$ is an arbitrary real parameter.
Similarly,
the boundary condition (\ref{charge bc}) is invariant under the global $U(1)_s$
transformation generated by $e^{i \phi S_z}$
where $S_z$ is the total ``charge'' associated to $U(1)_s$ symmetry.
A crucial observation now is that,
while
the boundary condition (\ref{charge bc}) is invariant under the global $U(1)_s$,
the corresponding ground state may carry an ``anomalous'' $S_z$ quantum number,
\begin{align}
e^{i Q \phi} |\mathrm{GS} \rangle_{\alpha} & = e^{i \times 0 \times \phi } |\mathrm{GS} \rangle_{\alpha} = |\mathrm{GS} \rangle_{\alpha} ,
\nonumber\\
e^{i S_z \phi }|\mathrm{GS} \rangle_{\alpha} & = e^{i\times 2\alpha \times \phi} |\mathrm{GS} \rangle_{\alpha} =
e^{2i\alpha \phi} |\mathrm{GS} \rangle_{\alpha}.
\end{align}
(The quantum numbers of the ground state can be computed explicitly, by using, for example, bosonization and identifying
an operator corresponding to the ground state $|\mathrm{GS}\rangle_{\alpha}$.
More precisely, the operator is given by $\sim \exp(i \alpha (\varphi_L + \varphi_R))$,
where the left- and right-moving electrons are identified as $\psi_{L/R} \sim \exp(\pm i\varphi_{L/R})$.
For details, see Ref.\ \onlinecite{Cho2014}.)
The anomalous phase is nothing but chiral anomaly;
in the presence of a $U(1)_c$ magnetic flux twisting the boundary condition,
the $S_z$ quantum number, which is conserved at the classical level,
is not conserved at the quantum level.
Hence, only way to reconcile the $U(1)_c\times U(1)_s$ symmetry and quantum mechanics is
to realize this (1+1)d theory as a boundary theory of a higher-dimensional system,
a (2+1)d bulk SPT phase respecting the $U(1)_c \times U(1)_s$ symmetry.
Let us now be more general.
Consider an edge theory, which is written in terms of a fundamental quantum field $\Phi(x)$
and is defined on a spatial circle $x \sim x+\ell$.
Suppose there is a set of non-spatial unitary symmetry operators, ${G}$,
which leave the edge theory invariant.
We consider a twisting boundary condition by a group element ${g}_1 \in {G}$,
\begin{align}
\Phi(x+\ell) = \mathscr{G}^{\ }_1 \Phi(x) \mathscr{G}^{-1}_1 = U_{g_1}\cdot \Phi (x),
\label{twisting g1}
\end{align}
where $\mathscr{G}_1$ is the operator which implements a symmetry operation $g_1$ in
the Hilbert space of the edge theory, and
$U_{g_1}$ is a unitary matrix acting on the internal index of the field $\Phi(x)$,
i.e., a unitary matrix representation of the symmetry.
(The field $\Phi(x)$ can carry a set of indices representing internal degrees of freedom, which are suppressed in
the equation above and in the following.)
With twisting,
states in the Hilbert space,
the ground state $|\mathrm{GS}\rangle_{g_1}$ in particular,
obey
\begin{align}
\left[
\Phi(x + \ell) - U_{g_1} \Phi(x)
\right]
|\mathrm{GS} \rangle_{g_1} =0.
\end{align}
As a next step, we consider the action of another symmetry $g_2 \in G$.
($g_2$ can be equal to $g_1$, which is the situation relevant to the quantum Hall effect,
but in our examples below, $g_2\neq g_1$.)
At the classical level, the $g_1$-twisted boundary condition (\ref{twisting g1})
may be invariant under $g_2 \in {G}$,
\begin{align}
0&=
\Phi(x + \ell) - U_{g_1} \Phi(x)
\nonumber \\
&=
\mathscr{G}^{\ }_2 \left[ \Phi(x + \ell) - U_{{g}_1} \Phi(x) \right] \mathscr{G}^{-1}_2.
\end{align}
When this is the case,
one may expect the twisted theory after quantization is invariant under $g_2$ as well.
In particular, one expects the partition function and/or the ground state of the twisted edge theory
is invariant under $g_2$.
When this expectation is betrayed, there is a quantum anomaly.
Before leaving this subsection,
we comment on a connection between
the twisting procedure and orbifolding/gauging.
Gauging and orbifolding
have a similar (identical) effect in that we focus on a gauge singlet ($G$-invariant) sector of the theory
[although the gauging means in general imposing the singlet condition locally (e.g., at each site of a lattice), while
the projection in orbifolding is enforced only globally].
Let us again consider an edge theory with symmetry group $G$.
By state-operator correspondence in quantum field theories,
corresponding to the state $|\mathrm{GS}\rangle_g$ in the twisted theory,
there is an operator, the twist operator $\sigma_{g}$,
which, when acting on the ground state of the untwisted theory,
creates $|\mathrm{GS}\rangle_g$
\cite{CFTbook, PolchinskiStringbook,Ginsparg91}:
\begin{align}
|\mathrm{GS}\rangle_g= \sigma_{g} (0)|0\rangle.
\end{align}
(The location of the insertion of the operator $\sigma_{g}$ is taken to be the origin in the radial quantization.)
The twist operator, when inserted in correlation functions,
implements the symmetry twist by $g$,
and satisfy the following algebraic relation with $\Phi(x)$
on the complex plane:
\begin{align}
\Phi(z,\bar{z}) \sigma_g(w, \bar{w})
=
\sigma_g(w,\bar{w}) U_g \cdot \Phi(z,\bar{z}).
\end{align}
Starting from the original, untwisted, theory,
one can now consider including the ground states with twisted boundary conditions
to define an extended theory.
In the extended theory,
the ground states with twisted boundary conditions
(the ground states in the ``twisted sectors''),
and hence the corresponding twist operators,
are considered as an excitation.
This procedure to generate a new theory from the untwisted theory is called orbifold.
Now, by further invoking bulk-boundary correspondence,
there is a corresponding bulk excitation (anyon).
Bulk statistical properties of the gauged theory can be read off
from the operator product expansions and fusion rules obeyed by
the twist operator(s).
Now a given symmetry group $G$ can be implemented in various different ways
in different SPT (and trivial) phases protected by $G$
leading to different choices of $U_g$, and to different twist operators.
By studying statistical (braiding) properties of the twist operator(s),
one can distinguish different ungauged (original) theories.
\cite{Levin2012}
\subsection{Spatial symmetry}
When considering SPT phases protected by spatial symmetries,
one can consider twisting the spatial symmetries.
Let us consider a parity symmetry:
\begin{align}
\mathscr{P}\Phi(x)\mathscr{P}^{-1}
= U_{{P}} \Phi(\ell-x),
\end{align}
where the space is defined on a circle $x \sim x+ \ell$.
Twisting by parity symmetry can be introduced in the following way.
\cite{Brunner02, Brunner03, Yamaguchi,Horava89, Hori2003, Blumenhagen2009}
\paragraph{Loop channel}
We consider Euclidean spacetime $[0, \ell] \times [0, \beta]$ parameterized by $(x_1,x_2)$,
where $x_1$ is the spatial coordinate and $x_2$ is the imaginary time coordinate.
In the path integral picture, the fields obey the following twisted boundary conditions
\begin{align}
\Phi(x_1, x_2+ \beta)
&=
\mathscr{P} \Phi(x_1, x_2)\mathscr{P}^{-1}
=
U_{P} \Phi(\ell-x_1, x_2),
\nonumber \\
\Phi(\ell, x_2) &= \mathscr{G} \Phi(0,x_2) \mathscr{G}^{-1},
\label{BC_xCap}
\end{align}
in which $\mathscr{G}$ may be implementing a non-spatial unitary symmetry $g \in G$,
e.g., the fermion number parity.
\cite{Hsieh2014}
In the operator language,
twisting by parity symmetry can be introduced as a projection operation.
This amounts to inserting the parity operator into the partition function,
\begin{align}
Z^{K}
&=
\mathrm{Tr}_{g}
\left[
\mathscr{P} e^{-\beta H^{\ }_{\mathrm{loop}}(\ell) }
\right],
\end{align}
where $H_{\mathrm{loop}}\equiv H$ is the Hamiltonian that generates time-translation
in the $x_2$-direction.
The trace is taken in the Hilbert space defined for the quantum field obeying the
boundary condition $\Phi(x_1+\ell) = \mathscr{G} \Phi(x_1) \mathscr{G}^{-1}$,
as indicated by the subscript $g$.
(For later use, we also introduce another parity operator,
$\mathscr{P}'$, such that
$\mathscr{P}^{\prime}\cdot \mathscr{P}^{-1} =\mathscr{G}$.)
This representation of the partition function with the choice of $x_2$
as a direction of time-evolution will be called the ''loop channel'' picture.
\paragraph{Tree channel}
As we have seen in the case of non-spatial symmetries,
it is useful to discuss the twisted partition function
in terms of the twist operator and the corresponding state.
One may want to develop a similar alternative picture for the case of twisting by parity.
To this end, one needs to find a convenient and proper
``time-slice'' that allows us to define a quantum state
($|\mathrm{GS}\rangle_g$ in the notation of the previous section for the non-spatial symmetry),
which implements the twisting condition.
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth]{Xcap}
\caption{
Rearrangement of spacetime.
(a) The original loop channel (Euclidean) spacetime.
Here the points $(A,B)$ and $(A',B')$ in the spacetime are identified due to the boundary conditions Eq.\ \eqref{BC_xCap}.
The line segments with the same symbols are also identified by the boundary conditions Eq.\ \eqref{BC_xCap}.
(b) We shift the box of the spacetime formed by $A'B'D'C'$ along $x_2$ by $\beta$.
(c) After shifting the box, we flip the orientation of the box with respect to $x_1 = 3\ell/4$.
(d) We slide the box along $x_1$ by $\ell/2$.
Then we obtain the crosscap geometry.
Notice that any $x_2 \in [0, \beta]$ is identified with its antipodal point at the slice of ``time" at $x_1 = 0$
and $x_1 = \ell/2$.
By renaming the variables $x_1 \to \sigma_1$ and $x_2 \to \sigma_2$, we obtain the spacetime of the tree channel. Here the $\sigma_1$ direction is taken to be a direction of
(fictitious) time-evolution. At the boundary of spacetime located at $\sigma_1=0$ and $\sigma_1=\ell/2$,
there are cross-caps.
\label{fig:xCap}
}
\end{center}
\end{figure}
It is convenient to recall that
any 2d compact unoriented surface without boundary can be generated from a sphere $S^2$
by adding handles and cross-caps, which can be thought of as a real projective plane $\mathbb{RP}_2$.
In particular,
the Klein bottle can be generated by first cutting out two discs from $S^2$,
and then identifying antipodal points of the resulting holes.
In fact, one can rearrange the path integral on the Klein bottle
such that the fields are now defined on $[0, \ell/2] \times [0, 2\beta]$.
The precise steps for rearranging the spacetime is described in Fig.\ref{fig:xCap}.
We use $(\sigma_1,\sigma_2)$ to represent this rearranged spacetime.
The fields now obey the cross-cap boundary conditions
\begin{align}
\Phi(0, \sigma_2+\beta)
&=
U_{P'} \Phi(0, \sigma_2),
\nonumber \\
\Phi(\ell/2, \sigma_2+\beta)
&=
U_{{P}} \Phi(\ell/2, \sigma_2),
\nonumber \\
\Phi(\sigma_1, \sigma_2+2\beta)
&=
\mathscr{G}' \Phi(\sigma_1, \sigma_2)\mathscr{G}^{\prime -1},
\end{align}
where
\begin{align}
\mathscr{P}^2 = \mathscr{P}^{\prime 2} =\mathscr{G}^{\prime}.
\label{cross cap formula 1}
\end{align}
By further rotating $(\sigma_1,\sigma_2)$ coordinates by $90^{\circ}$,
we can take $\sigma_1$ as a time-coordinate.
In this coordinate system, the time-evolution is generated by
a (fictitious) Hamiltonian (denoted by $H_{\mathrm{tree}}$ in the following).
The fact that the system is defined on an unoriented surface
is now encoded in boundary conditions at $\sigma_1=0$ and $\sigma_1=\ell/2$.
We then introduce cross-cap states which obey these boundary conditions as:
\begin{align}
&
\left[
\Phi(0, \sigma_2+\beta)
-
U_{P'} \Phi(0, \sigma_2)
\right]|C_{P} \rangle
=0,
\nonumber \\
&
\left[
\Phi(\ell/2, \sigma_2+\beta)
-
U_{{P}} \Phi(\ell/2, \sigma_2)
\right]|C_{P^{\prime}}\rangle=0.
\end{align}
The partition function can be written in terms of the cross-cap states as
\begin{align}
Z^{K}=
\langle {C}_{P^{\prime}} | e^{-\frac{\ell}{2} H^{\ }_{\mathrm{tree}} (2\beta)}
|{C}_{P} \rangle,
\end{align}
where $H^{\ }_{\mathrm{tree}}$ is the Hamiltonian that generates time-translation
in the tree channel picture.
The loop channel-tree channel duality (also known as open-closed duality)
asserts that the partition functions computed in the loop and tree channels agree.
\paragraph{Interplay with non-spatial symmetries}
By constructing the cross-cap states,
we have ``gauged'' parity symmetry.
As in the case of the SPT phases with non-spatial symmetries,
we now consider the effects of another non-spatial symmetry,
represented by $ g\in {G}$, on the cross-cap states.
We act with $g$ on the cross-cap boundary conditions,
\begin{align}
&\quad
\left[
\Phi(0, \sigma_2 + \beta)
-
U_{{P}} \Phi(0, \sigma_2)
\right]
| {C}_{{P}}\rangle
=0
\nonumber \\
&\Rightarrow
\mathscr{G}
\left[
\Phi(0, \sigma_2 +\beta)
-
U_{{P}} \Phi(0, \sigma_2)
\right]
\mathscr{G}^{-1}
\mathscr{G}
| {C}_{{P}}\rangle
=0
\nonumber \\
&\Rightarrow
\left[
U_{{g}} \Phi(0, \sigma_2 + \beta)
-
U_{{P}} U_{{g}} \Phi(0, \sigma_2)
\right]
\mathscr{G}
| {C}_{{P}}\rangle
=0
\nonumber \\
&\Rightarrow
\left[
\Phi(0, \sigma_2 + \beta)
-
U^{-1}_{{g}}
U^{\ }_{{P}}
U^{\ }_{{g}} \Phi(0, \sigma_2)
\right]
\mathscr{G}| {C}_{{P}}\rangle
=0.
\end{align}
Thus we deduce the following relation:
\begin{align}
\mathscr{G} |{C}_{{P}}\rangle
=
|{C}_{{g}\cdot {P} \cdot {g}^{-1}} \rangle.
\end{align}
If the cross-cap condition is invariant under
$g \in {G}$,
$U_{P} =U^{-1}_{{g}}U^{\ }_{{P}}U^{\ }_{{g}}$,
then we may expect that so is the cross-cap state,
$|{C}_{{P}} \rangle = \mathscr{G} |{C}_{{P}}\rangle$,
classically.
However this expected invariance may be broken down
quantum mechanically.
This then signals that the theory is anomalous and should describe
the edge of a SPT phase defined in one higher dimension.
\section{Bosonic SPT phases}
\label{Bosonic SPT phases}
In this section,
we apply our strategy above to edge theories of bosonic SPT phases
consisting of a single component non-chiral boson
(i.e., a two-component chiral boson with $2\times 2$ K-matrix).
We consider situations where
parity ($P$) or a combination of parity and charge conjugation ($CP$) is a
part of the full symmetry group, which can potentially protect the edge theory
from gap-opening.
In Ref.\ \onlinecite{Hsieh2014},
SPT phases protected by $P$ or $CP$ symmetry together with a continuous $U(1)$ symmetry
were studied by using a generalization of Laughlin's argument.
It was found that non-trivial $\mathbb{Z}_2$ bosonic SPT phases can exist
in the presence of the following combinations of
symmetries:
\begin{itemize}
\item
$P \rtimes U(1)_A$ (``spin $U(1)$''),
\item
$CP \rtimes U(1)_V$ (``charge $U(1)$''),
\end{itemize}
where the subscript $A/V$ represents the ``axial/vectorial'' nature of the $U(1)$ symmetry.
In Ref. \onlinecite{Hsieh_CPT},
following the spirit of Refs.\ \onlinecite{Lu2012a, Lu2013},
a microscopic stability analysis for the bosonic edge theory
is carried out by enumerating possible gapping potentials
in the presence of various discrete symmetries.
It was found that
the bosonic edge theory can be ingappable (protected)
in the presence of the following symmetries:
\begin{itemize}
\item
$P\times C, P\times T$
\item
$P \times TC$,
$CP \times T$,
$T \times C$.
\end{itemize}
SPTs protected by these symmetries are all classified by $\mathbb{Z}_2$.
In addition, it was also found that there are SPT phases protected by
\begin{itemize}
\item $T \times C \times P$
\end{itemize}
The classification of these SPT phases is ${\mathbb Z}^{4}_2$.
In the following,
we will study these SPT phases by using cross-cap states and by identifying quantum anomalies.
As we will show, this analysis reproduces exactly the same classification as
in Ref.\ \onlinecite{Hsieh_CPT},
and hence it gives us a perspective complimentary to the generalized Laughlin's argument on
the Klein bottle in the ``loop channel'' picture, and to microscopic stability analysis.
\subsection{Free compactified boson}
We start from the free boson theory on a spatial ring of circumference $\ell$
defined by the partition function $Z=\int\mathcal{D}[\phi] \exp (iS)$ with the action
\begin{align}
S=\frac{1}{4\pi\alpha'}\int dt \int_{0}^{\ell} dx
\left[
\frac{1}{v}( \partial_{t}\phi )^{2}
-v( \partial_{x}\phi )^{2}
\right],
\end{align}
where
the spacetime coordinate of the edge theory is denoted by $(t,x)$,
$v$ is the velocity,
$\alpha'$ is the coupling constant,
and the $\phi$-field is compactified as
\begin{align}
\phi \sim \phi + 2\pi R,
\end{align}
with the compactification radius $R$.
The canonical commutation relation is
\begin{align}
\left[
\phi(x,t), \partial_t {\phi}(x^{\prime},t)
\right]
=
i
2\pi \alpha' v
\sum_{n\in \mathbb{Z}}\delta(x-x^{\prime}-n\ell).
\end{align}
We use the chiral decomposition of the boson field, and introduce the dual field $\theta$ as
\begin{align}
\phi=\varphi_L+\varphi_R,
\quad
\theta =\varphi_L-\varphi_R.
\end{align}
The mode expansion of the chiral boson fields is given by ($x^{\pm}=vt\pm x$)
\begin{align}
\varphi_{L}(x^{+})
&=
x_L
+
\pi \alpha' p_L
\frac{ x^{+}}{\ell}
+
i
\sqrt{ \frac{\alpha'}{2} }
\sum^{n\neq 0}_{n\in \mathbb{Z}}
\frac{\alpha_{n}}{{n}}e^{-\frac{2\pi i nx^{+}}{\ell}}
,
\nonumber \\
\varphi_{R}(x^{-})
&=
x_R
+
\pi \alpha' p_R
\frac{x^{-}}{\ell}
+
i \sqrt{ \frac{\alpha'}{2} }
\sum^{n\neq 0}_{n\in \mathbb{Z}}
\frac{\tilde{\alpha}_{n}}{{n}}e^{-\frac{2\pi i nx^{-}}{\ell}} ,
\end{align}
where
$
\left[\alpha_m, \alpha_{-n}\right] =
\left[\tilde{\alpha}_m, \tilde{\alpha}_{-n}\right] =
m\delta_{mn}$ and $ [x_L, p_L]= [x_R, p_R]= {i}$.
The compactification condition on the boson fields
implies the allowed momentum eiganvalues are given by
\begin{align}
&p=\frac{1}{2}\left(p_L + p_R\right) = \frac{k}{R},
\quad
\tilde{p} = \frac{1}{2} \left( p_L - p_R \right) = \frac{R}{\alpha'}w,
\nonumber \\
&p_L = \frac{k}{R}+\frac{R}{\alpha'}w,
\quad
p_R = \frac{k}{R}-\frac{R}{\alpha'}w,
\end{align}
where $k$ and $w$ are an integer.
In terms of these momentum eigenvalues,
the compactification conditions on
the boson fields are
\begin{align}
\varphi_L(x+\ell) - \varphi_L(x)&=+\pi \alpha' p_L,
\nonumber\\
\varphi_R(x+\ell) - \varphi_R(x)&= -\pi \alpha' p_R,
\nonumber\\
\phi (x+\ell) - \phi(x)& = \pi \alpha'(p_L-p_R)=2\pi R w,
\nonumber \\
\theta(x+\ell) - \theta(x) &= \pi \alpha'(p_L+p_R) = 2\pi \frac{\alpha'}{R}k.
\end{align}
The Hilbert space is constructed as a tensor product of
the bosonic oscillator Fock spaces,
each of which generated by
pairs of creation and annihilation operators
$\{ \alpha_{m}, \alpha_{-m}\}_{m>0}$
and
$\{\tilde{\alpha}_{m}, \tilde{\alpha}_{-m}\}_{m>0}$,
and
the zero mode sector associated to ${x}_{L,R}$ and ${p}_{L,R}$.
We will denote states in the zero mode sector by specifying
their momentum eigenvalues as
\begin{align}
|p, \tilde{p}\rangle = | k/R, R w/\alpha'\rangle,
\quad
k, w\in \mathbb{Z},
\end{align}
or more simply as $|k,w\rangle$.
Alternatively,
the Fourier transformation of the momentum eigenkets
defines the ``position'' eigenkets, which we denote by
\begin{align}
|\phi_0, \theta_0\rangle
\quad
0 <\phi_0 \le 2\pi R,
\quad
0< \theta_0 \le 2 \pi \alpha'/R.
\end{align}
The two basis are related by
\begin{align}
| p, \tilde{p}\rangle
=
\int^{2\pi R}_0 d\phi_0
\int^{2\pi \alpha'/R}_0 d\theta_0
e^{ -i p \phi_0 -i \tilde{p}\theta_0}
|\phi_0,\theta_0\rangle.
\end{align}
\subsection{Symmetries}
Various symmetries of the single-component compactified boson theory
are listed below.
\paragraph{$U(1)\times U(1)$ symmetry}
In the free boson theory, when there is no perturbation,
there are two conserved $U(1)$ charges, one for each left- and right-moving sector,
defined by
\begin{align}
N_{L,R}&= \int^{\ell}_0 dx\, \partial_x \varphi_{L,R} =
\alpha' \pi p_{L,R},
\end{align}
They satisfy
\begin{align}
\left[
\varphi_L, N_L
\right]
=
\left[
\varphi_R, N_R
\right]
=
\alpha' \pi {i}.
\end{align}
Correspondingly to these conserved quantities,
the free boson theory is invariant under the following
$U(1)\times U(1)$ symmetry
\begin{align}
\mathscr{U}_{\delta\phi, \delta \theta}&:
\phi \to \phi + \delta \phi,
\quad
\theta \to
\theta +
\delta \theta,
\nonumber \\
&:\varphi_L
\to
\varphi_L
+
\delta \varphi_L,
\quad
\varphi_R
\to
\varphi_R
+
\delta \varphi_R.
\end{align}
In terms of the conserved charges,
the generators of the $U(1)\times U(1)$ transformations are given by
\begin{align}
\mathscr{U}^L_{\delta \varphi_L}
& =
e^{i \delta \varphi_L
N_{L}/(\alpha' \pi)
}
=
e^{
i \delta \varphi_L
p_{L}
},
\nonumber \\
\mathscr{U}^R _{\delta \varphi_R}
& =
e^{
i \delta \varphi_R
N_{R}/(\alpha' \pi)
}
=
e^{
i\delta \varphi_R p_{R}
},
\nonumber \\
\mathscr{U}_{\delta\phi,\delta\theta} &=
\mathscr{U}^L_{\delta \varphi_L}
\mathscr{U}^R_{\delta \varphi_R}
=
e^{
i (\delta \phi p + \delta \theta \tilde{p})}.
\end{align}
Note also that
$\mathscr{U}_{\delta \phi, \delta \theta}$
acts on the momentum eigenkets as
\begin{align}
\mathscr{U}_{\delta \phi, \delta \theta} |p,\tilde{p}\rangle =
e^{i (p\delta \phi + \tilde{p} \delta\theta)} |p, {\tilde p} \rangle.
\end{align}
\paragraph{$C$-symmetry}
Particle-hole symmetry or charge conjugation ($C$-symmetry) is unitary and acts on
the bosonic fields as
\begin{align}
\mathscr{C}&:\phi \rightarrow - \phi + n_{c}\pi R,
\quad
\theta \rightarrow -\theta + \frac{m_{c}\pi \alpha'}{R} \nonumber\\
&: (x_1, x_2) \rightarrow (x_1, x_2),
\label{csymmetry:phase}
\end{align}
where $(n_{c}, m_{c}) \in \{0,1\}$.
From these transformation laws of the boson fields,
we read off the action of $C$-symmetry on the position basis as
\begin{align}
\mathscr{C}| \phi_0,\theta_0\rangle
=
e^{i\delta}
\left|-\phi_{0} + n_{c}\pi R, -\theta_{0} + m_{c}\pi \alpha'/R
\right\rangle,
\label{phase ambiguity, C-symmetry}
\end{align}
where $e^{i \delta}$ is an unknown phase factor.
In order to have the relation
$\mathscr{C}|p, \tilde{p}\rangle \propto |-p, -\tilde{p}\rangle$,
expected from the commutation relation between $\mathscr{C}$ and $p,\tilde{p}$,
the phase $\delta$ has to be a constant (independent of $\phi_{0}$ and $\theta_0$).
The action of $C$-symmetry on the momentum eigenstates is given by
\begin{align}
\mathscr{C}
|p, {\tilde p} \rangle
&=
e^{i \delta}
e^{-ipn_{c}\pi R - i{\tilde p}\frac{m_{c}\pi \alpha'}{R}}
|-p, -{\tilde p} \rangle
\nonumber \\
&=
e^{i \delta}
e^{-i\pi kn_{c} - i \pi w m_{c}}
|-p, -{\tilde p} \rangle,
\label{csymmetry:momentum}
\end{align}
where $p=k/R$ and $\tilde{p}=w R/\alpha'$.
Since $\delta$ is constant, the phase ambiguity is fixed once we specify the action of
$\mathscr{C}$ on a {\it reference state}, e.g., $|p,\tilde{p}\rangle = |0,0\rangle$.
In our analysis presented below, the reference state and its charge conjugation parity $e^{i\delta}$
plays an important role.
\paragraph{$T$-symmetry}
Antiunitary time-reversal operator $\mathscr{T}$ acts on the boson fields as
\begin{align}
\mathscr{T}&: \phi \rightarrow \phi + n_{T}\pi R,
\quad
\theta \rightarrow -\theta + \frac{m_{T} \pi \alpha'}{R} \nonumber\\
&: (x_1,x_2) \rightarrow (x_1, -x_2),
\end{align}
where $n_{T}, m_{T} \in \{0 ,1 \}$.
\paragraph{$P$-symmetry}
Parity $\mathscr{P}$ is defined by
\begin{align}
\mathscr{P}
&: \phi \rightarrow \phi + n_{p}\pi R,
\quad
\theta \rightarrow -\theta + \frac{m_{p}\pi \alpha'}{R}\nonumber\\
&: (x_{1},x_{2}) \rightarrow (\ell -x_{1},x_{2}),
\label{psymmetry}
\end{align}
where $n_{p}, m_{p} \in \{0 ,1 \}$.
\paragraph{$CP$-symmetry}
The above symmetries can be combined.
For example, $CP$-symmetry is a non-local unitary symmetry,
and defined by
\begin{align}
\mathscr{CP}
&: \phi \rightarrow
-\phi + n_{cp}\pi R,
\quad
\theta \rightarrow \theta + \frac{m_{cp}\pi \alpha'}{R}\nonumber\\
&: (x_{1},x_{2}) \rightarrow (\ell -x_{1},x_{2}),
\label{cpsymmetry}
\end{align}
where $n_{cp}, m_{cp} \in \{0 ,1 \}$.
\subsection{Crosscap States}
We now move on to the tree channel picture and construct cross-cap states
by twisting $P$- and $CP$- symmetries.
After rearranging spacetime and exchanging the role of space and time,
the spacetime is,
$\mbox{space}\times \mbox{time} = 2\beta \times \ell/2$.
We parameterize this Euclidean spacetime by $(\sigma_1,\sigma_2)$.
The original spacetime of $[0, \ell] \times [0, \beta]$ is parameterized by $(x_1, x_2)$
(for the mapping between the two spacetime see Fig.\ \ref{fig:xCap}).
\subsubsection{$P$-symmetry}
We first consider the cross-cap state obtained by twisting parity symmetry,
defined in Eq.\ (\ref{psymmetry}).
The corresponding cross-cap states are
defined by
\begin{align}
&\bigg[ \phi(\sigma_{2}) - \phi(\sigma_{2} + \beta) - n_{p}\pi R \bigg]|C_{p} (n_{p}, m_{p}) \rangle =0,
\nonumber\\
&\bigg[\theta(\sigma_{2}) +\theta (\sigma_{2} + \beta) - \frac{m_{p}\pi \alpha'}{R}\bigg]|C_{p} (n_{p}, m_{p}) \rangle =0.
\label{xcapcondition:p}
\end{align}
By mode-expansion,
the cross-cap condition (\ref{xcapcondition:p}) translates into the corresponding condition for each mode.
To find an explicit form of the cross-cap states, we first focus on the zero mode sector of the boson fields:
\begin{align}
\phi(\sigma_{2}) &= x_{L} + x_{R} + \frac{\pi \alpha' {\tilde p} \sigma_{2}}{\beta}+\cdots,
\nonumber\\
\theta(\sigma_{2}) &= x_{L} - x_{R} + \frac{\pi \alpha' p \sigma_{2}}{\beta} +\cdots.
\end{align}
Within the zero mode sector, we solve the cross-cap conditions \eqref{xcapcondition:p}.
The first condition \eqref{xcapcondition:p} can be reduced to
\begin{align}
&\quad \pi \alpha' {\tilde p} -n_{p}\pi R = 0 \text{ mod } 2\pi R \nonumber\\
&\Rightarrow {\tilde p} = \frac{R}{\alpha'} (2N + n_{p}),\quad N \in {\mathbb Z}.
\label{xcapsolution1:p}
\end{align}
Similarly, the second condition can be solved as:
\begin{align}
&2(x_{L}- x_{R}) + \frac{2\pi \alpha' p \sigma_2}{\beta} + \pi \alpha' p = \frac{m_{p}\pi \alpha'}{R} \text{ mod } \frac{2\pi \alpha'}{R} \nonumber\\
& \Rightarrow p=0,
\quad (x_{L} - x_{R}) = (2{\tilde N} + m_{p}) \frac{\pi \alpha'}{2R}, ~ {\tilde N} \in {\mathbb Z}.
\label{xcapsolution2:p}
\end{align}
Solving the conditions \eqref{xcapsolution1:p} and \eqref{xcapsolution2:p},
we find the cross-cap state
$|C_{p} (n_{p}, m_{p}) \rangle$
in terms of the momentum eigenket $\{ |p, {\tilde p} \rangle \}$ as
\begin{equation}
|C_{p} (n_{p}, m_{p}) \rangle = \sum_{N \in {\mathbb Z}} (-1)^{m_{p}N} |0, 2N+n_{p} \rangle.
\label{xcapstate:p}
\end{equation}
The full cross-cap state is obtained by including the parts related to the oscillatory modes:
\begin{align}
\sqrt{R\sqrt{2} }
\exp \left[
-\sum_{n=1}^{\infty} \frac{(-1)^n}{n} \alpha_{-n}\tilde{\alpha}_{-n}
\right]
|C_p (n_p, m_p)\rangle.
\end{align}
For our purpose of diagnosing SPT phases, however, it turns out that it is enough
to focus on the zero-mode sector.
\subsubsection{$CP$-symmetry}
Next we consider $CP$-symmetry.
The corresponding cross-cap conditions are given by
\begin{align}
&\bigg[ \phi(\sigma_{2}) + \phi(\sigma_{2} + \beta) - n_{cp}\pi R \bigg]|C_{cp} (n_{cp}, m_{cp}) \rangle =0,
\nonumber\\
&\bigg[\theta(\sigma_{2}) -\theta (\sigma_{2} + \beta) - \frac{m_{cp}\pi \alpha'}{R}\bigg]|C_{cp} (n_{cp}, m_{cp}) \rangle =0.
\nonumber\\
\label{xcapcondition:cp}
\end{align}
By solving these conditions within the zero mode sector,
we obtain, as the cross-cap state,
\begin{equation}
|C_{cp} (n_{cp}, m_{cp}) \rangle = \sum_{N \in {\mathbb Z}} (-1)^{n_{cp}N} |2N+m_{cp},0 \rangle.
\label{xcapstate:cp}
\end{equation}
\subsection{Diagnosis of SPT phases}
\label{diagnosis}
\subsubsection{$P \rtimes U(1)_A$ and $CP \rtimes U(1)_V$ symmetries}
We start with the case of the symmetry group $P\rtimes U(1)_A$,
which has been studied in Ref.\ \onlinecite{Hsieh2014}
by using a generalization of Laughln's gauge argument.
In the generalized Laughlin's argument,
the edge theory is put on an unoriented spacetime, such as the Klein bottle,
and then the invariance under flux threading
(a large gauge transformation of $U(1)_A$ symmetry)
of the partition function of the edge theory is investigated.
More specifically,
the analysis in Ref.\ \onlinecite{Hsieh2014} is performed
in the loop-channel channel picture.
In the following, we will reproduce this result in terms of the tree-channel calculations,
i.e., by using the cross-cap state.
The relevant cross-cap state is $|C_p (n_p,0)\rangle$ presented in Eq.\ (\ref{xcapstate:p}).
We set $m_p =0$ since $n_{p} \in \{0,1\}, m_p =0$ is enough to classify the $P\rtimes U(1)_A$-symmetric SPT phases
according to Ref.\ \onlinecite{Hsieh_CPT}.
However, it is straightforward to generalize the analysis below to more general sets of $\{n_p, m_p \}$. The $U(1)_{A}$ symmetry is generated by
$\mathscr{U}^A_{\delta \theta} = \exp(i{\tilde p} \delta \theta)$:
\begin{align}
{\mathscr U}^{A}_{\delta \theta}&:
\theta \rightarrow \theta+\delta \theta,
\quad
\phi \rightarrow \phi.
\end{align}
Let us act with $\mathscr{U}^A_{\delta\theta}$ on the cross-cap condition (\ref{xcapcondition:p}).
The cross-cap condition for $\phi$ is trivially invariant under $\mathscr{U}^A_{\delta\theta}$.
For the condition written in terms of the dual field $\theta$ (with $m_{p}=0$),
\begin{align}
&
\mathscr{U}_{\delta\theta}^{A}
\left\{
\theta(\sigma_{2}) +\theta (\sigma_{2} + \beta)
\right\}
(\mathscr{U}_{\delta\theta}^{A})^{-1}
\mathscr{U}_{\delta\theta}^{A}
|C_{p} (n_{p}, m_{p}) \rangle =0
\nonumber \\
\Rightarrow
&
\left\{
\theta(\sigma_{2}) +\theta (\sigma_{2} + \beta)
+2\delta \theta
\right\}
\mathscr{U}^{A}_{\delta\theta}
|C_{p} (n_{p}, m_{p}) \rangle =0.
\end{align}
Thus, the cross-cap condition is transformed into
\begin{align}
&
\quad
\theta(\sigma_{2}) +\theta (\sigma_{2} + \beta)
=0
\nonumber \\
&
\Rightarrow
\theta(\sigma_{2}) +\theta (\sigma_{2} + \beta)
+2\delta \theta
=0,
\end{align}
which is invariant when
\begin{align}
2 \delta \theta & = 0 \mod \frac{2\pi \alpha'}{R}.
\end{align}
Thus, at least classically, we expect the theory to be invariant when $\delta \theta =\pi \alpha'/R$.
On the other hand, this invariance may not be maintained at the quantum level.
The cross-cap state may not be invariant under $\mathscr{U}^A_{\delta \theta = \pi \alpha'/R}$
and can pick up an anomalous phase;
one can easily check
\begin{align}
{\mathscr U}^A_{\pi \alpha'/R} |C_p (n_{p}, 0) \rangle = (-1)^{n_{p}}| C_p (n_{p},0) \rangle.
\end{align}
Thus the edge theory, when enforced parity symmetry with $n_{p}=1$,
is anomalous for $U(1)_A$ symmetry.
This invariance/non-invariance is equivalent to the
invariance/non-invariance of the Klein bottle partition function under
large gauge transformations discussed in Ref.\ \onlinecite{Hsieh2014}.
Observe also that the anomalous phase here is ${\mathbb Z}_{2}$-valued (i.e., a sign),
and this implies that the classification is ${\mathbb Z}_{2}$ since
the two copies of the theory is trivial.
This agrees with the loop channel calculation.\cite{Hsieh2014, Hsieh_CPT}
SPT phases protected by $CP \rtimes U(1)_V$
can be discussed in the same manner as $P \rtimes U(1)_A$.
The cross-cap state obtained by twisting $CP$ is given by
$|C_{cp} (0, m_{cp})\rangle$ in Eq.\ (\ref{xcapstate:cp}).
The $U(1)_{V}$ symmetry is generated by
${\mathscr U}^{V}_{\delta \phi} = \exp(ip \delta \phi)$ as
\begin{align}
{\mathscr U}^{V}_{\delta \phi}&: \theta \rightarrow \theta,
\quad
\phi \rightarrow \phi + \delta \phi.
\end{align}
Under this symmetry action, the cross-cap condition \eqref{xcapcondition:cp} with $n_{cp}=0$
is invariant when
\begin{align}
2\delta \phi & = 0 \mod 2\pi R.
\end{align}
On the other hand, we find
\begin{align}
{\mathscr U}^{V}_{\delta \phi = \pi R} |C_{cp} (0, m_{cp}) \rangle = (-1)^{m_{cp}}| C_{cp} (0,m_{cp}) \rangle.
\end{align}
Thus the edge theory with parity symmetry with $m_{cp}=1$ and $U(1)_V$ is anomalous.
\cite{Hsieh2014, Hsieh_CPT}
Here the anomalous phase is the ${\mathbb Z}_{2}$ sign,
and thus classification is ${\mathbb Z}_{2}$ as in the previous case.
The above analysis in terms of the cross-cap states is a reformulation of Ref. \onlinecite{Hsieh2014};
while in Ref. \onlinecite{Hsieh2014} the Klein bottle partition functions are computed in the loop channel picture,
here we have considered and studied the cross-cap states in the tree channel.
A few technical remarks are in order.
(i)
We have considered adiabatic transformations of the cross-cap states
by acting with
$\mathscr{U}^A_{\delta\theta}$ or $\mathscr{U}^V_{\delta\phi}$
and by continuously changing
$\delta \theta$ or $\delta \phi$, respectively.
In terms of the original, loop channel picture,
these twisting parameters should appear as
a twisting angle in twisting boundary conditions in the spatial direction.
This can be seen explicitly as follows.
By using the formula
$
\mathscr{G}|C_P\rangle = |C_{ g P g^{-1}}\rangle,
$
the partition function in the tree channel can be written as,
\begin{align}
Z^K &=
\langle C_{P'} |
e^{ -\frac{\ell}{2} H_{\mathrm{tree}} }
\mathscr{U}^A_{\delta \theta}
|C_P\rangle
\nonumber \\
&= \langle C_{P'} |
e^{-\frac{\ell}{2} H_{\mathrm{tree} } }
| C_{ U^A_{\delta \theta} \cdot P \cdot U^{A}_{-\delta \theta}} \rangle.
\end{align}
This can be written in the loop channel as
\begin{align}
Z^K &=
\mathrm{Tr}^{\ }_{ P' U^A_{\delta\theta} P^{-1} U^{A}_{-\delta \theta} }\, e^{-\beta H_{\mathrm{loop}} }
\end{align}
By noting
$U^A_{\delta \theta} P^{-1} U^A_{-\delta\theta}= P^{-1} U^A_{-2\delta \theta}$,
and taking $P'=P$,
\begin{align}
Z^K = \mathrm{Tr}_{U^{A}_{-2\delta\theta}}
e^{-\beta H_{\mathrm{loop} }},
\end{align}
i.e., $2\delta \theta$ (not $\delta \theta$)
appears, in the loop channel, as a twisting angle for a spatial boundary condition.
When $2\delta \theta$ is an integer multiple of $2\pi$,
the twisted system is large gauge equivalent to the system without twisting.
(ii)
The current analysis in the tree-channel picture has a few advantages over the loop-channel calculations.
First of all,
in Ref.\ \onlinecite{Hsieh2014},
we rely heavily on the existence of a continuous $U(1)$ symmetry
(either $U(1)_A$ or $U(1)_V$) as in Laughlin's thought experiment in the quantum Hall effect.
Therefore, it is not entirely obvious how the methodology in Ref.\ \onlinecite{Hsieh2014}
can be generalized to SPT phases which lack continuous ($U(1)$) symmetries.
The reformulation in the tree channel and in term of cross-cap states, however,
indicates a natural way to deal with SPTs without $U(1)$ symmetry.
In fact, the procedure described above can be generalized to cases without
$U(1)$ symmetry.
(See the following sections.)
Second,
in the current reformulation in the tree channel,
there is no need to compute the full partition function,
although it is possible to compute the partition function by using cross-cap states.
\subsubsection{$P \times C$ symmetry}
To demonstrate that our methodology works in the absence of a continuous $U(1)$ symmetry,
we now consider SPT phases protected by $P \times C$.
Here the relevant cross-cap state is $|C_p (n_{p},m_{p}) \rangle$ in Eq.\ \eqref{xcapstate:p}.
We consider $C$-symmetry \eqref{csymmetry:phase} with $n_{c}=m_{c}=0$.
One can check easily that under the action of $C$-symmetry, the cross-cap condition \eqref{xcapcondition:p} is invariant.
On the other hand, the cross-cap state may not, as one can check explicitly as
\begin{align}
\mathscr{C} |C_p (n_{p}, m_{p}) \rangle = e^{i\delta} (-1)^{n_{p}m_{p}}| C_p (n_{p},m_{p}) \rangle,
\label{1}
\end{align}
by using Eq.\ \eqref{csymmetry:momentum}.
To judge if the cross-cap state is anomalous or not,
we need to know the overall phase $e^{i \delta}$, which originates from
our ignorance on the charge conjugation parity of the zero mode sector.
Depending on our choice of $\delta$,
either one of phases with
$(n_p,m_p)=(0,0),(0,1), (1,0)$ or $(n_p,m_p)=(1,1)$ is anomalous.
A natural choice would be $e^{i\delta}=1$,
as this means we assign the charge conjugation parity
$+1$ to the reference state $|p,\tilde{p}\rangle = |0,0\rangle$.
With this choice, the edge theory with the symmetry group $P\times C$
is anomalous when $(n_{p},m_{p})=(1,1)$:
this result is consistent with the microscopic analysis given in Ref. \onlinecite{Hsieh_CPT}.
The phase acquired by the cross-cap state under the action of $C$-symmetry is ${\mathbb Z}_{2}$ (i.e., sign),
and hence the classification is ${\mathbb Z}_{2}$.
Furthermore, one may twist $CP$ symmetry,
which can be generated by the multiplication of $C$ and $P$, i.e., $CP = C \times P$,
to form the cross-cap state $|C_{cp}, (n_p, m_p) \rangle$ and then study the action of $C$ on the cross-cap:
\begin{align}
\mathscr{C} |C_{cp} (n_{p}, m_{p}) \rangle = e^{i\delta'} (-1)^{n_{p}m_{p}}| C_{cp} (n_{p},m_{p}) \rangle,
\end{align}
where $e^{i\delta'}$ is an overall phase factor,
which, as in Eq.\ \eqref{1}, cannot be determined.
This again generates the same classification as Eq.\ \eqref{1}.
This situation should be contrasted with the cases with
a continuous $U(1)$ symmetry; in the latter case,
one can compare the phase
acquired by the cross-cap state when acting with
$\mathscr{U}^V_{\delta\theta}$
and
$\mathscr{U}^V_{\delta\theta+\pi R}$;
One can follow the evolution of $\mathscr{U}^V_{\delta\theta}|C_{cp} (n_{cp}, m_{cp})\rangle$
adiabatically.
In the present case, because of the discrete nature of $C$-transformation,
such comparison is not possible.
\subsubsection{$P\times T$ symmetry}
We now consider the cases where we have both $P/CP$ and $T/CT$ symmetries.
As in the case of $P\times C$, we can first twist $P/CP$ to obtain the corresponding cross-cap state.
The action of the remaining symmetry, $T/CT$, on the cross-cap state can be studied.
However, unlike $C$-symmetry, which is a non-spatial unitary symmetry,
how $T/CT$ symmetry acts in the tree-channel is non-trivial.
We illustrate this point for the case of $P\times T$ symmetry.
Let us consider the bosonic edge theory
in the presence of $P$-symmetry with $m_{p}=0$
and $T$-symmetry with $n_T=0$.
We twist $P$-symmetry and consider the corresponding cross-cap state
$|C_p (n_p,0)\rangle$ in Eq.\ \eqref{xcapstate:p}.
One can check the cross-cap condition is left invariant under $T$-symmetry:
the fields obey, in the loop channel, the following boundary conditions:
\begin{align}
\phi(x_1, x_2+\beta) &\equiv \phi(\ell-x_1, x_2) + n_p \pi R,
\nonumber \\
\phi(x_1+\ell, x_2) &\equiv \phi(x_1, x_2),
\nonumber \\
\theta(x_1, x_2+\beta) &\equiv -\theta(\ell-x_1, x_2),
\nonumber \\
\theta(x_1+\ell, x_2) &\equiv \theta(x_1, x_2).
\label{62}
\end{align}
Under time-reversal, these conditions are transformed into:
\begin{align}
& \phi(x_1, -x_2) \equiv \phi(\ell-x_1, \beta -x_2) + n_p \pi R,
\nonumber \\
& \phi(x_1+\ell, \beta-x_2) \equiv \phi(x_1, \beta-x_2),
\nonumber \\
& -\theta(x_1, -x_2)+m_T\frac{\pi \alpha'}{R} \equiv
\theta(\ell-x_1, \beta-x_2)-m_T \frac{\pi \alpha'}{R},
\nonumber \\
& -\theta(x_1+\ell, \beta-x_2)+m_T \frac{\pi\alpha'}{R} \equiv
-\theta(x_1, \beta-x_2) + m_T \frac{\pi\alpha'}{R}.
\label{63}
\end{align}
For example, we can derive the first line of Eq.\eqref{62} from the first line of Eq.\eqref{63} by following steps.
\begin{align}
&\phi(x_1, x_2+\beta) \equiv \phi(\ell-x_1, x_2) + n_p \pi R, \nonumber\\
&\leftrightarrow \phi(x_1, y_2) \equiv \phi(\ell-x_1, y_2 - \beta) + n_p \pi R, \nonumber\\
&\to \phi(x_1, -y_2) \equiv \phi(\ell-x_1, -y_2 + \beta) + n_p \pi R, \nonumber\\
&\leftrightarrow \phi(x_1, -x_2) \equiv \phi(\ell-x_1, \beta - x_2) + n_p \pi R,
\end{align}
where we renamed the variable $y_2 = x_2 + \beta$ inbetween the first and the second lines, and we acted $T$-symmetry inbetween the second and the third lines. With the compactification conditions, these are equivalent to
\begin{align}
\phi(x_1, -x_2) &\equiv \phi(\ell-x_1, \beta-x_2) + n_p \pi R,
\nonumber \\
\phi(x_1+\ell, \beta-x_2) &\equiv \phi(x_1, \beta-x_2),
\nonumber \\
\theta(x_1, -x_2) &\equiv -\theta(\ell-x_1, \beta-x_2),
\nonumber \\
\theta(x_1+\ell, \beta-x_2) &\equiv
\theta(x_1, \beta-x_2),
\end{align}
which, with mere relabeling $\beta-x_2\to x_2$,
are equivalent to the original boundary conditions.
Since $T$-transformation leaves the $P$-twisted boundary condition invariant,
we expect the cross-cap state is also invariant under $T$-transformation,
at least up to a phase factor.
(Recall that we are after this possible anomalous phase of the cross-cap state.)
To compute this phase, we need to know how $T$-transformation looks like
in the tree channel.
In the tree channel, $T$-transformation should look like a parity transformation
(in fact, $CP$-transformation, since $T$ flips the sign of $\phi$):
\begin{align}
\phi(\sigma_1,\sigma_2) &\to -\phi(\sigma_1, 2\beta-\sigma_2),
\nonumber \\
\theta(\sigma_1,\sigma_2)&\to \theta(\sigma_1, 2\beta-\sigma_2)
+m_T \frac{\pi \alpha'}{R}.
\end{align}
This $CP$-transformation in the tree-channel picture, obtained from $T$-transformation in the loop-channel picture,
is denoted by $\widetilde{CP}$ in the following.
If the phase field $\phi$ and its dual $\theta$ obey the standard commutation
relation, this $\widetilde{CP}$-transformation must be unitary,
i.e., it must preserve, in particular, the Heisenberg algebra obeyed by
the zero modes, $[x_L, p_L]=[x_R,p_R]=i$.
If it were defined in the original coordinates $(x_1,x_2)$,
this unitary $\widetilde{CP}$-transformation is CPT-dual of $T$-symmetry.
Since $T$-transformation in the loop channel
preserves the boundary condition, this should be so in tree channel as well.
The cross-cap conditions
\begin{align}
&
\left[
\phi(\sigma_2) - \phi(\sigma_2 + \beta) - n_p \pi R
\right]
| C_p (n_p,0) \rangle
=
0,
\nonumber \\
&
\left[
\theta(\sigma_2) + \theta(\sigma_2 + \beta)
\right]
| C_p (n_p,0) \rangle
=
0,
\end{align}
are transformed into, by the $\widetilde{CP}$ transformation,
\begin{align}
&
\left[
-\phi(2\beta-\sigma_2) + \phi(\beta-\sigma_2 ) - n_p \pi R
\right]
\widetilde{\mathscr{CP}}| C_p (n_p,0) \rangle
=
0,
\nonumber \\
&
\Big[
\theta(2\beta-\sigma_2)
+ m_T \frac{\pi \alpha'}{R}
\nonumber \\
&\qquad
+ \theta(\beta-\sigma_2)
+ m_T \frac{\pi \alpha'}{R}
\Big]
\widetilde{\mathscr{CP}}| C_p (n_p,0) \rangle
=
0.
\end{align}
Due to the compactification condition,
these cross-cap conditions are equivalent to the original conditions.
While the crosscap conditions are preserved by $\widetilde{CP}$,
the cross-cap state $|C_p (n_p ,0) \rangle$ \eqref{xcapstate:p} is not invariant
when $n_p =m_T = 1$,
\beq
\widetilde{\mathscr{CP}}|C_p (n_p,0) \rangle = (-1)^{m_T n_p} |C_p (n_p,0) \rangle.
\eeq
Thus $\widetilde{CP}$ or the original time-reversal symmetry is anomalous.
This result is consistent with the gapping potential analysis in Ref.\ \onlinecite{Hsieh_CPT}.
In Appendix \ref{Bosonic SPT phases app},
we present identification of quantum anomalies using cross-cap states
for other bosonic SPT phases studied in Ref.\ \onlinecite{Hsieh_CPT}.
\section{Real fermionic SPTs}
\label{Real fermionic SPTs}
In this section,
we discuss edge theories of (2+1)d topological superconductors in the presence of discrete symmetries.
In particular, we consider topological superconductors with reflection symmetry in
symmetry classes D+R$_{+}$ and BDI + R$_{++}$, and their CPT partners
discussed in Refs.\ \onlinecite{Chiu2013, Morimoto2013, Shiozaki2014},
At quadratic level, i.e., in the absence of interactions,
topological classification for these symmetry classes
are found to be $\mathbb{Z}_2$ and $\mathbb{Z}$, respectively.
For the latter, it was found in Ref.\ \onlinecite{yao2013interaction}, the $\mathbb{Z}$ classification collapses into
$\mathbb{Z}_8$ in the presence of interactions.
Our discussion in this section can trivially be extended to other fermionic SPT phases
supporting complex (Dirac) fermion edge modes.
\subsection{Majorana fermion edge states}
Consider an edge theory of a (topological) superconductor
consisting of $N_f$ flavors of non-chiral real (Majorana) fermions
described by the Hamiltonian
\begin{align}
H
&=
\sum^{N_f}_{a=1}
\int^{\ell}_0dx\,
\left[
\psi^a_L ( -v i\partial_x) \psi^a_L
+
\psi^a_R ( + v i \partial_x) \psi^a_R
\right].
\label{fermionic edge theory}
\end{align}
The fermi velocity $v$ is set to be identical for all fermion flavors
for simplicity.
The fermion fields obey the canonical anticommutation relations
\begin{align}
\{ \psi^a_L(x), \psi^b_L(x')\} &= 2\pi \delta^{ab} \sum_{m\in \mathbb{Z}} \delta(x-x'+\ell m),
\nonumber \\
\{ \psi^a_R(x), \psi^b_R(x')\} &= 2\pi \delta^{ab} \sum_{m\in \mathbb{Z}} \delta(x-x'+\ell m).
\end{align}
The (1+1)d non-chiral fermionic edge theory (\ref{fermionic edge theory})
can be realized at the edge of (2+1)d topological superconductors
in symmetry classes DIII, D+R$_{+}$ and BDI + R$_{++}$.
The fermionic edge theory (\ref{fermionic edge theory})
is invariant under the following three symmetries:
(i)
The fermion number parity conservation,
where the fermion parity operator is given by
\begin{align}
\mathscr{G}_f =
(-1)^{F},
\quad
F =\sum^{N_f}_{a=1} F_a,
\end{align}
where $F_a$ is the total fermion number operator for the $a$-th flavor,
\begin{align}
F_a = \frac{1}{2\pi} \int^{\ell}_0dx\, i \psi^a_L \psi^a_R.
\end{align}
The fermion number parity conservation is the most fundamental symmetry of
any fermionic system.
In the following, the fermion number parity conservation
is assumed to be always unbroken
(at least classically -- it may however be anomalous).
(ii)
Time-reversal symmetry.
We consider two-kinds of time-reversal:
one which squares to $+1$:
\begin{align}
&
\mathscr{T} \psi_L(t,x) \mathscr{T}^{-1} = \psi_R(-t,x),
\nonumber \\
&
\mathscr{T} \psi_R(t,x) \mathscr{T}^{-1} = \psi_L(-t,x),
\nonumber \\
& \quad \mathscr{T}^2=1,
\quad
\mathscr{T}i \mathscr{T}^{-1} = -i,
\end{align}
and
the other which squares to $-1$:
\begin{align}
&
\mathscr{T} \psi_L(t,x) \mathscr{T}^{-1} = \psi_R(-t,x),
\nonumber \\
&
\mathscr{T} \psi_R(t,x) \mathscr{T}^{-1} = -\psi_L(-t,x),
\nonumber \\
& \quad \mathscr{T}^2=\mathscr{G}_f,
\quad
\mathscr{T}i \mathscr{T}^{-1} = -i.
\end{align}
(The flavor index is suppressed.)
(iii)
Parity symmetries.
Similarly to time-reversal,
we consider two-kinds of parities:
one which squares to $+1$:
\begin{align}
&\mathscr{P} \psi_L(t,x) \mathscr{P}^{-1} = \psi_R(t,\ell-x),
\nonumber \\
&\mathscr{P} \psi_R(t,x) \mathscr{P}^{-1} = \psi_L(t,\ell-x),
\nonumber \\
&\quad
U_P = \sigma_x, \qquad \mathscr{P}^2=1,
\end{align}
and
the other which squares to $-1$:
\begin{align}
&\mathscr{P} \psi_L(t,x) \mathscr{P}^{-1} = \psi_R(t,\ell-x),
\nonumber \\
&\mathscr{P} \psi_R(t,x) \mathscr{P}^{-1} = -\psi_L(t,\ell-x).
\nonumber \\
& \quad
U_P = i \sigma_y, \qquad \mathscr{P}^2 = -1.
\end{align}
Note that the phase of parity (and hence parity squared) is arbitrary
for systems of complex fermions whereas
there is no such arbitrariness for real fermions.
Observe that
$\mathscr{T}^2=-1$ and $\mathscr{P}^2=1$ are CPT conjugate to each other.
Similarly,
$\mathscr{T}^2=1$ and $\mathscr{P}^2=-1$ are CPT conjugate to each other.
(These can easily be inferred from
the fact that they both prohibit the same kind of masses.)
In the absence of time-reversal and parity symmetries,
the non-chiral fermion edge state can easily be gapped by adding a mass term.
To realize a non-trivial SPT phase,
we need to impose $\mathscr{T}$, $\mathscr{P}$, or both $\mathscr{T}$ and $\mathscr{P}$
(see Table \ref{table}).
Imposing $\mathscr{T}^2=-1$ alone leads to
$\mathbb{Z}_2$ topological classification (class DIII)
while imposing $\mathscr{T}^2=1$ alone does not give rise to a SPT.
Similarly, imposing $\mathscr{P}^2=1$ alone leads to
$\mathbb{Z}_2$ topological classification (class D + R$_+$)
while imposing $\mathscr{P}^2=-1$ alone does not give rise to a SPT (class D+R$_-$).
Imposing $\mathscr{T}^{2}=-1$ together with $\mathscr{P}^2=-1$ (class DIII + R$_{--}$)
gives rise to $\mathbb{Z}_8$ topological distinction.
Similarly, its CPT conjugate,
$\mathscr{P}^2=1$ together with $\mathscr{T}^2=+1$ (class BDI + R$_{++}$)
gives rise to $\mathbb{Z}_8$ topological distinction.
In the following, our task is to understand
these $\mathbb{Z}_2$ (class D+R$_{+}$) and
$\mathbb{Z}_8$ (class BDI+R$_{++}$) classifications of SPTs in terms of
quantum anomalies.
We will apply our formalism using cross-cap states.
\begin{table}
\begin{tabular}{c|l|l|c}
Class & $T$ & $P$ & Classification
\\ \hline
DIII & $\mathscr{T}^2=\mathscr{G}_f$ & None & $\mathbb{Z}_2$ \\
BDI & $\mathscr{T}^2=1$ & None & 0 \\
D+R$_{+}$ & None & $\mathscr{P}^2=1$ & $\mathbb{Z}_2$ \\
D+R$_{-}$ & None & $\mathscr{P}^2=-1$ & 0 \\
DIII+R$_{--}$ & $\mathscr{T}^2=\mathscr{G}_f$ & $\mathscr{P}^2=-1$ & $\mathbb{Z}\to \mathbb{Z}_8$ \\
BDI+R$_{++}$ & $\mathscr{T}^2=1$ & $\mathscr{P}^2=1$ & $\mathbb{Z}\to \mathbb{Z}_8$ \\
\end{tabular}
\caption{
Edge theories of (2+1)d topological superconductors with various time-reversal and parity symmetries
studied in Sec.\ \ref{Real fermionic SPTs}.
\label{table}
}
\end{table}
\subsection{Cross-cap states}
\paragraph{Cross-cap condition}
In both symmetry classes D+R$_{+}$ and DIII+R$_{--}$,
parity $\mathscr{P}$ and fermion number parity $G_f$ are conserved.
Hence, theories in these symmetry classes can be twisted by
parity $\mathscr{P}$, and the combination of fermion number parity and spatial parity $G_f \mathscr{P}$.
In the fermionic edge theory (\ref{fermionic edge theory}),
twisting boundary conditions in time direction by $\mathscr{P}$ or $\mathscr{G}_f \mathscr{P}$
leads to the following conditions on the fermion fields:
\begin{align}
\psi_L(x_1, x_2+\beta) =\eta_1 \psi_R(\ell-x_1, x_2),
\nonumber \\
\psi_R(x_1, x_2+\beta) = \eta_2 \psi_L(\ell-x_1, x_2).
\end{align}
Here, $\eta_{1,2}$ are given by
\begin{align}
(\eta_1, \eta_2) =
\left\{
\begin{array}{ll}
(+,+)& \mbox{twisting by $\mathscr{P}$ with $\mathscr{P}^2=1$}
\\
(-,-)& \mbox{twisting by $\mathscr{G}_f \mathscr{P}$ with $\mathscr{P}^2=1$}
\\
(+,-)& \mbox{twisting by $\mathscr{P}$ with $\mathscr{P}^2=-1$}
\\
(-,+)& \mbox{twisting by $\mathscr{G}_f \mathscr{P}$ with $\mathscr{P}^2=-1$}
\end{array}
\right.
\end{align}
In the rearranged geometry (see Fig.\ \ref{fig:xCap}),
these twisted boundary conditions are given by
\begin{align}
\psi_L(\sigma_1, \sigma_2) &=\eta_1 \psi_R(\sigma_1, \sigma_2+\beta),
\nonumber \\
\psi_R(\sigma_1, \sigma_2) &= \eta_2 \psi_L(\sigma_1, \sigma_2+\beta).
\end{align}
By making a 90 degree rotation in the $(\sigma_1, \sigma_2)$ plane,
$(\sigma_1, \sigma_2) \to (\sigma_1', \sigma_2') = (\sigma_2, -\sigma_1)$,
we exchange the role of time and space coordinates and regard $\sigma_1$ direction as a fictitious time direction.
The fermion fields $\psi^{\prime}_{L,R}$ with respect to the rotated coordinate system $(\sigma_1', \sigma_2')$
are related to the original fermion fields as
\begin{align}
\psi^{\prime}_L = e^{i\pi/4} \psi_L,
\quad
\psi^{\prime}_R = e^{-i\pi/4} \psi_R.
\end{align}
The cross-cap states are defined by the cross-cap conditions
\begin{align}
\left[
\psi_L (\sigma_2)- i \eta_1 \psi_R (\sigma_2+\beta)
\right]
|C (\eta_1,\eta_2) \rangle &=
0,
\nonumber \\
\left[
\psi_R (\sigma_2)+ i \eta_2 \psi_L (\sigma_2+\beta)
\right]
|C (\eta_1, \eta_2) \rangle &=0,
\end{align}
where for notational simplicity we have removed
${ }^{\prime}$ and simply
write $\psi^{\prime}_{L,R}\to \psi_{L,R}$
and $\sigma^{\prime}_{1,2}\to \sigma_{1,2}$.
A crucial observation is that
for the cross-cap conditions generated by $\mathscr{P}$ with $\mathscr{P}^2=+1$,
there are Majorana fermion zero modes in the mode expansion of the
fermion fields,
while there are no Majorana zero modes when $\mathscr{P}^2=-1$.
This follows from the general formula (\ref{cross cap formula 1}),
which suggests that when $\mathscr{P}^2=1(-1)$ the fermion fields obey periodic (anti-periodic) boundary condition.
(This can also be understood by first assuming
the existence of zero modes, $\psi_{L0}$ and $\psi_{R0}$ for
the left- and right-moving fermion fields, respectively.
Then, the cross-cap condition tells us
$\psi_{L0} - i\eta_1 \psi_{R0}=0$
and
$\psi_{R0} + i\eta_2 \psi_{L0}=0$.
With $\eta_1= -\eta_2$, these two conditions
are not compatible with each other.)
This distinction between $\mathscr{P}^2=1$ and $\mathscr{P}^2=-1$ are directly related
to the $\mathbb{Z}_2$ classification in class D+R$_+$ and
to the absence of SPT phases in class D+R$_-$.
For class D+R$_-$ ($P^2=-1$), the fermion zero modes are not allowed,
and hence the corresponding cross-cap state is constructed entirely
in terms of fermionic oscillator modes (non-zero modes).
It can be checked easily that this cross-cap state is anomaly free.
On the other hand,
for class D+R$_+$ ($P^2=+1$),
the construction of the corresponding cross-cap state is more non-trivial
because of the presence of fermion zero modes.
We will see the action of the fermion number parity operator on the cross-cap
state, $|C (\eta_1, \eta_2)\rangle$ with $\eta_1=\eta_2$,
give rise to an anomalous phase (sign), indicative of the expected $\mathbb{Z}_2$
classification.
\paragraph{Construction of cross-cap states}
Let us now construct the cross-cap state focusing on $\eta_1 = \eta_2= \eta$:
\begin{align}
\left[
\psi^a_L (\sigma_2)- i \eta \psi^a_R (\sigma_2+\beta)
\right]
|C, \eta\rangle &=
0,
\nonumber \\
\left[
\psi^a_R (\sigma_2)+ i \eta \psi^a_L (\sigma_2+\beta)
\right]
|C, \eta\rangle &=0,
\end{align}
where $a=1,\ldots, N_f$.
When $\eta_1=\eta_2$, the cross-cap condition is compatible with the periodic boundary
condition for the fermion field in $\sigma_1$ direction.
Within the zero mode sector, the cross-cap condition reads
\begin{align}
\left[
\psi^a_{0L} - i \eta \psi^a_{0R}
\right]
|C, \eta\rangle
&=
0,
\end{align}
where $\psi^a_{0L/R}$ is the zero mode for the $a$-th flavor,
satisfying $(\psi^{a}_{0L/R})^2=1$.
In the following, we construct the cross-cap states within the zero-mode sector explicitly,
together with a reference state.
[Recall the important role played by the reference state for the case of the bosonic SPT phases
discussed below Eqs.\ (\ref{csymmetry:momentum}) and (\ref{1})].
The choice of the reference state, however, is far from obvious,
due to the degeneracy in the zero mode sector.
We will illustrate this point by contrasting two constructions.
In the first construction,
we construct the Hilbert space of the zero modes
by considering
the following fermion creation/annihilation operators:
\begin{align}
f^{\dag}_{a} &=
\frac{1}{{2}}(\psi^a_{0L} + {i}\psi^a_{0R}),
\quad
f^{\ }_{a} =
\frac{1}{{2}}(\psi^a_{0L} - {i}\psi^a_{0R}),
\end{align}
and the Fock vacuum $|0_f\rangle$
of the $f$-fermions.
The cross-cap state $|C, \eta=+\rangle$ is then nothing
but
$|0_f\rangle$ itself,
\begin{align}
|C, \eta =+\rangle = e^{ i\phi_+}|0_f\rangle.
\end{align}
On the other hand,
the cross-cap state $|C, \eta=-\rangle$ can be constructed as
\begin{align}
|C, \eta =-\rangle = e^{i\phi_-} \prod_{a=1}^{N_f} f^{\dag}_a |0_f\rangle.
\end{align}
[N.B.
In the above representation of $|C,\eta\rangle$,
the ambiguous phases $\phi_{\pm}$ are not fixed by the cross-cap condition.
These phases will not affect our later analysis, and hence will be set to zero henceforth.
One common convention for the phase is
$
|C, \eta = + \rangle = e^{-i\frac{\pi}{8}}|0_f\rangle
$,
$
|C, \eta = - \rangle = e^{i\frac{\pi}{8}} f^{\dag}|0_f \rangle,
$
where we focus on the case of single flavor $N_f=1$ for simplicity.
Then we notice the following identities
$
\psi_{0L} |C, \eta \rangle = e^{-i\eta \frac{\pi}{4}}|C, -\eta \rangle
$,
$
\psi_{0R} |C, \eta \rangle = e^{-i\eta \frac{\pi}{4}}|C, -\eta \rangle.
$
One motivation for this phase convention is that
it well compares with the operator product expansions of the Ising CFT,
\cite{CFTbook}
$\psi_{L} \sigma \sim e^{i\frac{\pi}{4}}\mu$,
$\psi_{L} \mu \sim e^{-i \frac{\pi}{4}}\sigma$,
$\psi_{R} \sigma \sim e^{-i\frac{\pi}{4}} \mu$,
$\psi_{R} \mu \sim e^{i \frac{\pi}{4}} \sigma$,
where $\sigma$ and $\mu$ are the Ising spin operator and the disorder operator, respectively.
]
Alternatively,
when $N_f=\mbox{even}$,
one can introduce the following fermion creation operators
(see, for example, Ref. \onlinecite{Bergman1997}):
\begin{align}
d^{\dag}_{L j} =
\frac{1}{{2}}(\psi^{2j-1}_{0L} + {i} \psi^{2j}_{0L}),
\quad
{d}^{\dag}_{R j} =
\frac{1}{{2}}({\psi}^{2j-1}_{0R} + {i} {\psi}^{2j}_{0R}),
\end{align}
and the Fock vacuum $|0_d\rangle$ annihilated by $d_{Lj}$ and $d_{Rj}$.
We observe that
$
d_{Lj}- {i}{d}_{Rj}
=
f_{2j-1} - {i} f_{2j}
$,
$(d^{\dag}_{jL} - i d^{\dag}_{jR}) |0_f\rangle=0$,
and
\begin{align}
|C, + \rangle = |0_f\rangle = \prod_j (d^{\dag}_{jL} - i d^{\dag}_{jR}) |0_d\rangle.
\end{align}
Similarly,
\begin{align}
|C, - \rangle = \prod_{a=1}^{N_f} f^{\dag}_a |0_f\rangle = \prod_j (d^{\dag}_{j R} - i d^{\dag}_{j L}) |0_d\rangle.
\label{2}
\end{align}
(Similar construction is possible even for $N_f=\mbox{odd}$
by adding an extra Majorana fermion as in Ref. \onlinecite{Fidkowski2011}.
We will not dwell on this point
as the identification of a quantum anomaly when $N_f=\mbox{odd}$
is rather straight forward and does not depend on the choice of the reference state.)
One important feature of the second construction is the clear factorization of
the vacuum $|0_d\rangle$ into the left- and right-moving sectors,
$|0_d\rangle = |0_d\rangle_L\otimes |0_d\rangle_R$,
as it is annihilated by $d_{Lj}$ and $d_{Rj}$ separately.
The entire Hilbert space built out of $|0_d\rangle$ also factorizes into the left- and right-moving sectors.
This factorization allows us to introduce the action of reflection (time-reversal) in a transparent way.
Such factorization is also expected for general construction of
cross-cap states and boundary states in boundary conformal field theories.
(See, for example, Ref. \onlinecite{Bates2006}.)
\subsection{Fermion number parity - $\mathbb{Z}_2$ classification in class D+R$_{+}$}
We now ask the properties of the cross-cap states
under the action of symmetry generators.
As a start, let us consider the fermion number parity.
In the tree channel, its explicit form within the zero mode sector is given by
\begin{align}
\mathscr{G}_f
=
(i \psi^1_{0L}\psi^1_{0R})
(i \psi^2_{0L}\psi^2_{0R})
\cdots
(i \psi^{N_f}_{L}\psi^{N_f}_{0R}).
\end{align}
The fermion number parity acting on the cross-cap states gives
\begin{align}
\mathscr{G}_f|C, \pm \rangle = (\pm)^{N_f} |C, \pm\rangle.
\label{sign D+R+}
\end{align}
Therefore,
when $N_f=\mbox{even}$, there is no anomaly.
On the other hand,
when $N_f=\mbox{odd}$,
one would conclude,
$|C, -\rangle$ is anomalous
while
$|C, +\rangle$ is not.
This is consistent with the $\mathbb{Z}_2$ classification of
(2+1)d topological superconductors.
in symmetry class D+R$_{+}$.
Upon closer inspection, however, Eq.\ (\ref{sign D+R+}) would look strange
since the two cross-cap states $|C, \pm \rangle$
should be treated on the equal footing:
If the cross-cap $|C, \pm\rangle$ is obtained
by twisting $\mathscr{P}$,
$|C, \mp\rangle$ is obtained from simply from $\mathscr{G}_f \mathscr{P}$.
Hence, both cross-cap states $|C, \pm\rangle$ belong/correspond to the same phase.
In fact, it should be noted that there is a phase ambiguity
in defining the cross-cap states and the fermion number parity operator.
In the above analysis, we implicitly made a particular choice where
the fermion number parity of the ground state $|0_f\rangle$ is $+1$.
In principle, one could assign a different fermion number parity eigenvalue,
e.g., by modifying the definition of the fermion number parity operator,
$\mathscr{G}_f\to -\mathscr{G}_f$.
Alternatively, instead of using $f^{\dag}, f$,
one could define $c:=f^{\dag}$ and $c^{\dag}:=f$,
which leads to
$|C, -\rangle =|0_c\rangle$
and
$|C,+\rangle =\prod_a c^{\dag}|0_c\rangle$.
In this convention, it is tempting to
claim
$\mathscr{G}_f|C,-\rangle = +|C,-\rangle$
while
$\mathscr{G}_f|C,+\rangle = (-1)^{N_f}|C,+\rangle$.
Thus, there is some ambiguity when deducing
the fermion number parity eigenvalue.
We have encountered similar ambiguity when
dealing with CP symmetric bosonic SPT phases.
Such ambiguity of the fermion number parity eigenvalue of the ground state,
however, does not affect our conclusion,
since, independent of the phase choice,
when $N_f=\mbox{odd}$,
we cannot make both $|C,\pm\rangle$ anomaly-free.
Hence we conclude the $\mathbb{Z}_2$ classification of symmetry class D+R$_{+}$.
\subsection{Time-reversal - $\mathbb{Z}_8$ classification in class BDI + R$_{++}$}
We now include time-reversal symmetry.
After $\pi/2$ rotation of spacetime,
a time-reversal operator is transformed into a parity operator (unitary) (denoted by $\widetilde{\mathscr{P}}$ in the following).
Correspondingly to two time-reversal operators $\mathscr{T}^2=\pm 1$ in the loop channel picture,
there are two kinds of parity operators in the tree channel picture.
They act on the fermion zero modes as:
\begin{align}
\widetilde{\mathscr{P}} \psi_{0L} \widetilde{\mathscr{P}}^{-1} = \psi_{0R},
\quad
\widetilde{\mathscr{P}} \psi_{0R} \widetilde{\mathscr{P}}^{-1} = \psi_{0L},
\end{align}
and
\begin{align}
\widetilde{\mathscr{P}} \psi_{0L} \widetilde{\mathscr{P}}^{-1} = \psi_{0R},
\quad
\widetilde{\mathscr{P}} \psi_{0R} \widetilde{\mathscr{P}}^{-1} = -\psi_{0L}.
\end{align}
Explicitly, they can be written as
\begin{align}
\widetilde{\mathscr{P}}
=
e^{i\delta}
\prod_{a}
\frac{1}{\sqrt{2}} \left(\psi_{0L} + \psi_{0R} \right)^a
\end{align}
and
\begin{align}
\tilde{ \mathscr{P}} =
e^{i\delta}
\prod_a
\frac{1}{\sqrt{2}}
\left(
1-
\psi_{0L}\psi_{0R}
\right)^a,
\end{align}
respectively,
where $e^{i\delta}$ is an unknown phase factor and will be discussed in more detail shortly.
The first parity operator does not preserve the cross-cap condition with $\eta_1=\eta_2$,
so we will focus on the second one.
One also verifies
\begin{align}
\widetilde{\mathscr{P}}^2 = e^{2i\delta} (i)^{N_f} \mathscr{G}_f,
\label{projective}
\end{align}
and
$ \mathscr{G}_f \widetilde{\mathscr{P}}= \widetilde{\mathscr{P}}\mathscr{G}_f$.
Let us now calculate the action of $\widetilde{\mathscr{P}}$ on
the cross-cap states.
By using the representation in terms of the $f$-fermions,
$\psi_{0L}\psi_{0R} = i (2 f^{\dag} f - 1)$,
$\widetilde{\mathscr{P}}$ can be written as
\begin{align}
\widetilde{\mathscr{P}} =
e^{i\delta}
\prod_a
\frac{1}{\sqrt{2}}
\left[
1- i (2 n_a -1)
\right],
\end{align}
where $n_a = f^{\dag}_a f_a$.
Then, the action of $\widetilde{\mathscr{P}}$ on the cross-cap states
is given by
\begin{align}
\widetilde{\mathscr{P}} |C, +\rangle
&=
e^{i\delta}
\prod_a
\frac{1}{\sqrt{2}}
\left[
1- i (2 n_a -1)
\right]
|0_f\rangle
\nonumber \\
&=
e^{i\delta}
\prod_a
\frac{1}{\sqrt{2}}
\left[
1+ i
\right]
|0_f\rangle
=
e^{ +i\frac{\pi}{4}N_f}
e^{i\delta}
|C, +\rangle,
\nonumber \\
\widetilde{\mathscr{P}} |C, -\rangle
&=
e^{i\delta}
\prod_a
\frac{1}{\sqrt{2}}
\left[
1- i
\right]
|C, -\rangle
=
e^{ -i\frac{\pi}{4}N_f}
e^{i\delta}
|C, -\rangle.
\end{align}
The relative phase between
$\widetilde{\mathscr{P}}|C, +\rangle$
and
$\widetilde{\mathscr{P}}|C, -\rangle$
is $e^{ + i \pi N_f/2}$,
which
is independent of the choice of $e^{i\delta}$ (the choice of the action of $\widetilde{\mathscr{P}}$ on the reference state),
and
vanishes when $N_f = 4 \times \mbox{integer}$.
In other words,
one cannot make both cross-cap states anomaly-free unless $N_f = 4\times \mbox{integer}$.
One then immediately concludes the classification is {\it at least} $\mathbb{Z}_4$.
On the other hand,
as we have seen for the case of the bosonic SPT phases,
a proper choice of the phase $e^{i\delta}$, if exists, leads to a refined classification.
If we choose $|0_f\rangle$ as the reference state and demand $|0_f\rangle$ transform trivially under $\widetilde{\mathscr{P}}$,
$\widetilde{\mathscr{P}}|0_f\rangle = |0_f\rangle$,
we obtain $\mathbb{Z}_4$ classification.
Alternatively,
if we choose $|0_d\rangle$ as the reference state,
and demand $\widetilde{\mathscr{P}}|0_d\rangle=|0_d\rangle$,
which may be obtained from
$\widetilde{\mathscr{P}}|0_d\rangle_{L,R} = |0_d\rangle_{R,L}$,
\begin{align}
\widetilde{\mathscr{P}}|C, + \rangle
&= \widetilde{\mathscr{P}} |0_f\rangle \nonumber \\
&= \widetilde{\mathscr{P}} \prod_{j=1}^{N_f/2} (d^{\dag}_{jL} - i d^{\dag}_{jR} )|0_d\rangle
\nonumber \\
&=
\prod_{j=1}^{N_f/2} (d^{\dag}_{jR} +i d^{\dag}_{jL} )|0_d\rangle
\nonumber \\
&=
(i)^{N_f/2}|0_f\rangle.
\end{align}
I.e., $e^{i\delta} = 1$.
It can also be checked, straightforwardly,
\beq
\widetilde{\mathscr{P}}|C, - \rangle = (i)^{N_f/2} |C, - \rangle,
\eeq
following Eq.\ \eqref{2}.
Thus, with this choice, the two cross-cap states $|C,\eta\rangle$ can be both made anomaly free
only when $N_f=8\times \mbox{integer}$, i.e., $\mathbb{Z}_8$ classification.
As a final comment, we provide a yet another point of view by using Eq.\ (\ref{projective}).
Equation (\ref{projective}) suggests, within the zero mode sector, the symmetry is realized projectively.
The ``unwanted'' phase $e^{2i \delta} (i)^{N_f}$ can be removed by
choosing $e^{i\delta} = e^{- i \pi N_f/4}$.
However, with this choice, the reference state now acquires an anomalous phase
$\widetilde{\mathscr{P}}|0_d\rangle = e^{ -i \pi N_f/4} |0_d\rangle$.
This conflict between the two demands,
one to represent the symmetry group non-projectively
and
the other to make the reference state transform trivially under $\widetilde{\mathscr{P}}$,
can be considered as a form of quantum anomaly.
\section{Conclusion}
There are three known ways of symmetry breaking in nature.
First, a symmetry can be broken {\it explicitly} by adding a symmetry-breaking perturbation to the Hamiltonian.
Second, a symmetry can be broken {\it spontaneously} in many-body systems and in field theories
because of interactions.
These two forms of symmetry breaking occur both in classical and quantum systems.
The third way of symmetry breaking is more subtle and happens only in quantum many-body systems;
A symmetry can be {\it anomalous}, meaning it can be broken because of quantum effects.
In discussing SPT phases, we are to distinguish different quantum phases of matter, all respecting the same set of symmetries.
Therefore, by definition, the Landau paradigm, while powerful in discussing phases with spontaneous symmetry breaking,
cannot be applied to SPT phases.
On the other hand, as we demonstrated,
anomalous symmetry breaking (quantum anomalies)
can be useful in diagnosing and perhaps even classifying SPT phases.
The idea of using quantum anomalies to characterize topological phases of matter
goes back to Laughlin's thought experiment (Laughlin's gauge argument) in the quantum Hall states.
Our methodology can be considered as a proper generalization of Laughlin's gauge argument to SPT phases
protected by parity (reflection) and other symmetries.
In Laughlin's thought experiment,
quantized charge pumping caused by flux threading (large gauge transformation)
characterizes the quantum Hall effect even in the presence of interactions and disorder.
Within edge theories of the quantum Hall effect, the charge pumping by a large gauge transformation
appears as a quantum anomaly, i.e., non-conservation of the electric $U(1)$ charge.
The edge theories of SPT phases (if exist) differ from the edge theories of the quantum Hall systems.
First, quite generically, edge theories supported by SPT phases are non-chiral (having vanishing thermal Hall conductance).
Second, edge theories of SPT phases may be protected by discrete symmetries, and may not have a continuous $U(1)$ symmetry.
For these reasons, diagnosing edge theories of SPT phases requires to identify
quantum anomalies (breaking down of symmetries of SPT phases by quantum effects) of a subtler kind than in the quantum Hall effect.
By twisting parity in edge theories of (2+1)d topological phases,
we obtain a cross-cap state and investigate possible quantum anomalies in terms of the cross-cap state.
This method is applied explicitly to the several examples, including bosonic and fermionic SPT phases,
and reproduces the known classifications in Refs.\ \onlinecite{Hsieh_CPT, yao2013interaction}.
In conclusion,
we have provided a way to gauge the spatial symmetries to efficiently diagnose SPT phases,
which is beyond the cohomology classification tables.
We close with a few comments.
(i)
In our approach, it is crucial to fix the action of symmetry operators on a reference state in the tree-channel picture.
In particular, possible phase ambiguity should be fixed, although
in many cases a partial result is obtained even without fixing this ambiguity.
While we have provided a proposal to fix this phase ambiguity by specifying reference states,
it is desirable to have a more convincing and convenient method.
One possible approach to this problem is to use the state-operator correspondence
and assign the phase which is consistent with the operator algebra of the edge theories.
For example, such argument is used in Ref. \onlinecite{Gimon1996} in somewhat similar context.
Following this type of approach is left for the future research.
(ii)
We treated SPT phases where parity and time-reversal are in conflict quantum mechanically,
in that enforcing parity by orientifolding leads to breakdown of time-reversal,
e.g., class BDI+R$_{++}$ topological superconductors.
We choose to twist parity since twisting unitary symmetry appears to be more straightforward,
while, in principle, we could equally consider twisting by time-reversal.
In fact, after exchanging the role of space and time coordinates and when considering
cross-cap states,
what appears to be time-reversal symmetry in the original, loop-channel picture
is represented as a parity (or $CP$) symmetry in the tree-channel picture.
This indicates that by exchanging the role of space and time coordinates,
twisting time-reversal is neither more nor less than twisting parity, and should lead to an orientifold theory.
Our analysis in this paper thus provides an insight into SPT phases protected, not by parity, but by time-reversal symmetry
and other symmetry,
e.g., (2+1)d time-reversal symmetric topological insulators (the quantum spin Hall effect),
which are protected by $U(1)_V$ and time-reversal symmetries.
Orientifolds of gapless edge states discussed in this work
may provide a complementary view to, e.g.,
Ref. \onlinecite{Chen_Gauging_Tsymmetry},
which discusses a way to gauge time-reversal in gapped bulk SPT phases
by using the tensor network representation of quantum states.
(iii)
In our approach,
the use of cross-cap states, upon going from the loop to tree channel picture,
most clearly demonstrates how quantum anomalies may appear.
In doing so, we exchange the role of space and time coordinates.
The fictitious time-evolution and the fictitious Hilbert space in the tree channel picture
is in fact akin to
the row-to-row transfer matrix and to the auxiliary space of matrix product states (MPSs), respectively.
The projective representation of the symmetry group of SPT phases on the auxiliary space of MPSs has been successfully used
to diagnose and classify (1+1)d gapped SPT phases.
Our use of cross-cap states and the associated anomalous phases have some similarities with the
classification of (1+1)d gapped SPT phases by MPSs.
(iv)
In our previous calculations \cite{Hsieh2014},
the full partition functions on the Klein bottle are computed
explicitly and checked for the presence/absence of quantum anomalies.
In the current reformulation in the tree channel,
there is no need to compute the partition function,
although it is possible to compute the partition function by using cross-cap states.
In addition, we do not rely on the continuous $U(1)$ symmetry, and hence can treat
a wider class of SPT phases that are not discussed in Ref. \onlinecite{Hsieh2014}.
The tree-channel formulation also somewhat liberates us from the unwanted reliance on relativistic and conformal symmetries
when discussing topological classification of phases of matter.
As an MPS can be constructed for an arbitrary (1+1)d (gapped as well as gapless) quantum states,
cross-cap states can in principle be constructed for general edge theories which lack relativistic invariance.
If it comes to compute the full partition function, the lack of the relativistic invariance may be inconvenient,
as the Hamiltonian in the loop and tree channels do not agree.
The partition function in such situations would not be expressed in terms of well-known functions such as the Jacobi theta functions.
The importance, necessity and limitations of relativistic and conformal invariance,
when following our approach to SPT phases,
are however not entirely understood currently, and are left for the future studies.
It would be interesting to note that, in constructing cross-cap states and, similarly,
in constructing conformal invariant boundary states in boundary conformal field theories,
the Virasoro algebra plays a major role.
\acknowledgements
We thank Xie Chen and Andreas Ludwig
for helpful discussion.
This work is supported
by
the NSF under Grant No. DMR-1064319 (GYC), No. DMR 1408713 (GYC), and DMR-1455296 (SR),
the Brain Korea 21 PLUS Project of Korea Government (GYC),
and Alfred P. Sloan foundation (SR).
|
\section{Introduction}
\label{sc:introduction}
Let $\bp = (p,q,r)$ be a triple of positive integers satisfying
\begin{align}
\frac{1}{p} + \frac{1}{q} + \frac{1}{r} < 1.
\end{align}
The \emph{Schwarz triangle group}
\begin{align}
\Delta = \la a, b, c \relmid a^2 = b^2 = c^2 = (ab)^p = (bc)^q = (ca)^r = e \ra
\end{align}
is the reflection group generated by reflections
along edges of the hyperbolic triangle
in the upper half plane $\bH$
with angles $\pi/p$, $\pi/q$, and $\pi/r$.
The \emph{von Dyck group}
\begin{align}
\Gamma = \la x, y, z \relmid x^p = y^q = z^r = x y z = e \ra
\end{align}
is the subgroup of $\Delta$ of index two
consisting of products of even numbers of reflections.
It is a cocompact Fuchsian group,
and the orbifold quotient
$\bX = [\bH/\Gamma]$
has three orbifold points
with stabilizers $\bZ/p \bZ$,
$\bZ/q \bZ$, and $\bZ/r \bZ$.
The triple $\bp = (p,q,r)$ is called the \emph{signature}
of the Fuchsian group $\Gamma$.
Smooth rational orbifolds are studied in detail
by Geigle and Lenzing \cite{Geigle-Lenzing_WPC}
under the name of \emph{weighted projective lines}.
In particular,
the orbifold $\bX$ can be described as
\begin{align}
\bX = [ (\Spec T \setminus \bszero) / K],
\end{align}
where
\begin{align}
T
= \bC[X,Y,Z]/(X^p+Y^q+Z^r)
= \bigoplus_{\veck \in L} T_\veck
\end{align}
is a two-dimensional ring
graded by the abelian group
\begin{align}
L = \bZ \vecX \oplus \bZ \vecY \oplus \bZ \vecZ \oplus \bZ \vecc /
(p \vecX - \vecc, q \vecY - \vecc, r \vecZ - \vecc)
\end{align}
of rank one,
which is the group of characters of the algebraic group
\begin{align}
K = \Spec \bC[L].
\end{align}
Here the grading is given by $X \in T_\vecX$, etc.
The \emph{dualizing element} is defined by
\begin{align}
\vecomega = \vecc - \vecX - \vecY - \vecZ,
\end{align}
and the canonical ring of the orbifold $\bX$ is given by
\begin{align}
R = \bigoplus_{k=0}^\infty
R_k, \qquad R_k =
T_{k \vecomega}.
\end{align}
The isolated singularity at the origin of the scheme $\Spec R$
is called the \emph{triangle singularity}
with signature $\bp=(p, q, r)$.
Let
$
\frakm
= \bigoplus_{k=1}^\infty R_k
$
denote the irrelevant ideal of $R$.
The dimension of the vector space $\frakm / \frakm^2$ is called
the \emph{embedding dimension}
of $R$.
It is known that the embedding dimension of $R$ coincides with the minimum number of generators of $R$ (see \pref{lm:embdim}).
A graded ring is said to be a \emph{hypersurface}
if the embedding dimension is greater
than the Krull dimension by one.
\begin{theorem}[{\cite{MR0568895}, cf.~also~\cite{Milnor_3BM,MR586722}}]
\label{th:Dolgachev}
The ring $R$ is a hypersurface
if and only if the signature $\bp$ is one of the 14 signatures
shown in \pref{tb:Dolgachev}.
\end{theorem}
\begin{table}[t]
\begin{align*}
\begin{array}{cccc}
\toprule
\text{Signature} & \text{Generators} & \text{Weights} & \text{Relation} \\
\midrule
(2, 3, 7) & (X,Y,Z) & (21, 14, 6;42) & x^2+y^3+z^7 \\
(2, 3, 8) & (XZ,Y,Z^2) & (15, 8, 6;30) & x^2+y^3 z+z^5 \\
(2, 3, 9) & (X,YZ,Z^3) & (9, 8, 6;24) & x^2 z+y^3+z^4 \\
(2, 4, 5) & (XY,Y^2,Z) & (15, 10, 4;30) & x^2+y^3+y z^5 \\
(2, 4, 6) & (XYZ,Y^2,Z^2) & (11, 6, 4;22) & x^2+y^3 z+y z^4 \\
(2, 4, 7) & (XY,Y^2 Z,Z^3) & (7, 6, 4;18) & x^2 z+y^3+y z^3 \\
(2, 5, 5) & (Y^5,X,YZ) & (10, 5, 4;20) & x^2+x y^2+z^5 \\
(2, 5, 6) & (Y^4,XZ,Y Z^2) & (6, 5, 4;16) & x y^2+x^2 z+z^4 \\
(3, 3, 4) & (X^3,XY,Z) & (12, 8, 3;24) & x^2+x z^4+y^3 \\
(3, 3, 5) & (X^3 Z,XY,Z^2) & (9, 5, 3;18) & x^2+x z^3+y^3 z \\
(3, 3, 6) & (X^3,XYZ,Z^3) & (6, 5, 3;15) & x^2 z+x z^3+y^3 \\
(3, 4, 4) & (X Y^4,X^2,YZ) & (8, 4, 3;16) & x^2+x y^2+y z^4 \\
(3, 4, 5) & (X Y^3,X^2 Z,Y Z^2) & (5, 4, 3;13) & x y^2+x^2 z+y z^3 \\
(4, 4, 4) & (X^4,Y^4,XYZ) & (4, 4, 3;12) & x^2 y+x y^2+z^4 \\
\bottomrule
\end{array}
\end{align*}
\caption{Hypersurface triangle singularities}
\label{tb:Dolgachev}
\end{table}
\pref{tb:Dolgachev} coincides
with the list of weighted homogeneous exceptional unimodal singularities
classified by Arnold \cite{Arnold_ICM}.
The signature in this context is called the \emph{Dolgachev number}
of the exceptional unimodal singularity.
Besides the Dolgachev number,
each exceptional unimodal singularity comes with another triple of positive integers
called the \emph{Gabrielov number},
which describes the Milnor lattice of the singularity
\cite{MR0367274}.
This leads to the discovery of \emph{strange duality}
by Arnold,
which states that exceptional unimodal singularities come in pairs
in such a way that the Dolgachev number and the Gabrielov number are interchanged.
Strange duality is described in terms of an exchange of the algebraic lattice
and the transcendental lattice of a K3 surface
by Dolgachev and Nikulin \cite{Dolgachev_IQF,Nikulin_ISBF}
and Pinkham \cite{Pinkham_strange-duality},
and is now considered as a precursor of \emph{mirror symmetry}.
In this paper,
we consider the following generalization
of triangle singularities.
Let $n$ be an integer greater than 3 and
$\bp = (p_1, \ldots, p_n)$ be a sequence of integers
called a \emph{signature}.
In what follows, we assume that $\bp$ satisfies
\begin{align}
\sum_{i=1}^n \frac{1}{p_i} < 1.
\end{align}
Consider the ring
\begin{align}
T = \bC[X_1, \ldots, X_n] \left/ \lb \sum_{i=1}^n X_i^{p_i} \rb \right.,
\end{align}
which is graded by the abelian group
\begin{align} \label{eq:L}
L = \bZ \vecX_1 \oplus \cdots \oplus \bZ \vecX_n \oplus \bZ \vecc \left/ \lb p_i \vecX_i - \vecc \rb \right.
\end{align}
of rank one.
Here the grading is given by $X_i \in T_{\vecX_i}$ for all $i \in \{ 1, \ldots, n \}$.
The \emph{dualizing element} is defined by
\begin{align}
\vecomega = \vecc - \sum_{i=1}^n \vecX_i,
\end{align}
and the \emph{canonical ring} is defined by
\begin{align}
R = \bigoplus_{k=0}^\infty
R_k, \qquad R_k =
T_{k \vecomega}.
\end{align}
The singularity of $\Spec R$ is
a higher-dimensional generalization of triangle singularities.
The main result in this paper is the finiteness
of hypersurface singularities of this type:
\begin{theorem} \label{th:main}
For any integer $n$ greater than 3,
there are only finitely many signatures $\bp = (p_1,\ldots,p_n)$
such that $R$ is a hypersurface.
\end{theorem}
Our proof of \pref{th:main} gives an algorithm
to classify all hypersurface generalized triangle singularities
for any given $n \ge 4$.
The list of hypersurface generalized triangle singularities for $n=4$
is shown in \pref{tb:dim3}.
Our proof of \pref{th:main} also gives the following:
\begin{theorem} \label{th:isolated}
For any integer $n$ greater than 3,
the generalized triangle singularity
associated with signature $\bp = (p_1, \ldots, p_n)$
is an isolated hypersurface singularity
if and only if
\begin{align} \label{eq:fraction}
\sum_{i=1}^n \frac{1}{p_i} + \frac{1}{\prod_{i=1}^n p_i} = 1.
\end{align}
\end{theorem}
A slight generalization
\begin{align} \label{eq:fraction2}
\sum_{i=1}^n \frac{1}{p_i} + \frac{1}{\prod_{i=1}^n p_i} \in \bZ
\end{align}
of the Diophantine equation
\eqref{eq:fraction} is equivalent to the \emph{improper Zn\'am problem},
which asks for a sequence $(p_1, \ldots, p_n)$
of integers satisfying
$p_i \mid \prod_{j \in \{ 1, \ldots, n \} \setminus \{ i \}} p_j + 1$
for all $i \in \{ 1, \ldots, n \}$.
It is not known whether \eqref{eq:fraction2} implies \eqref{eq:fraction}.
The \emph{$a$-invariant} $a=a(R) \in \bZ$
of a graded Gorenstein ring $R$
is defined by
\begin{align}
K_R = R(a),
\end{align}
where $K_R$ is the graded canonical module of $R$.
If $R$ is a hypersurface
generated by $n$ elements of degrees $a_1, \ldots, a_n$
with one relation of degree $h$,
then the $a$-invariant of $R$ is given by
\begin{align}
a = h-a_1-\cdots-a_n.
\end{align}
\begin{theorem} \label{th:a}
If $R$ is a hypersurface generalized triangle singularity,
then one has
\begin{align} \label{eq:a}
a(R) = 1.
\end{align}
\end{theorem}
The equality \eqref{eq:a} also holds for
hypersurface (ordinary) triangle singularities (cf.\ \cite[Proposition 2.8]{MR586722}).
It follows from \eqref{eq:a} that one can compactify $\Spec R$
to a weighted projective hypersurface
with trivial canonical bundle
by adding one variable of degree one.
It is an interesting problem to study
when this hypersurface admits a smoothing.
\ \\
\emph{Acknowledgements}:
We thank Korea Institute for Advanced Study
for financial support and excellent research environment.
K.~U. is supported by JSPS Grant-in-Aid for Young Scientists No.~24740043.
\begin{table}[htbp]
\begin{align*}
\begin{array}{cccc}
\toprule
\text{Signature} & \text{Generators} & \text{Weights} & \text{Relation} \\
\midrule
( 2 , 3 , 7 , 43 ) & (X,Y,Z,W) & ( 903 , 602 , 258 , 42 ; 1806 ) & x^2+y^3+z^7+w^{43} \\
( 2 , 3 , 7 , 44 ) & (XW,Y,Z,W^2) & ( 483 , 308 , 132 , 42 ; 966 ) & x^2+w(y^3+z^7+w^{22}) \\
( 2 , 3 , 7 , 45 ) & (X,YW,Z,W^3) & ( 315 , 224 , 90 , 42 ; 672 ) & y^3+w(x^2+z^7+w^{15}) \\
( 2 , 3 , 7 , 49 ) & (X,Y,ZW,W^7) & ( 147 , 98 , 48 , 42 ; 336 ) & z^7+w(x^2+y^3+w^7) \\
( 2 , 3 , 8 , 25 ) & (XZ,Y,Z^2,W) & ( 375 , 200 , 150 , 24 ; 750 ) & x^2+z(y^3+z^4+w^{25}) \\
( 2 , 3 , 8 , 26 ) & (XZW,Y,Z^2,W^2) & ( 207 , 104 , 78 , 24 ; 414 ) & x^2+zw(y^3+z^4+w^{13}) \\
( 2 , 3 , 9 , 19 ) & (X,YZ,Z^3,W) & ( 171 , 152 , 114 , 18 ; 456 ) & y^3+z(x^2+z^3+w^{19}) \\
( 2 , 3 , 9 , 21 ) & (X,YZW,Z^3,W^3) & ( 63 , 62 , 42 , 18 ; 186 ) & y^3+zw(x^2+z^3+w^{7}) \\
( 2 , 3 , 10 , 16 ) & (XZW,Y,Z^2,W^2) & ( 159 , 80 , 48 , 30 ; 318 ) & x^2+zw(y^3+z^5+w^8) \\
( 2 , 3 , 13 , 13 ) & (W^{13},X,Y,ZW) & ( 78 , 39 , 26 , 12 ; 156 ) & w^{13}+x(x+y^2+z^3) \\
( 2 , 4 , 5 , 21 ) & (XY,Y^2,Z,W) & ( 315 , 210 , 84 , 20 ; 630 ) & x^{2}+y(y^2+z^5+w^{21}) \\
( 2 , 4 , 5 , 22 ) & (XYW,Y^2,Z,W^2) & ( 175 , 110 , 44 , 20 ; 350 ) & x^{2}+yw(y^2+z^5+w^{11}) \\
( 2 , 4 , 6 , 13 ) & (XYZ,Y^2,Z^2,W) & ( 143 , 78 , 52 , 12 ; 286 ) & x^2+yz(y^2+z^3+w^{13}) \\
( 2 , 4 , 6 , 14 ) & (XYZW,Y^2,Z^2,W^2) & ( 83 , 42 , 28 , 12 ; 166 ) & x^2+yzw(y^2+z^3+w^{7}) \\
( 2 , 4 , 7 , 10 ) & (XYW,Y^2,W^2,Z) & ( 119 , 70 , 28 , 20 ; 238 ) & x^2+yz(y^2+z^5+w^{7}) \\
( 2 , 5 , 5 , 11 ) & (Y^5,X,YZ,W) & ( 110 , 55 , 44 , 10 ; 220 ) & z^5+x(x+y^2+w^{11}) \\
( 2 , 5 , 5 , 15 ) & (Y^5,X,YZW,W^5) & ( 30 , 15 , 14 , 10 ; 70 ) & z^5+xw(x+y^2+w^{3}) \\
( 2 , 5 , 6 , 8 ) & (XZW,Z^2,W^2,Y) & ( 95 , 40 , 30 , 24 ; 190 ) & x^2+yz(y^3+z^4+w^{5}) \\
( 2 , 5 , 7 , 7 ) & (W^7,X,ZW,Y) & ( 70 , 35 , 20 , 14 ; 140 ) & z^7+x(x+y^2+w^{5}) \\
( 2 , 7 , 7 , 7 ) & (Z^7,Y^7,X,YZW) & ( 14 , 14 , 7 , 6 ; 42 ) & w^7+xy(x+y+z^2) \\
( 3 , 3 , 4 , 13 ) & (X^3,XY,Z,W) & ( 156 , 104 , 39 , 12 ; 312 ) & y^3+x(x+z^4+w^{13}) \\
( 3 , 3 , 4 , 15 ) & (X^3,XYW,Z,W^3) & ( 60 , 44 , 15 , 12 ; 132 ) & y^3+xw(x+z^4+w^{5}) \\
( 3 , 3 , 5 , 8 ) & (X^3,XY,Z,W) & ( 120 , 80 , 24 , 15 ; 240 ) & y^3+x(x+z^5+w^{8}) \\
( 3 , 3 , 5 , 9 ) & (X^3,XYW,W^3,Z) & ( 45 , 35 , 15 , 9 ; 105 ) & y^3+xz(x+z^3+w^{5}) \\
( 3 , 3 , 6 , 7 ) & (X^3,XYZ,Z^3,W) & ( 42 , 35 , 21 , 6 ; 105 ) & y^3+xz(x+z^2+w^{7}) \\
( 3 , 3 , 6 , 9 ) & (X^3,XYZW,Z^3,W^3) & ( 18 , 17 , 9 , 6 ; 51 ) & y^3+xzw(x+z^2+w^{3}) \\
( 3 , 4 , 4 , 8 ) & (Z^4,YZW,W^4,X) & ( 24 , 15 , 12 , 8 ; 60 ) & y^4+xz(x+z^2+w^{3}) \\
( 3 , 4 , 5 , 5 ) & (W^5,ZW,X,Y) & ( 60 , 24 , 20 , 15 ; 120 ) & y^5+x(x+z^3+w^{4}) \\
( 3 , 5 , 5 , 5 ) & (W^5,Z^5,YZW,X) & ( 15 , 15 , 9 , 5 ; 45 ) & z^5+xy(x+y+w^3) \\
( 4 , 4 , 4 , 5 ) & (Y^4,Z^4,XYZ,W) & ( 20 , 20 , 15 , 4 ; 60 ) & z^4+xy(x+y+w^5) \\
( 4 , 4 , 4 , 8 ) & (X^4,Y^4,XYZW,W^4) & ( 8 , 8 , 7 , 4 ; 28 ) & z^4+xyw(x+y+w^2) \\
( 5 , 5 , 5 , 5 ) & (X^5,Y^5,Z^5,XYZW) & ( 5 , 5 , 5 , 4 ; 20 ) & w^5+xyz(x+y+z) \\
\bottomrule
\end{array}
\end{align*}
\caption{Hypersurface generalized triangle singularities in dimension 3}
\label{tb:dim3}
\end{table}
\clearpage
\begin{comment}
\begin{shaded}
In this section,
we give a proof of \pref{th:Dolgachev}
as a warm-up exercise.
The `if' part is proved by a straightforward
case-by-case analysis.
For example,
if $(p,q,r)=(4,4,4)$,
then
the ring $R$ is generated by three elements
\begin{align}
x = XYZ, \quad
y = Y^4, \quad
z = Z^4
\end{align}
of degrees 3, 4, and 4 respectively,
with one relation
\begin{align}
x^4+y^2z+yz^2
= X^4Y^4Z^4+Y^8Z^4+Y^4Z^8
= X^4Y^4Z^4(X^4+Y^4+Z^4)
= 0.
\end{align}
To prove `only if' part,
we may assume $p \le q \le r$.
\begin{enumerate}[(i)]
\item
If $p \ge 4$, then one has
\begin{align}
\begin{array}{c|ccccc}
k & 0 & 1 & 2 & 3 & 4 \\ \hline
\dim R_k & 1 & 0 & 0 & 1 & 2
\end{array},
\end{align}
so that one needs
\begin{align}
\dim R_5 = 0
\end{align}
for $R$ to be a hypersurface,
since $R$ will be generated by more than three elements
otherwise.
Since we have $\dim R_5 \ge 1$
if $r \ge 5$,
the only possibility is $p=q=r=4$.
\item
Assume $p=3$.
\begin{enumerate}
\item
If $r \le 4$,
then one has $(p,q,r) = (3,3,4)$ or $(3,4,4)$,
and one can show that $R$ is a hypersurface in both cases.
\item
If $r \ge 5$, then one has
\begin{align}
\dim R_5 &=
\begin{cases}
1 & q \le 4, \\
2 & q \ge 5,
\end{cases}
\end{align}
so that one needs $q \le 4$
for $R$ to be a hypersurface.
This gives $(p,q,r) = (3,4,4)$ or $(3,4,5)$,
and one can show that $R$ is a hypersurface
in both cases.
\end{enumerate}
\item
If $p=2$, then one has
\begin{align}
4 \vecomega &= (q-4) \vecy + (r-4) \vecz, \\
5 \vecomega &= \vecx + (q-5) \vecy + (r-5) \vecz, \\
6 \vecomega &= \vecc + (q-6) \vecy + (r-6) \vecz.
\end{align}
\begin{enumerate}
\item
If $q \ge 6$, then one has
\begin{align}
\begin{array}{c|ccccccc}
k & 0 & 1 & 2 & 3 & 4 & 5 & 6\\ \hline
\dim R_k & 1 & 0 & 0 & 0 & 1 & 1 & 2
\end{array},
\end{align}
so that $R$ cannot be a hypersurface.
\item
If $q=5$, then one has
\begin{align}
7 \vecomega = \vecx + 3 \vecy + (r-7) \vecz,
\end{align}
so that one needs $r \le 6$
for $R$ to be a hypersurface.
Then one has $(p,q,r) = (2,5,5)$ or $(2,5,6)$,
and $R$ is a hypersurface in both cases.
\item
If $q=4$, then one has
\begin{align}
8 \vecomega = \vecc + (r-8) \vecz.
\end{align}
If $r \ge 8$,
then one has
\begin{align}
\begin{array}{c|ccccccccc}
k & 0 & 1 & 2 & 3 & 4 & 5 & 6 & 7 & 8 \\ \hline
\dim R_k & 1 & 0 & 0 & 0 & 1 & 0 & 1 & 0 & 2
\end{array},
\end{align}
so that $R$ cannot be a hypersurface.
If $r \le 7$,
then one has $(p,q,r)=(2,4,4)$, $(2,4,5)$, $(2,4,6)$, $(2,4,7)$,
and $R$ is a hypersurface unless $(p,q,r) = (2,4,4)$.
\item
If $q=3$, then one has
\begin{align}
6 \vecomega &= (r-6) \vecz, \\
9 \vecomega &= \vecx + (r-9) \vecz, \\
10 \vecomega &= 2 \vecy + (r-10) \vecz, \\
12 \vecomega &= \vecc + (r - 12) \vecz.
\end{align}
If $r \ge 12$, then one has
\begin{align}
\begin{array}{c|ccccccccccccc}
k & 0 & 1 & 2 & 3 & 4 & 5 & 6 & 7 & 8 &9 & 10 & 11 & 12 \\ \hline
\dim R_k & 1 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 1 & 1 & 0 & 2
\end{array},
\end{align}
so that $R$ cannot be a hypersurface.
If $r \le 11$, then a case-by-case analysis shows that
$R$ is a hypersurface if and only if
$(p,q,r) = (2,3,7)$, $(2,3,8)$ or $(2,3,9)$.
\end{enumerate}
\end{enumerate}
\end{shaded}
\end{comment}
\section{Proof of \pref{th:main}}
\label{sc:main}
Let $\bp = (p_1, \ldots, p_n)$ be a signature
such that $R$ is a hypersurface, and
\begin{align}
\Tbar
= \bigoplus_{\vecv \in L} \Tbar_\vecv
= \bC \ld \Xbar_1,\ldots,\Xbar_n \rd
\end{align}
be a polynomial ring
graded by the abelian group $L$
in \eqref{eq:L}.
The Veronese subring of $\Tbar$
over $\bZ \, \vecomega$ is denoted by
\begin{align}
\Rbar = \bigoplus_{k \in \bZ} \Rbar_k, \qquad
\Rbar_k = \Tbar_{k \vecomega}.
\end{align}
We write
\begin{align}
[i,j]=\{i,i+1,\ldots,j\}
\end{align}
for a pair $(i,j)$ of integers with $i \le j$.
Let
$
\varphi \colon \Tbar \to T = \Tbar/ \lb \Fbar \rb
$
be the natural projection,
where
\begin{align}
\Fbar
= \sum_{i=1}^n \Ybar_i,
\qquad
\Ybar_i = \Xbar_i^{p_i}
\text{ for } i \in [1,n].
\end{align}
The restriction
$
\varphi|_\Rbar \colon \Rbar \to R = \Rbar / \lb \Fbar \rb
$
will also be denoted by $\varphi$
by abuse of notation.
We write $X_i=\varphi(\Xbar_i)$ for $i \in [1,n]$.
Any element $\vecv \in L$ can be written uniquely as
\begin{align} \label{eq:ka}
\vecv = \ell \vecc + \sum_{i=1}^n a_i \vecX_i,
\end{align}
where $\ell \in \bZ$ and
$0 \le a_i \le p_i-1$ for any $i \in [1,n]$.
This defines functions $\ell \colon L \to \bZ$
and $a_i \colon L \to [0,p_i-1]$
for $i \in [1,n]$.
Any element of $\Tbar_\vecv$
for $\vecv \in L$ can be written as the product
$
\Mbar(\vecv) \Pbar
$, where $\Mbar(\vecv)$ is the monomial defined by
\begin{align} \label{eq:Mbar}
\Mbar(\vecv) = \prod_{i=1}^n \Xbar_i^{a_i(\vecv)}
\end{align}
and
$
\Pbar \in \bC \ld \Ybar_1,\ldots,\Ybar_n \rd_\ell
$
is a homogeneous polynomial of degree $\ell$.
\begin{comment}
\begin{shaded}
Since $\{ Y_i \}_{i=1}^n$ has one relation
\begin{align}
\sum_{i=1}^n Y_i = 0,
\end{align}
we have
\begin{align}
\dim R_k = \binom{\ell(k \vecomega)+n-1}{n-1}
\end{align}
where
\begin{align}
\ell(k \vecomega)
= k - \sum_{i=1}^n \myceil{\frac{k}{p_i}}.
\end{align}
We also set
$
M = \varphi \circ \Mbar \colon L \to T.
$
\end{shaded}
\end{comment}
Let
$
\frakm
= \bigoplus_{k=1}^\infty R_{k}
$
be the irrelevant ideal of $R$.
Since we assume that $R$ is a hypersurface,
that is, $\dim_\bC \frakm/\frakm^2 = n$,
we can choose a set
$
\Xibar = \lc \xbar_i \rc_{i=1}^n \subset \Rbar
$
of monomials
whose image
$
\Xi = \varphi \lb \Xibar \rb
= \lc x_i = \varphi \lb \xbar_i \rb \rc_{i=1}^n
$
forms a basis of the vector space
$\frakm/\frakm^2$.
By the following \pref{lm:embdim}, the ring $R$ is generated by $\Xi$ over $\bC$.
\begin{lemma} \label{lm:embdim}
Let $S$ be a subset of $\frakm$ consisting of homogeneous elements,
that is, $S \subset \bigcup_{k=1}^\infty R_k$.
Then $S$ generates the ring $R$ over $\bC$ if and only if
the image of $S$ spans the vector space $\frakm / \frakm^2$.
\end{lemma}
\begin{proof}
Let $\bC S$ denote the vector space spanned by $S$.
If $S$ generates $R$, that is, $\bC[S]=R$, then
$\frakm/\frakm^2=(\bC[S]\cap \frakm)/\frakm^2=(\bC S+\frakm^2)/\frakm^2$.
Hence the image of $S$ spans $\frakm/\frakm^2$.
Conversely, assume that the image of $S$ spans $\frakm/\frakm^2$.
In order to prove that $S$ generates $R$ by induction,
suppose that $\bC[S]$ contains $R_k$ for all $k<n$.
Then $\bC[S] \supset \frakm^2 \cap R_n$.
Since the image of $S$ spans $\frakm/\frakm^2$, we have
$R_n=\bC S \cap R_n + \frakm^2 \cap R_n$,
which implies that $\bC[S]$ contains $R_n$.
Therefore we obtain $\bC[S]=R$ by induction.
\end{proof}
We also set
\begin{align} \label{eq:df}
\begin{split}
\nu=1-\sum_{i=1}^n \frac{1}{p_i}, \quad
N= \lcm \lc p_i \relmid i \in [1,n] \rc, \\
N_i=\lcm \lc p_j \relmid j \in [1,n] \setminus \{ i \} \rc, \quad
q_i=\nu p_i N_i.
\end{split}
\end{align}
\begin{lemma} \label{lm:power}
For any $i \in [1,n]$ and any $m \in \bN$,
we have $\Xbar_i^m \in \Rbar$ if and only if $q_i \mid m$.
\end{lemma}
\begin{proof}
Note that $\Rbar_k$ contains a pure power of $\Xbar_i$
if and only if
\begin{align} \label{eq:a=0}
a_j(k \vecomega) = 0
\qquad \text{for any } j \in [1,n] \setminus \{i\},
\end{align}
where $a_j \colon L \to [0,p_j-1]$ are defined
by \eqref{eq:ka}.
Since
\begin{align}
a_j(k \vecomega) = p_j \myceil{\frac{k}{p_j}} - k
\qquad \text{for any } j \in [1,n],
\end{align}
the condition \eqref{eq:a=0} holds
if and only if $k$ is an integer multiple of $N_i$.
It follows from
\begin{align}
N_i \vecomega
=N_i \vecc -
\sum_{j \in [1,n] \setminus \{ i \}} \frac{N_i}{p_j} \vecc
- N_i \vecX_i
&= \lb \nu + \frac{1}{p_i} \rb N_i \vecc - N_i \vecX_i
= q_i \vecX_i
\end{align}
that the pure power of $\Xbar_i$ contained in $\Rbar_{N_i}$ is $\Xbar_i^{q_i}$, and
\pref{lm:power} is proved.
\end{proof}
\begin{corollary} \label{cr:puregen}
If $\Xbar_i^m \in \Xibar$ for some $i \in [1,n]$ and $m \in \bN$,
then one has $m=q_i$.
\end{corollary}
\begin{proof}
This is immediate from \pref{lm:power}.
\end{proof}
\begin{lemma} \label{lm:MPQ}
For any $k \in \bN$
and any $\Zbar \in \Rbar_k$,
there exist
$
\Pbar
\in \bC \ld \Ybar_1,\ldots,\Ybar_n \rd_{\ell(k \vecomega)-1}
$
and
$
\Qbar \in \bC \ld \xbar_1,\ldots,\xbar_n \rd \cap \Rbar_k
$
such that
$
\Zbar = \Mbar(k \vecomega) \Pbar \Fbar + \Qbar.
$
\end{lemma}
\begin{proof}
Since $\Xi$ generates $R$ as a ring,
there exist $\Pbar' \in \Tbar_{k \vecomega - \vecc}$
and $\Qbar \in \bC \ld \xbar_1, \ldots, \xbar_n \rd$
such that
$
\Zbar = \Pbar' \Fbar + \Qbar.
$
It follows from the definition of $\ell$ and $\Mbar$ that
$
\ell(k \vecomega-\vecc) = \ell(k \vecomega)-1
$
and
$
\Mbar(k \vecomega-\vecc) = \Mbar(k \vecomega).
$
Hence there exists
$
\Pbar \in \bC \ld \Ybar_1,\ldots,\Ybar_n \rd_{\ell(k \vecomega)-1}
$
such that
$
\Pbar' = \Mbar(k \vecomega) \Pbar.
$
\end{proof}
\begin{lemma} \label{lm:R=T}
If $\xbar_i=\Xbar_i^{q_i}$ for all $i \in [1,n]$,
then we have $R=T$ and $\xbar_i=\Xbar_i$ for all $i \in [1,n]$.
\end{lemma}
\begin{proof}
For each $i$, fix a sufficiently large $k$
with $k\equiv -1 \bmod p_i$.
Since
\begin{align}
k \vecomega = k \vecc - k \sum_{i=1}^n \vecX_i,
\end{align}
we have
$
a_i(k \vecomega) = 1,
$
so that there exists a monomial $\Gbar \in \Tbar_{k \vecomega - \vecX_i}$
such that $\Xbar_i \Gbar \in \Rbar_k$ and $\Xbar_i \nmid \Gbar$.
By applying \pref{lm:MPQ}, we have
\begin{align} \label{eq:XG2}
\Xbar_i \Gbar = \Mbar(k\vecomega) \Pbar \Fbar + \Qbar, \quad
\Xbar_i \mid \Mbar(k\vecomega), \quad \text{and} \quad \Xbar_i^2 \nmid \Mbar(k\vecomega).
\end{align}
Assume for contradiction that $q_i>1$.
By comparing terms of degree $1$ in the variable $X_i$
in \eqref{eq:XG2}, we obtain
\begin{align} \label{eq:XG}
\Xbar_i \Gbar
= \Mbar(k\vecomega) \cdot \Pbar \big|_{\Xbar_i=0} \cdot \Fbar \big|_{\Xbar_i=0}.
\end{align}
Since
$
\Fbar \big|_{\Xbar_i=0}
= \sum_{j \in [1,n] \setminus \{ i \}} \Xbar_j^{p_j}
$
is not a monomial,
the right hand side of \eqref{eq:XG} is not a monomial.
This contradicts the fact that the left hand side is a monomial,
and \pref{lm:R=T} is proved.
\end{proof}
\begin{lemma} \label{lm:XiXj}
If there exist $i, j \in [1,n]$ such that
$i \ne j$, $\Xbar_i^{q_i} \nin \Xibar$, and
$\Xbar_i^a \Xbar_j^b \nin \Xibar$ for all $a,b \ge 1$,
then $\Xbar_j^{q_j} \in \Xibar$ and $p_i \mid q_i$.
\end{lemma}
\begin{proof}
By \pref{lm:power},
we have $\Xbar_i^{q_i} \in R$.
Hence we have
\begin{align} \label{eq:power_x_k}
\Xbar_i^{q_i} = \Mbar \lb q_i \vecX_i \rb \Pbar \Fbar + \Qbar
\end{align}
by \pref{lm:MPQ}.
Let
$
\pi \colon \bC \ld \Xbar_1, \ldots, \Xbar_n \rd
\to \bC \ld \Xbar_i, \Xbar_j \rd
$
be the surjective ring homomorphism
defined by
\begin{align}
\pi \lb \Xbar_k \rb =
\begin{cases}
\Xbar_k & k = i, j, \\
0 & \text{otherwise}.
\end{cases}
\end{align}
By projecting \eqref{eq:power_x_k} by $\pi$,
we obtain
\begin{align} \label{eq:Xq}
\Xbar_i^{q_i}
= \pi \lb M \lb q_i \vecX_i \rb \Pbar \rb \cdot \lb \Xbar_i^{p_i}
+ \Xbar_j^{p_j} \rb + \pi \lb \Qbar \rb.
\end{align}
It follows from the assumption of \pref{lm:XiXj}
that the only element in $\Xibar$
whose image by $\pi$ does not vanish
is a polynomial in $\Xbar_j$.
Hence we have $\pi \lb \Qbar \rb \in \bC \ld \Xbar_j \rd$.
If $\pi \lb \Qbar \rb = 0$,
then the right hand side of \eqref{eq:Xq} is not a monomial,
which contradicts the fact that the left hand side is a monomial.
Hence we have $\pi \lb \Qbar \rb \ne 0$,
so that $\Xbar_j^m \in \Xibar$ for some $m \in \bN$.
This implies $m=q_j$ by \pref{cr:puregen}.
It follows from \eqref{eq:Xq} that
$\Xbar_i^{q_i}-\pi \lb \Qbar \rb$ is divisible
by $\pi \lb \Mbar(q_i \vecX_i) \rb$.
Together with the fact that $\pi \lb \Qbar \rb \in \bC \ld \Xbar_j \rd$,
this implies that $\Mbar \lb q_i \vecX_i \rb = 1$.
Hence $p_i$ divides $q_i$, and
\pref{lm:XiXj} is proved.
\end{proof}
\begin{lemma} \label{lm:XiXjXk}
Let $i,j,k$ be distinct elements of $[1,n]$.
If $\Xbar_i^{q_i}, \Xbar_j^{q_j} \nin \Xibar$ and $\Xbar_k^{q_k} \in \Xibar$,
then there exists an element in $\Xibar$
of the form $\Xbar_i^a \Xbar_j^b \Xbar_k^c$
with $(a,b) \ne (0,0)$ and $c \ge 1$.
\end{lemma}
\begin{proof}
Assume for contradiction that $\Xbar_i^a \Xbar_j^b \Xbar_k^c \nin \Xibar$
for all $(a,b,c)$ with $(a,b)\ne (0,0)$ and $c \ge 1$.
In particular, we have $\Xbar_i^a \Xbar_k^c \nin \Xibar$ for all $a, c \ge 1$.
Together with the assumption that $\Xbar_i^{q_i} \nin \Xibar$,
this implies $p_i \mid q_i$ by \pref{lm:XiXj}.
If we set $q=q_i/p_i$,
then we have
\begin{align}
\Ybar_i^{q} = \Pbar \Fbar + \Qbar
\end{align}
by \pref{lm:MPQ},
since $\Ybar_i^q \in T_{q \vecc}$ and $\Mbar(q \vecc) = 1$.
Let
$
\pi \colon \bC \ld \Xbar_1, \ldots, \Xbar_n \rd
\to \bC \ld \Xbar_i, \Xbar_j, \Xbar_k \rd
$
be the surjective homomorphism
defined by
\begin{align}
\pi(\Xbar_l) =
\begin{cases}
\Xbar_l & l = i, j, k, \\
0 & \text{otherwise}.
\end{cases}
\end{align}
We can write
\begin{align}
\pi \lb \Pbar \rb
= \Ybar_i^{q-1} + \Abar_1 \Ybar_i^{q-2} + \cdots + \Abar_{q-1},
\end{align}
where
$
\Abar_l \in \bC \ld \Ybar_j,\Ybar_k \rd
$
for $l \in [1,q-1]$.
Then we have
\begin{align*}
\pi \lb \Pbar \Fbar \rb
&= \pi \lb \Pbar \rb \pi \lb \Fbar \rb \\
&= \lb \Ybar_i^{q-1} + \Abar_1 \Ybar_i^{q-2} + \cdots + \Abar_{q-1} \rb
\lb \Ybar_i+\Ybar_j+\Ybar_k \rb \\
&= \Ybar_i^q + \lb \Abar_1+\Ybar_j+\Ybar_k \rb \Ybar_i^{q-1}
+\lb \Abar_2 + \lb \Ybar_j+\Ybar_k \rb \Abar_1 \rb \Ybar_i^{q-2}
+\cdots+ \lb \Ybar_j+\Ybar_k \rb \Abar_{q-1}.
\end{align*}
We shall show that $\Xbar_j \mid \Abar_l$ for all $l \in [1, q-1]$.
It follows from the assumption
that every monomial appearing in $\pi(\Qbar)$
is either divisible by $\Xbar_i \Xbar_j$ or
consists only of $\Xbar_k$.
Since all monomials appearing in
$
\lb \Ybar_j+\Ybar_k \rb \Abar_{q-1}
$
are not divisible by $\Xbar_i$,
they must be in $\bC[\Xbar_k]$.
This implies that $\Abar_{q-1}=0$.
Since all monomials appearing in
$
\lb \Abar_{q-1} + \lb \Ybar_j + \Ybar_k \rb \Abar_{q-2} \rb \Ybar_i
$
contains $\Xbar_i$,
they must be divisible by $\Xbar_i \Xbar_j$.
Hence we must have $\Xbar_j \mid \Abar_{q-2}$.
By repeating the same argument,
we obtain $\Xbar_j \mid \Abar_l$ for all $l \in [1, q-1]$.
In particular, one has $\Xbar_j \mid \Abar_1$.
It follows that the monomial $\Ybar_i^{q-1} \Ybar_k$
from $(\Abar_1+\Ybar_j+\Ybar_k) \Ybar_i^{q-1}$ do not cancel with any other terms.
Since this monomial is neither divisible by $\Xbar_i \Xbar_j$
nor consists only of $\Xbar_k$,
this is a contradiction,
and \pref{lm:XiXjXk} is proved.
\end{proof}
The following lemma is the key to proving \pref{th:main}.
Set
\begin{align} \label{eq:I1}
I := \lc i \in [1,n] \relmid \Xbar_i^{q_i} \in \Xibar \rc.
\end{align}
\begin{lemma} \label{lm:gen}
If $n \ge 4$,
then we have $|I| \ge n-1$.
\end{lemma}
\begin{proof}
Assume for contradiction
that $n \ge 4$ and $r :=n - |I| \ge 2$.
If $i \ne j$ and $i, j \nin I$,
then we have $\Xbar_i^a \Xbar_j^b \in \Xibar$ for some $a, b \ge 1$
by \pref{lm:XiXj}.
It follows that
\begin{align}
\# \lc \Xbar_i^a \Xbar_j^b \in \Xibar \relmid i,j \nin I, \
a,b \ge 1 \rc
\ge \binom{r}{2}.
\end{align}
Similarly, \pref{lm:XiXjXk} implies that
\begin{align} \label{eq:I3}
\# \lc \Xbar_i^a \Xbar_j^b \Xbar_k^c \in \Xibar \relmid
i,j \nin I, \
k \in I, \
(a,b) \ne (0,0), \
c \ge 1
\rc
\ge |I| = n - r.
\end{align}
It follows from \pref{eq:I1}--\pref{eq:I3} that
\begin{align}
\# \Xibar
= n
\ge |I|+\binom{r}{2}+|I|
= (n-r) + \frac{1}{2}r(r-1)+(n-r),
\end{align}
and hence
\begin{align} \label{eq:nr}
n \le - \frac{1}{2}r(r-5).
\end{align}
Since
\begin{align}
\max \lc -\frac{1}{2}r(r-5) \relmid r \in [0,n] \rc
= 3,
\end{align}
the inequality \eqref{eq:nr} contradicts the assumption $n \ge 4$,
and \pref{lm:gen} is proved.
\end{proof}
\begin{proposition} \label{pr:nu_N}
If $n\geq 4$, we have $\nu=1/N$.
\end{proposition}
\begin{proof}
If $R=T$, we have $q_i=1$ for all $i \in [1,n]$ by \pref{lm:power}.
Hence
\begin{align}
q_i
=\nu p_i N_i
=\nu \cdot \lcm(p_i,N_i) \cdot \gcd(p_i,N_i)
=\nu N \cdot \gcd(p_i, N_i)
=1.
\end{align}
Since $\nu N$ is an integer,
it follows that
\begin{align} \label{eq:nuN}
\nu N
= \gcd (p_i, N_i)
=1.
\end{align}
If $R \subsetneq T$,
then we have $\Xbar_i^{q_i} \nin \Xibar$
for some $i \in [1,n]$ by \pref{lm:R=T}.
Since we have $\Xbar_l^{q_l}\in \Xibar$ for all $l \in [1,n] \setminus \{ i \}$
by \pref{lm:gen},
we can set $\xbar_l = \Xbar_l^{q_l}$ for $l \in [1,n] \setminus \{ i \}$.
Then we have $\xbar_i \ne \Xbar_i^{q_i}$.
Let $m$ be an element of $[1,n] \setminus \{ i \}$
such that the variable $\Xbar_m$ appears in $\xbar_i$,
and fix any distinct elements $j$ and $k$ of $[1,n] \setminus \{ i, m \}$.
Then it follows that $\Xbar_i^a \Xbar_j^b\nin \Xibar$ for all $a,b \ge 1$.
This implies that $p_i \mid q_i$ by \pref{lm:XiXj}.
Let
$
\pi \colon \bC \ld \Xbar_1, \ldots, \Xbar_n \rd
\to \bC \ld \Xbar_i, \Xbar_j, \Xbar_k \rd
$
be the surjective homomorphism
defined by
\begin{align}
\pi \lb \Xbar_l \rb =
\begin{cases}
\Xbar_l & l = i, j, k, \\
0 & \text{otherwise},
\end{cases}
\end{align}
just as in the proof of \pref{lm:XiXjXk}.
It follows from the choice of $j$ and $k$ that
\begin{align} \label{eq:pixi}
\pi \lb \xbar_l \rb = 0 \text{ for any } l \in [1,n] \setminus \{ j, k \}.
\end{align}
If we write $q=q_i/p_i$,
then the same argument as in the proof of \pref{lm:XiXjXk}
shows
\begin{align} \label{eq:piQ}
\Ybar_i^q = \Pbar \Fbar + \Qbar
\end{align}
and
\begin{align} \label{eq:PF}
\pi(\Pbar \Fbar)
&= \Ybar_i^q + \lb \Abar_1 + \Ybar_j + \Ybar_k \rb \Ybar_i^{q-1}
+ \lb \Abar_2 + \lb \Ybar_j + \Ybar_k \rb \Abar_1 \rb \Ybar_i^{q-2}
+ \cdots + \lb \Ybar_j + \Ybar_k \rb \Abar_{q-1}.
\end{align}
If follows from \eqref{eq:pixi} that
\begin{align} \label{eq:piQ2}
\pi \lb \Qbar \rb \in \bC \ld \Xbar_j^{q_j},\Xbar_k^{q_k} \rd.
\end{align}
By projecting \eqref{eq:piQ} by $\pi$,
we obtain
\begin{align}
\pi \lb \Qbar \rb=\Ybar_i^q-\pi \lb \Pbar \Fbar \rb,
\end{align}
which together with \eqref{eq:PF} and \eqref{eq:piQ2} gives
\begin{align}
\Abar_l=(-\Ybar_j-\Ybar_k)^l \quad \text{for any } l \in [1,q-1]
\end{align}
and
\begin{align} \label{eq:piQ3}
\pi(Q)=(-\Ybar_j-\Ybar_k)^q.
\end{align}
It follows from \eqref{eq:piQ2} and \eqref{eq:piQ3}
that
$
\Ybar_j^a \Ybar_k^{q-a} \in \bC \ld \Xbar_j^{q_j},\Xbar_k^{q_k} \rd
$
for any $a \in [0,q]$.
Since
$
\Ybar_j^a \Ybar_k^{q-a}
= \Xbar_j^{a p_j} \Xbar_k^{(q-a) p_k},
$
this implies $q_j \mid p_j$ and $q_k \mid p_k$.
Hence both $p_j/q_j=1/\nu N_j$ and
$p_k/q_k=1/\nu N_k$ are integers.
The product $u := \nu N$ is an integer
by the definitions of $\nu$ and $N$ in \eqref{eq:df}.
The definitions of $N_j$, $N_k$ and $N$
in \eqref{eq:df} and
the fact that both $1/\nu N_j=N/u N_j$ and
$1/\nu N_k=N/u N_k$ are integers
imply $u=1$,
and \pref{pr:nu_N} is proved.
\end{proof}
\begin{corollary} \label{cr:piqi}
We have $q_i=\gcd(p_i,N_i)$ and, in particular, $q_i \mid p_i$ for any $i \in [1,n]$.
\end{corollary}
\begin{proof}
It follows from \eqref{eq:df} and $\nu = 1/N$ that
\begin{align}
q_i
=\nu p_i N_i
=\nu \cdot \lcm(p_i,N_i) \cdot \gcd(p_i,N_i)
=\nu N \cdot \gcd(p_i, N_i)
=\gcd(p_i, N_i).
\end{align}
\end{proof}
\begin{corollary} \label{cr:gen}
We have either
\begin{enumerate}[(1)]
\item \label{it:1}
$R=T$ and $x_i = X_i$ for all $i \in [1,n]$, or
\item \label{it:2}
$R \ne T$ and
there exists $i \in [1,n]$ such that $q_i=p_i$ and
$
\displaystyle{
x_j =
\begin{cases}
X_j^{q_j} & j \ne i, \\
\prod_{q_k \ne 1} X_k & j = i.
\end{cases}
}
$
\end{enumerate}
\end{corollary}
\begin{proof}
It follows from \eqref{eq:df} and \pref{pr:nu_N} that
\begin{align}
\frac{1}{N} = 1 - \sum_{i=1}^n \frac{1}{p_i}.
\end{align}
Hence we have
\begin{align}
(N-1) \vecomega
&= (N-1) \vecc - (N-1) \sum_{i=1}^n \vecX_i \\
&= \lc (N-1)-\sum_{i=1}^n \frac{N}{p_i} \rc \vecc + \sum_{i=1}^n \vecX_i \\
&= \sum_{i=1}^n \vecX_i,
\end{align}
so that
$
\prod_{i=1}^n X_i \in R_{N-1}.
$
If $R=T$,
then we can set $x_i = X_i$ for all $i \in [1,n]$.
If $R \ne T$,
then we have $|I|=n-1$ by \pref{lm:gen},
and there exists $i \in [1,n]$
such that $x_j = X_j^{q_j}$ for $j \in [1,n] \setminus \{ i \}$.
Note that we have $\Fbar\in \Tbar_\vecc$ and $\vecc=N \vecomega$.
We can remove $X_k$ such that $q_k=1$
from $\prod_{k=1}^n X_k$
to obtain an element
$
\prod_{q_k \ne 1} X_k,
$
which is one of the generators of $R$.
Hence $x_i=\prod_{q_k \ne 1} X_k$.
Since we have $X_i^{q_i}\in R$ (\pref{lm:power}) and $q_i \mid p_i$ (\pref{cr:piqi}), it follows that $q_i=p_i$.
\end{proof}
\begin{lemma} \label{lm:c}
For any positive integer $n$,
there exist only finitely many sequences
$(p_1, \ldots, p_n)$
of $n$ positive integers
satisfying
\begin{align}
\sum_{i=1}^n \frac{1}{p_i}+\frac{1}{N}=1,
\end{align}
where
$
N = \lcm \{ p_i \mid i \in [1,n] \}.
$
\end{lemma}
\begin{proof}
We may assume $p_1\le p_2 \le \cdots \le p_n$.
Then we have $p_1 \le N$ and
\begin{align}
\frac{n+1}{p_1} \ge \sum_{i=1}^n \frac{1}{p_i}+\frac{1}{N}=1,
\end{align}
so that
\begin{align}
p_1 \le n+1.
\end{align}
Hence there are only finitely many possibilities for $p_1$.
If we fix $p_1$,
then we have
\begin{align}
\frac{n}{p_2} \ge \sum_{i=2}^n \frac{1}{p_i}+\frac{1}{N}=1 - \frac{1}{p_1}
\end{align}
since $p_2 \le N$.
This leaves only finitely many possibilities for $p_2$.
By repeating the same argument,
we can see that there are only finitely many possibilities
for $(p_1, \ldots, p_n)$.
\end{proof}
\begin{proof}[Proof of \pref{th:main}]
For any $n \ge 4$,
\pref{pr:nu_N} and \pref{lm:c}
give a finite list of possible signatures
for hypersurface generalized triangle singularities, and
\pref{th:main} is proved.
\end{proof}
For each signature $\bp=(p_1, \ldots, p_n)$
with $\nu=1/N$,
we can check if $R$ is a hypersurface as follows:
If $q_i=1$ for any $i \in [1,n]$,
then we have $R = T$ and $R$ is a hypersurface.
\begin{comment}
\begin{shaded}
Even if $q_i \ne 1$ for some $i \in [1,n]$,
\pref{cr:piqi} shows that
we need $q_i | p_i$ for any $i \in [1,n]$
in order for $R$ to be a hypersurface.
[$\leftarrow$ $q_i\mid p_i$ follows from $\nu=1/N$]
\end{shaded}
\end{comment}
If $R$ is a hypersurface and $R \ne T$,
then there exists $i \in [1, n]$ such that $q_i=p_i$ and
\begin{align} \label{eq:xj}
x_j =
\begin{cases}
X_j^{q_j} & j \ne i, \\
\prod_{q_k \ne 1} X_k & j = i
\end{cases}
\end{align}
by \pref{cr:gen}.
\begin{comment}
\begin{shaded}
For this to be the case,
we need $q_i=p_i$,
so that $X_i^{q_i}$ can be written in terms of
$X_j^{q_j}$ for $j \in [1,n] \setminus \{ i \}$.
\end{shaded}
\end{comment}
Assume that there exists $i \in [1,n]$ such that $q_i=p_i$.
Fix any such $i$ and define $\{ x_j \}_{j=1}^n$ by \eqref{eq:xj}.
Let $R'$ be the subring of $T$
generated by $\{ x_j \}_{j=1}^n$,
so that $R$ is a hypersurface
if and only if $R = R'$.
It follows from $\nu=1/N$ that $q_j \mid p_j$ for any $j \in [1,n]$ (cf.\ \pref{cr:piqi}).
Hence $Y_j := X_j^{p_j}$ is contained in $R'$ for any $j \in [1,n]$.
Note that any element of $T_{\vecv}$ for $\vecv \in L$
can be written as the product $M(\vecv) P$,
where $M(\vecv):=\varphi(\Mbar(\vecv))$ is the image of $\Mbar(\vecv)$ defined by \eqref{eq:Mbar},
and $P$ is a homogeneous element of $\bC[Y_1, \ldots, Y_n]/(Y_1+\cdots+Y_n)$.
Since $M((k+N) \vecomega) = M(k \vecomega)$ for any $k \in \bZ$,
the ring $R$ is generated by $\{ Y_j \}_{j=1}^{n}$ and $\{ M(k\vecomega) \}_{k=0}^{N-1}$.
Therefore we have $R = R'$
if and only if $M(k \vecomega) \in R'$ for $0 \leq k \leq N-1$.
\pref{tb:dim3} is obtained in this way.
\section{Proof of \pref{th:isolated}}
\label{sc:isolated}
We keep the same notations as in \pref{sc:main}.
Given a signature $\bp = (p_1, \ldots, p_n)$,
we define a group $G \subset \GL_n(\bC)$ by
\begin{equation}
G = \lc \diag(\alpha_1, \ldots, \alpha_n) \relmid
\alpha_1^{p_1}=\cdots=\alpha_n^{p_n} = \prod_{i=1}^{n} \alpha_i=1 \rc.
\end{equation}
The group $G$ acts naturally on $T$ in such a way that
$\diag(\alpha_1, \ldots, \alpha_n) \in G$ maps
$X_i \in T$ to $\alpha_i X_i$ for $i \in [1,n]$.
\begin{lemma} \label{lm:invariant_ring}
If $n\geq 4$ and $R$ is a hypersurface,
then $R$ coincides with the invariant ring $T^G$.
\end{lemma}
\begin{proof}
We have
$
N \vecomega = N \nu \vecc,
$
which is equal to $\vecc$ by \pref{pr:nu_N}.
This shows that
$
\bZ \vecomega \supset \bZ \vecc.
$
Note that
\begin{align}
L / \bZ \vecc \cong \bigoplus_{i=1}^n \bZ \vecX_i / ( p_i \vecX_i )
\end{align}
and
\begin{align}
\vecomega \equiv - \sum_{i=1}^n \vecX_i \mod \vecc.
\end{align}
It follows that
\begin{align}
L / \bZ \vecomega
&\cong \left. \lb \bigoplus_{i=1}^n \bZ \vecX_i \rb \right/ \lb p_i \vecX_i, \sum_{i=1}^n \vecX_i \rb.
\end{align}
This allows us to identify $L/\bZ \vecomega$
with the group of characters of $G$,
so that the ring $R$, which is the Veronese subring over $\vecomega$, is exactly
the $G$-invariant part of $T$.
\end{proof}
\begin{proposition} \label{pr:isolated}
If $n\geq 4$ and $R$ is a hypersurface,
then $R$ has an isolated singularity if and only if $R=T$.
\end{proposition}
\begin{proof}
The `if' part is clear since $T$ has an isolated singularity.
To prove the `only if' part,
assume that $R \subsetneq T$.
Then we have $s := \gcd(p_i,p_j) \ne 1$
for some $1 \le i < j \le n$ by \pref{lm:power} and \pref{cr:piqi}.
Define a subset of
\begin{align}
\Spec T = \{ (X_1, \ldots, X_n) \in \bA^n \mid X_1^{p_1} + \cdots + X_n^{p_n} = 0 \}
\end{align}
by
\begin{align}
P = \{ (X_1, \ldots, X_n) \in \Spec T \mid X_i = X_j = 0 \text{ and }
X_k \ne 0 \text{ for any } k \ne i, j \}.
\end{align}
Then the stabilizer subgroup of any point in $P$
with respect to the action of $G$
is given by
\begin{align}
\lc \diag(\alpha_1,\ldots,\alpha_n) \relmid
\alpha_i^{s}=1, \
\alpha_i \alpha_j=1 \text{ and }
\alpha_k=1 \text { for any } k \ne i, j \rc.
\end{align}
This is isomorphic to a cyclic subgroup of $\SL_2(\bC)$,
so that $\Spec R = (\Spec T)/G$ has a non-isolated family
of $A_{s-1}$-singularities along $P/G$.
\end{proof}
Now we prove \pref{th:isolated}:
\begin{proof}[Proof of \pref{th:isolated}]
To prove the `if' part,
assume that we have \eqref{eq:fraction}.
Then we have
\begin{align}
\nu
:= 1 - \sum_{i=1}^n \frac{1}{p_i}
= \frac{1}{\prod_{i=1}^n p_i}.
\end{align}
It follows that for any $i \in [1,n]$,
we have
\begin{align}
q_i
&:= \nu p_i N_i \\
&= \frac{1}{\prod_{i=1}^n p_i} \cdot p_i \cdot \lcm \lc p_j \relmid j \in [1,n] \setminus \{ i \} \rc \\
&\le \frac{1}{\prod_{i=1}^n p_i} \cdot p_i \cdot \prod_{j \in [1,n] \setminus \{ i \}} p_j \\
&= 1.
\end{align}
This implies $q_i=1$
since $q_i$ is a positive integer
by definition.
Hence we have $R=T$ (\pref{lm:power}),
which clearly has an isolated singularity at the origin.
To prove the `only if' part,
assume that $R$ has an isolated hypersurface singularity.
Then \pref{pr:isolated} shows $R=T$,
which implies $q_i=1$ for any $i \in [1,n]$
by \pref{lm:power}.
Then one has $\nu = 1/N$ and $\gcd(p_i, N_i) = 1$ for any $i \in [1,n]$
by \eqref{eq:nuN},
which implies $N = \lcm \lc p_i \relmid i \in [1,n] \rc = \prod_{i=1}^n p_i$
and \eqref{eq:fraction}
by \eqref{eq:df}.
\end{proof}
\section{Proof of \pref{th:a}}
\label{sc:a}
Since we assume that $n\geq 4$ and $R$ is a hypersurface, we have
\begin{equation} \label{eq:nu_N_is_1}
\nu:=1-\sum_{i=1}^{n}\frac{1}{p_i}=\frac{1}{N}
\end{equation}
by \pref{pr:nu_N}.
Define a function
$
m \colon \bZ \to \bZ
$
by
\begin{align} \label{eq:m}
m(k)
:= \ell(k \vecomega)
=k-\sum_{i=1}^{n} \myceil{\frac{k}{p_i}},
\end{align}
where the function
$
\ell \colon L \to \bZ
$
is defined by \eqref{eq:ka}.
Recall that we have
\begin{equation}
R_k=\Mbar(k \vecomega) \cdot
\lb \bC[\Ybar_1,\ldots,\Ybar_n] / \lb \Ybar_1+\cdots+\Ybar_n \rb \rb_{m(k)},
\end{equation}
where $\Mbar(k \vecomega)$ is defined by \eqref{eq:Mbar}.
Therefore the Hilbert series of $R$ is given by
\begin{align} \label{g_j}
F(t)
:= \sum_{k=0}^\infty (\dim_\bC R_k) \, t^k
= \sum_{k=0}^\infty c(k) t^n,
\end{align}
where
\begin{align}
c(k) = \begin{cases}
\displaystyle{\binom{m(k)+n-2}{n-2}} & m(k) \geq 0, \\
0 & m(k)<0.
\end{cases}
\end{align}
We can write
\begin{align}
F(t) =\sum_{j=0}^{N-1} g_j(t) t^j,
\end{align}
where
\begin{align}
g_j(t) = \sum_{k=0}^\infty c(j+Nk) t^{Nk}.
\end{align}
It follows from \eqref{eq:nu_N_is_1} and \eqref{eq:m} that
\begin{align}
m(j+N k)
=m(j)+\nu N k=m(j)+k.
\end{align}
Hence we have
\begin{align}
c(j+Nk) = \begin{cases}
\displaystyle{\binom{m(j)+k+n-2}{n-2}} & m(j)+k \geq 0, \\
0 & m(j)+k<0.
\end{cases}
\end{align}
Therefore, by the following \pref{lm:value_m_k},
\begin{align} \label{eq:gj}
g_j(t) = \sum_{k=0}^\infty \binom{k+n-2}{n-2} t^{N(k-m(j))}.
\end{align}
\begin{lemma} \label{lm:value_m_k}
We have $m(0)=0$, $m(1)=-(n-1)$, and
\begin{align}
-(n-2) \leq m(k) \leq 0
\end{align}
for $2 \leq k \leq N-1$.
\end{lemma}
\begin{proof}
It is clear that $m(0)=0$ and $m(1)=-(n-1)$.
For $k\leq N-1$, we have
\begin{align*}
m(k)
\le k - \sum_{i=1}^n \frac{k}{p_i}
= \nu k
= \frac{k}{N}
< 1,
\end{align*}
so that $m(k)\leq 0$.
On the other hand, for $k\geq 2$, we have
\begin{align*}
m(k)
\ge k - \sum_{i=1}^n \lb \frac{k}{p_i} + \frac{p_i - 1}{p_i} \rb
= \nu k-n-(\nu - 1)
= \nu(k-1)-(n-1)
> -(n-1),
\end{align*}
so that $m(k) \ge -(n-2)$.
\end{proof}
\begin{lemma} \label{lm:deg_F}
$\lb 1-t^N \rb^{n-1} F(t)$ is a polynomial of degree $(n-1)N+1$.
\end{lemma}
\begin{proof}
By \pref{eq:gj}, we have
\begin{align}
\lb 1-t^N \rb^{n-1} F(t)
&= \lb 1-t^N \rb^{n-1} \sum_{j=0}^{N-1} g_j(t) t^j \\
&= \lb 1-t^N \rb^{n-1} \sum_{j=0}^{N-1} t^j
\sum_{k=0}^\infty \binom{k+n-2}{n-2} t^{N(k-m(j))} \\
&= \sum_{j=0}^{N-1} t^{j-N \cdot m(j)} \lb 1-t^N \rb^{n-1}
\sum_{k=0}^\infty \binom{k+n-2}{n-2} t^{N k} \\
&= \sum_{j=0}^{N-1} t^{j-N \cdot m(j)},
\label{F_sum}
\end{align}
which is a polynomial of degree $(n-1)N+1$ by \pref{lm:value_m_k}.
\end{proof}
Now we prove \pref{th:a}:
\begin{proof}[Proof of \pref{th:a}]
If $R$ is generated by elements
of degrees $a_1, \ldots, a_n$
with one relation of degree $h$,
then the Hilbert series of $R$ is given by
\begin{align} \label{F_prod}
F(t)=\frac{1-t^h}{\prod_{i=1}^n \lb 1-t^{a_i} \rb}.
\end{align}
By \pref{lm:deg_F}, we have
\begin{align}
h=\sum_{i=1}^n a_i + 1.
\end{align}
This concludes the proof of \pref{th:a}.
\end{proof}
\begin{remark}
By (\ref{F_prod}), we have
\begin{align}
\lim_{t \to 1} (1-t)^{n-1} F(t)
=\lim_{t \to 1} \frac{1+t+\cdots+t^{h-1}}{\prod_{i=1}^n (1+t+\cdots+t^{a_i-1})}
=\frac{h}{\prod_{i=1}^n a_i}.
\end{align}
Similarly, by (\ref{F_sum}), we have
\begin{align}
\lim_{t \to 1} (1-t)^{n-1} F(t)
&= \lim_{t \to 1} \frac{(1-t)^{n-1}}{(1-t^N)^{n-1}} \sum_{j=0}^{N-1} t^{j-N \cdot m(j)} \\
&=\frac{1}{N^{n-1}} \cdot N \\
&=\frac{1}{N^{n-2}}.
\end{align}
Hence the following equality holds:
\begin{align}
\frac{h}{\prod_{i=1}^n a_i}=\frac{1}{N^{n-2}}.
\end{align}
A related discussion can be found
in \cite[Theorem 2.6]{MR586722}.
\end{remark}
\bibliographystyle{amsalpha}
|
\section{Introduction}\label{s:intro}
~~~~A fundamental task in statistics is to determine whether two
groups of variables are dependent. For example, in genomic analysis,
we might want to test whether two groups of genes are associated to
identify dependence between genetic pathways. In the brain imaging
research, we may want to discover whether sets of voxels from
different parts of the brain are related to explore functional
connectivity. In general, high-dimensional data analysis can be
simplified by identifying sets of independent variables.\\
\indent Testing of dependence is often reduced to testing for linear
dependence. Pearson correlation coefficient is a classical and
widely-used method for quantifying the strength of linear dependence
between two univariate variables. Spearman's rank correlation
coefficient \citep{Spearman} is a ranked-based version of Pearson
correlation coefficient which quantifies monotone correlation. Tests
based on correlation are powerful for testing
specific types of association, but lose power for other general types.\\
\indent For testing more general associations, the $\chi^2$ test of
independence and Hoeffding's test of independence
\citep{Dstatistics} are two classical nonparametric methods. These
tests are based on partitioning data into a contingency table. The
main drawback for $\chi^2$ test is that the result is sensitive to
the way the data are partitioned. Several approximations of the test
statistics of the Hoeffding's test are studied: \cite{Hoeffding2}
introduce an approximation by the concordances and discordances of a
$2\times 2$ contingency tables, and \cite{Wilding2008160} propose an
approximation by using two Weibull extensions. A relation between
the Hoeffding's test and the $\chi^2$ test statistics was noted by
\cite{modHoeff}, and they also suggested extending the idea of
\cite{Hoeffding2} to a $k\times k$ contingency tables, for $k>2$.
More recent methods related to the Hoeffding's test have been
proposed by \cite{HHG} and \cite{DDP}. Both of these tests are
consistent under general types of associations. Other methods for
testing for independence include the distance correlation test of
\cite{dCov} and the maximal information coefficient of \cite{MIC}.
Both the tests of \cite{HHG} and \cite{dCov} can be extended to
higher dimensions for testing joint independence of two or more
random vectors. Several Bayesian methods are available for testing
of independence. The simplest test of linear dependence between two
univariate random variables can be achieved by fitting a linear
model and inspecting the posterior distribution of the correlation
coefficient. Other methods were proposed for testing of independence
based on a contingency table
\citep{BayesContingency,BayesContingency3,BayesContingency2}.\\
\indent In this article, we propose a nonparametric Bayesian test of
independence between two groups of variables. We test the null
hypothesis of independence and the alternative hypothesis of
dependence. We specify nonparametric Bayesian models for the
response density under both hypotheses. Under the null hypothesis,
the joint distribution is taken to be the product of two independent
densities, both with nonparametric priors; under the alternative,
the full joint density has a nonparametric prior. The test is based
on the posterior probability of the alternative hypothesis. By
specifying nonparametric Bayesian models under each hypothesis, we
obtain an extremely flexible test which can capture both linear and
complex nonlinear relationships between groups of variables.\\
\indent The remainder of the article proceeds as follows. In Section
\ref{s:sm}, we introduce the statistical algorithm. The details of
the reversible jump MCMC algorithm use to compute the posterior
probability of the alternative hypothesis are provided in Section
\ref{s:Cd}. In Section \ref{s:sim}, we present a simulation study to
compare the power of the proposed test with other tests of linear
and nonlinear relationships. The method is illustrated using a
genetic data analysis in Section \ref{s:real}. Section
\ref{s:conclusion} concludes.
\section{Statistical model}\label{s:sm}
~~~~Let $\mat{X}_1\in\mathbb{R}^{D_1}$ and
$\mat{X}_2\in\mathbb{R}^{D_2}$ be random vectors in $D_1$ and $D_2$
dimensions, respectively, and denote
$\mat{X}=(\mat{X}_1,\mat{X}_2)$. The objective is to test whether
$\mat{X}_1$ and $\mat{X}_2$ are independent. The hypotheses are
\beqn
\mbox{H}_0&:&\mbox{$\mat{X}_1$ and $\mat{X}_2$ are independent and $f(\mat{X})=f_1(\mat{X}_1)f_2(\mat{X}_2)$}\nonumber\\
\mbox{H}_1&:&\mbox{$\mat{X}_1$ and $\mat{X}_2$ are dependent and
$f(\mat{X})$ cannot be factorized}\nonumber\eeqn In other words,
when they are independent, the joint density can be factorized as
the product of
two lower-dimensional densities.\\
\indent Under both hypotheses, the densities are modeled using
Dirichlet process mixture (DPM) prior. Under $\mbox{H}_0$,
$f_1(\mat{X}_1)$ and $f_2(\mat{X}_2)$ follow independent DPM priors;
under $\mbox{H}_1$ when $\mat{X}_1$ and $\mat{X}_2$ are not
independent, the joint distribution is assumed to follow a DPM
prior. The following subsections describe the independent and joint
DPM priors.
\subsection{The independent DPM prior}
~~~~When $\mat{X}_1$ and $\mat{X}_2$ are independent,
$f_j(\mat{X}_j)$, $j=1,2$, are assumed to follow the DPM prior
independently. The DPM prior can be written as the infinite mixture
\beq\label{indepDPM}
f_j(\mat{X}_j)=\sum_{l=1}^{\infty}w_{lj}\phi_j(\mat{X}_j~|~\mat{\mu}_{lj},\mat{\Sigma}_j),\eeq
where $w_{lj}$ is the mixture weight, $\phi_j$ is assigned to be the
$D_j-$dimensional multivariate normal distribution (MVN) in this
analysis, $\mat{\mu}_{lj}$ is the mean vector of the $l^{th}$
mixture
component, and $\mat{\Sigma}_j$ is the covariance matrix.\\
\indent The mixture weights $w_{lj}$ are modeled by the
stick-breaking construction with concentration parameter $d_j$. The
weights $w_{lj}$ are modeled in terms of latent
$v_{lj}\sim\mbox{Beta}(1,d_j)$. The first weight is $w_{1j}=v_{1j}$.
The remaining elements are modeled as
$w_{lj}=v_{lj}\prod_{i=1}^{l-1}(1-v_{ij})$, where
$\prod_{i=1}^{l-1}(1-v_{ij})=1-\sum_{i=1}^{l-1}w_{ij}$ is the
remaining probability after accounting first $l-1$ mixture weights.
The number of mixture components is truncated by a sufficiently
large number $K$ (i.e. $l=1,...,K$), where the last term $v_K$ is
fixed to be 1 to ensure that $\sum_{l=1}^Kw_{lj}=1$.\\
\indent The mean vectors $\mat{\mu}_{lj}$ have priors
$\mat{\mu}_{lj}\sim\mbox{MVN}(\mat{0},\mat{\Omega}_j)$. The
covariance matrices $\mat{\Sigma}_j$ and $\mat{\Omega}_j$ are
parameterized as $\mat{\Sigma}_j=r\mat{S}_j$, and
$\mat{\Omega}_j=(1-r)\mat{S}_j$. Under this model, $\mat{S}_j$ is
the covariance matrix for $\mat{X}_j$ marginally over the mixture
means $\mat{\mu}_{lj}$, and $r$ is the proportion of the total
variance attributed to the variance within each mixture component.
The marginal covariance $\mat{S}_j$ is assigned to have inverse
Wishart prior distribution, and to facilitate computing, the prior
of $r$ is a discrete uniform distribution with support
$r\in\{0,0.01,...,1\}$. The concentration parameter $d_j$ has prior
distribution $\mbox{Gamma}(a,b)$.
\subsection{The joint DPM prior}
~~~~When $\mat{X}_1$ and $\mat{X}_2$ are not independent,
$f(\mat{X})$ is assumed to follow the joint DPM
prior\beq\label{jointDPM}
f(\mat{X})=\sum_{l=1}^{\infty}w_{l}\phi(\mat{X}~|~\mat{\mu}_{l},\mat{\Sigma}),\eeq
where $w_{l}$ is the mixture weight, $\phi$ is the
$(D_1+D_2)-$dimensional MVN distribution, $\mat{\mu}_{l}$ is the
mean vector of the $l^{th}$ mixture component, and $\mat{\Sigma}$ is
the covariance matrix. The number of mixtures is truncated by the
same number $K$ as in the independent model. The mixture weights
$w_{l}$ are again modeled by the stick-breaking algorithm with
concentration parameter $d$. The mean vectors $\mat{\mu}_l$ have
priors $\mat{\mu}_l\sim \phi(\mat{0},\mat{\Omega})$. The covariance
matrices $\mat{\Sigma}$ and $\mat{\Omega}$ are modeled as
$\mat{\Sigma}=r\mbox{diag}(\mat{S})$ and
$\mat{\Omega}=(1-r)\mat{S}$, where $\mat{S}$ is the covariance
matrix for $\mat{X}$, and $\mbox{diag}(\mat{S})$ is the diagonal
form of $\mat{S}$. In other words, under the joint DPM prior, we
assign non-diagonal structure for the $\mat{\Omega}$, and diagonal
structure for the $\mat{\Sigma}$. We found that diagonalizing
$\mat{\Sigma}$ greatly improved computational stability . The priors
for $\mat{S}$, $r$, and $d$ are the same as in the independent DPM
prior.
\subsection{Bayesian test of independence}\label{ss:transition}~~~~The Bayesian hypothesis test of
independence is based on the Bayes factor (BF)\beq
\mbox{BF}=\frac{P(\mbox{H}_1~|~\mat{X})/P(\mbox{H}_0~|~\mat{X})}{P(\mbox{H}_1)/P(\mbox{H}_0)}=\frac{P(\mat{X}~|~\mbox{H}_1)}{P(\mat{X}~|~\mbox{H}_0)}.\eeq
The null is rejected if $\mbox{BF}>T$, where $T$ is a threshold
parameter. The threshold parameter $T$ can be chosen based on rules
of thumb about the weight of evidence favoring $\mbox{H}_1$. For
example, \cite{BF} suggest that $\mbox{BF}=10$ is a strong evidence
for $\mbox{H}_1$. Alternatively, in the simulation study in Section
\ref{s:sim}, we select $T$ to control the Type I error rate. In the
analysis of genetic data in Section \ref{s:real}, multiple tests are
performing simultaneously, therefore we select $T$ to control the
Bayesian false discovery rate.
\section{Computing details}\label{s:Cd}
~~~~Computing the Bayes factor requires computing the posterior
probability of each hypothesis. This is accomplished using a
reversible jump MCMC (RJMCMC) algorithm as described below.
\subsection{Reparameterization and hyperparameters}
~~~~The updating algorithm of the DPM prior is facilitated by
introducing the equivalent clustering model. The mixture form in
(\ref{indepDPM}) can be written as
$$f_j(\mat{X}_j~|~g_j=l)=\phi_j(\mat{X}_j~|~\mat{\mu}_{lj},\mat{\Sigma}_j),$$ which draws an
auxiliary cluster label $g_j\in\{1,...,K\}$ with $P(g_j=l)=w_{lj}$.
Similarly, the model in (\ref{jointDPM}) is equivalent to
$$f(\mat{X}~|~g=l)=\phi(\mat{X}~|~\mat{\mu}_{l},\mat{\Sigma}),$$ with cluster label $g$ and
$P(g=l)=w_{l}$. Under the clustering model, the full conditionals of
all the parameters
are conjugate.\\
\indent In addition, we introduce model indicator parameter $M$,
where
\[M\in\left\{\begin{array}{ll}
I & \mbox{if $\mat{X}_1$ and $\mat{X}_2$ are independent ($\mbox{H}_0$ is true)}\\
J & \mbox{if $\mat{X}_1$ and $\mat{X}_2$ are not independent ($\mbox{H}_1$ is true)}.\end{array} \right.\]
Under each MCMC step, we propose a new indicator $M^{'}$ in the
Markov chain, and decide whether to accept the new status $M^{'}$.
The probability $P(\mbox{H}_1~|~\mat{X})$ is then approximated by
$\sum_{i=1}^NI(M^{(i)}=J)/N$, where $N$ is the number of MCMC
samples and $M^{(i)}$ is the model status for the $i^{th}$ MCMC
sample.\\
\indent Throughout this article, we let the number of mixture
components truncated at $K=20$ and the hyperparameters in the
stick-breaking procedure $(a,b)$ are fixed under different sample
sizes $n$ as presented in Table \ref{abtable}.
\begin{table}[h]
\caption{Hyperparameters $(a,b)$ under different sample sizes
$n$.}\label{abtable}
\begin{center}
\begin{tabular}
{|c|c|c|} \hline $n$ & $a$ & $b$\\
\hline 100 & 1.5 & 2.5\\
200 & 1.0 & 4.0\\
300 & 1.0 & 4.5\\
500 & 0.8 & 4.6\\ \hline
\end{tabular}
\end{center}
\end{table}
\subsection{Pseudo code for the DPM test of independence algorithm}\label{ss:pseudo}
~~~~Let $\mat{\Theta}_M$ denote the DPM parameters
($\mat{\Theta}_M=\{\mat{\mu}_{11},...,\mat{\mu}_{K2}$, $r$,
$\mat{S}_1$, $\mat{S}_2$, $w_{11},...,w_{K2}$, $d_1,d_2$\} if $M=I$,
and $\mat{\Theta}_M=\{\mat{\mu}_{1},...,\mat{\mu}_{K}$, $r$,
$\mat{S}$, $w_{1},...,w_{K}$, $d$\} if $M=J$). The algorithm of the
DPM test of
independence is described as follows:\\
\textbf{Step 0}: Select initial values for $M$ and
$\mat{\Theta}_M$.\\
\textbf{Step 1}: Update $\mat{\Theta}_M$ given $M$ using the Gibbs
sampling.\\
\textbf{Step 2}: Update $M$ given the parameters $\mat{\Theta}_M$.\\
~~\textbf{Step 2.1}: Generate proposed model status $M^{'}$ with
$P(M^{'}=I)=P(M^{'}=J)=0.5$.\\
~~\textbf{Step 2.2}: If $M=M^{'}$, then so back to \textbf{Step 1}.\\
~~\textbf{Step 2.3}: If $M=I$ and $M^{'}=J$, then propose
$\mat{\Theta}_{M^{'}}$ required for the joint DPM prior ($\mbox{H}_1$).\\
~~\textbf{Step 2.4}: If $M=J$ and $M^{'}=I$, then propose
$\mat{\Theta}_{M^{'}}$ required for the independent DPM prior ($\mbox{H}_0$).\\
~~\textbf{Step 2.5}: Accept $M^{'}$ with probability
min$\{1,\alpha(M,M^{'})\}$.\\
\textbf{Step 3}: Back to \textbf{Step 1}.\\
The full conditionals requires for Step 1 are all standard and are
given in Appendix \ref{MCMCindep} for $M=I$, and Appendix
\ref{MCMCjoint} for $M=J$. Details on the RJMCMC steps are provided
below in Section \ref{ss:RJMCMCsteps}.
\subsection{Steps of the RJMCMC algorithm}\label{ss:RJMCMCsteps}
~~~~The parameter spaces under the independent and the joint DPM
priors are different, so moving between these two parameter spaces
becomes a trans-dimensional problem. Reversible jump MCMC (RJMCMC)
was first introduced by \cite{RJMCMC1}, which can be thought of as a
generalized Metropolis-Hastings algorithm for the trans-dimensional
updates.\\
\indent Under the current model status $M$, the propose model status
$M^{'}$ is randomly assigned to be either $I$ or $J$ with acceptance
probability min$\{1,\alpha(M,M^{'})\}$, where
\beq\label{RJMCMCacc}\alpha(M,M^{'})=\frac{l_{M^{'}}\cdot\pi_{M^{'}}\cdot
q_{M~|~M^{'}}\cdot p_{M^{'}\rightarrow M}}{l_{M}\cdot\pi_{M}\cdot
q_{M^{'}~|~M}\cdot p_{M\rightarrow
M^{'}}}\left|\mathfrak{J}\right|,\eeq where $l_M$ and $\pi_M$ are
the likelihood function and the prior distribution under model $M$,
$q_{M^{'}~|~M}$ is the candidate distribution of the parameters when
proposing for model $M^{'}$ under model $M$, $p_{M\rightarrow
M^{'}}$ is the probability of proposing $M^{'}$ conditional on the
current status $M$, and $\left|\mathfrak{J}\right|$ is the Jacobian.
As $M^{'}$ is randomly picked from $\{I,J\}$, $p_{M\rightarrow
M^{'}}$ and $p_{M^{'}\rightarrow M}$ are equal in the algorithm.
Note that when $M=M^{'}$, it becomes the usual fixed-dimensional
MCMC algorithm as $\alpha(M,M^{'})=1$; when $M\neq M^{'}$, the
candidate distribution of the parameters $q$ is then for balancing
the parameter spaces
between the independent and joint models.\\
\indent Recall that $\mat{\Theta}_M$ and $\mat{\Theta}_{M^{'}}$
denote the DPM parameters under models $M$ and $M^{'}$,
respectively, and the truncated number $K$ under both models are
assigned to be identical. We first examine the case when $X_1$ and
$X_2$ are univariate random variables ($D_1=D_2=1$) with the current
model status $M=I$, and the proposed model is $M'=J$. Denote the
covariance matrix under the joint
model as $\mat{S}=\bigl(\begin{smallmatrix} S_{11}^2&S_{11}S_{22}\rho_J\\
S_{11}S_{22}\rho_J&S_{22}^2
\end{smallmatrix} \bigr)$. We assign the $2\times K$ mean vector $\mat{\mu}$ to be the same in both the independent and joint DPM models.
Also, we assign the variances $S_{11}^2$ and $S_{22}^2$, and $r$ to
be the same across different model statuses. Therefore, this move
only requires proposing the parameters under the joint DPM prior in
(\ref{jointDPM}): the cluster label $g'_J$, $\rho'_J$, the
concentration parameter $d'_J$, and the mixture weights
$\mat{w}'_J$. The concentration parameter $d'_J$ is proposed by
$d'_J\sim\mbox{Gamma}(\bar{d_I},1)$, where $\bar{d_I}$ is the mean
of $d_I$, and then the mixture probabilities $\mat{w}'_J$ is
proposed from the stick-breaking procedure with concentration
parameter $d'_J$. The cluster label $g'_J$ is proposed from the full
conditional distribution given in Appendix \ref{fullcond}. The
details of the mapping for each parameter is described in the end of
this section.\\
\indent Conversely, if the current model status is $M=J$ and the
proposed model status is $M'=I$, the parameters of the independent
model described in (\ref{indepDPM}) are proposed as follow: The
concentration parameter $d'_I\sim\mbox{Gamma}(d_J,1)$, and the
mixture weights $\mat{w}'_I$ are again proposed by the
stick-breaking procedure with concentration parameter $d'_I$. The
cluster label $g'_I$ is again proposed by the full conditional
distribution given in Appendix \ref{fullcond}.\\
\indent For dimension matching under the RJMCMC algorithm, the
bijection map is described below for the case where $M=I$ and
$M'=J$. The reverse move uses the same map. Let \beqn
\theta_M&=&\{\mat{\mu},S_{11},S_{22},r,\mat{w}_I,d_I,g_I\}\nonumber\\
u&=&\{\rho'_J,\mat{w}'_J,d'_{J},g'_J\}\nonumber\\
\theta_{M'}&=&\{\mat{\mu},S_{11},S_{22},r,\mat{w}_J,\mat{d}_J,g_J,\rho_J\}\nonumber\\
u'&=&\{\mat{w}'_I,d'_{I},g'_I\}.\nonumber\eeqn Then we assign
$\mat{\Theta}_M=\{\theta_M,u\}$,
$\mat{\Theta}_{M'}=\{\theta_{M'},u'\}$. The bijection function $h$
has the form
$$h(\mat{\Theta}_M)=h(\theta_M,u)=\mat{\Theta}_{M'}=\{\theta_{M'},u'\},$$ which is
a one-to-one bijection map with: $\mat{w}_I\rightarrow\mat{w}'_I$,
$d_I\rightarrow d'_I$, $g_I\rightarrow g'_I$, $\rho'_J\rightarrow
\rho_J$, $\mat{w}'_J\rightarrow \mat{w}_J$, $d'_J\rightarrow d_J$,
and $g'_J\rightarrow g_J$. Hence, the Jacobian
$|\mathfrak{J}|=\left|\frac{\partial(\theta'_{M'},u')}{\partial(\theta_{M},u)}\right|=1$.\\
\indent When $D_1+D_2>2$, the transition of the covariance matrices
between the independent and joint models becomes more complicated as
the off-diagonal elements are harder to propose than in the
bivariate case. One way to alleviate this concern is to assume the
covariance matrix $\mat{S}$
under the joint model is a block-diagonal matrix $\mat{S}=\bigl(\begin{smallmatrix} \mat{S}_{1}&0\\
0&\mat{S}_2
\end{smallmatrix} \bigr)$, where $\mat{S}_i$ is a $D_i\times D_i$
covariance matrix of $\mat{X}_i$ for $i=1,2$. However, in the
simulation study and the real data analysis of this article, we will
focus on the case where $X_1$ and $X_2$ are univariate random
variables.
\section{Simulation Study}\label{s:sim}
~~~~The simulation study focuses on testing for dependence between
two univariate variables. The objective is to compare the power of
each method under linear and nonlinear dependence. In the following
subsections, we introduce the data generation procedure, the
competing methods, and the simulation results.
\subsection{Data generation}\label{ss:DG}
~~~~ The seven different types of data sets are simulated. Scenarios
5 and 6 are designed from \cite{DDP}.
\begin{enumerate}
\item Independent normal (Null): $X_j\sim\mbox{N}(0,1)$, for j=1,2.
\item Bivariate normal (BVN): $(X_1,X_2)\sim\mbox{BVN}\left[\mat{0},\bigl(\begin{smallmatrix} 1&\rho\\
\rho&1
\end{smallmatrix} \bigr)\right]$, where $\rho=0.2$.
\item Horseshoe (HS): $X_1\sim\mbox{N}(0,1)$, $X_2~|~X_1\sim\mbox{N}(\rho X_1^2,1)$, where $\rho=0.2$.
\item Cone: $X_1\sim\mbox{U}(0,1)$, $X_2~|~X_1\sim\mbox{N}\left[0,(\rho X^2_1+0.1)^2\right]$, where $\rho=0.1$.
\item W: $X_1\sim\frac{1}{n}\sum_{i=1}^{n}\mbox{U}(a_i,a_i+\frac{1}{3})$, $X_2~|~X_1\sim\mbox{U}\left[3(X_1^2-\frac{1}{2})^2,3(1+X_1^2-\frac{1}{2})\right]$,
where $a_1=-1$, $n$ is the number of samples, and
$a_i=a_{i-1}+\frac{2}{n}$, for $i>1$.
\item Circle: $(X_1,X_2)\sim\frac{1}{n}\sum_{i=1}^{n}\mbox{BVN}\left[\theta_i,\bigl(\begin{smallmatrix} \frac{1}{9}&0\\
0&\frac{1}{64}
\end{smallmatrix} \bigr)\right]$, where
$\theta_i=\left[\sin(a_i\pi),\cos(a_i\pi)\right]$, and $a_i$ is
defined as in W.
\end{enumerate}
~~~~Each scenario is generated with the algorithms introduced above
with sample size $n=100,200$, and 500. Then for each dimension, we
standardize the data to have mean zero and variance one. We plot the
data when $n=200$ in Figure \ref{SimData} along with the true
density. The responses are dependent for designs 2-6. Design 3-6 are
all examples of the challenging dependent but uncorrelated random
variables and thus the usual test of correlation will miss this
dependence.
\begin{figure}[h]
\setlength{\abovecaptionskip}{1ex} \centering
\includegraphics[height=13cm,width=15cm]{SimData.eps}
\caption{True log density (background color) and one simulated data
set (points) for each simulation design.}\label{SimData}
\end{figure}
\subsection{Methods for testing of independence}
~~~~We compare six methods in the simulation study (described in
detail in the Appendix). Each method is controlled to have type I
error rate approximately equal to 0.05.
\begin{enumerate}
\item Linear regression (LR): The model $X_2=\beta_0+\beta_1X_1+\epsilon$, $\epsilon\sim\mbox{N}(0,1)$ is fitted by least squares and the linear association is
determined by the test of $\beta_1=0$.
\item E-statistics (ES) \citep{dCov}: The testing procedure is by calculating the
distance covariance between $X_1$ and $X_2$.
\item Heller-Heller-Gorfine method (HHG) \citep{HHG}: The test statistic is
based on the sum of all likelihood ratio tests of $2\times 2$
contingency tables formed by the pairwise distances within each of
$X_1$ and $X_2$.
\item Data Derived Partitions method (DDP) \citep{DDP} with $3\times 3$ contingency tables:
The DDP method is similar to the HHG method, but only designed for
univariate random variables. The test statistic is based on the sum
of all likelihood ratio tests of $3\times 3$ contingency tables
formed by the observed values.
\item Maximal Information Coefficient method (MIC) \citep{MIC}: It is a rank-order test
statistic which is calculated from the largest achievable mutual
information under different grid sizes.
\item The DPM test of independence (DPM): The proposed test is described in
Section \ref{s:sm}. $X$ is first marginally transformed to be
standard normal distribution. The normal score transformation makes
the proposed method a distribution-free testing procedure.
Therefore, the threshold for the BF in Section \ref{s:sm} that
controls Type I error can be determined by the permutations of the
transformed data. The threshold $T$ for the Bayes factor is computed
from 300 permutations of the sample.
\end{enumerate}
\subsection{Simulation results}
~~~~The results are presented in Table \ref{simutable} with sample
sizes $n=100$, 200 and 500. The first three rows of the table are
the type I error rate for each method under different samples sizes,
which is controlled for all methods ($\mbox{Type I error rate is
between }0.03\mbox{ to }0.09$). The following rows give the power of
each method under different scenarios and sample sizes. It is clear
that as the sample size $n$ increases, the powers increase for all
the methods except the LR method under the HS, Cone, and Circle
scenarios because of the nonlinear associations of these
scenarios.\\
\indent When the data are generated from bivariate normal
distribution, the LR method has the highest power. This is expected
because the LR method is theoretically the most powerful test under
this scenario. The ES
and DPM tests are the second best among other comparing tests.\\
\indent The DPM test outperforms all other methods when data are
generated from the HS and the W shapes. Under the Cone shape data,
the HHG and the DPM tests both perform well. For the Circle design,
the HHG, DDP, and DPM tests all have power greater than 0.9 starting
from small sample sizes, and the ES and MIC have lower power.\\
\indent In summary, the LR method is able to capture linear
association but loses power in the nonlinear cases. The ES method is
able to capture linear and nonlinear associations, but loses power
in some of the nonlinear cases. The HHG and DDP methods both have
high power in testing of nonlinear associations, but lose power in
the linear association, especially the HHG method. The MIC method is
a relatively conservative test compared to all other methods, and
this problem is discussed by \cite{ComMIC}. The proposed method not
only shows the ability to capture the linear association, but is
also powerful for detecting nonlinear associations in the simulation
study.
\begin{table}[h]
\caption{Power of each test (columns) for each simulation settings
and sample size $n$ (rows). A $*$ indicates that the power is
significantly different than the power of DPM
test.}\label{simutable}
\begin{center}
\scalebox{0.95}{
\begin{tabular}
{|c|c|c|c|c|c|c|c|} \hline Type & $n$ & LR & ES & HHG & DDP & MIC & DPM\\
\hline Null & 100 & 0.06 & 0.04 & 0.03 & 0.02 & 0.03 & 0.03\\
& 200 & 0.06 & 0.05 & 0.03 & 0.02 & 0.07 & 0.05\\
& 500 & 0.09 & 0.08 & 0.05 & 0.09 & 0.02 & 0.09\\
\hline BVN & 100 & $0.57^{*}$ & 0.49 & $0.24^{*}$ & $0.38^{*}$ & $0.13^{*}$ & 0.43\\
& 200 & $0.84^{*}$ & 0.78 & $0.37^{*}$ & $0.61^{*}$ & $0.25^{*}$ & 0.76\\
& 500 & 0.99 & 0.99 & $0.83^{*}$ & 0.99 & $0.40^{*}$ & 0.99\\
\hline HS & 100 & $0.11^{*}$ & $0.22^{*}$ & 0.42 & 0.39 & $0.12^{*}$ & 0.44\\
& 200 & $0.05^{*}$ & $0.48^{*}$ & $0.53^{*}$ & $0.60^{*}$ & $0.25^{*}$ & 0.68\\
& 500 & $0.10^{*}$ & $0.96 $ & $0.99 $ & 1.00 & $0.42^{*}$ & 1.00\\
\hline Cone & 100 & $0.03^{*}$ & $0.25^{*}$ & $0.54^{*}$ & $0.33 $ & $0.17^{*}$ & 0.36\\
& 200 & $0.08^{*}$ & $0.56^{*}$ & 0.87 & $0.73^{*}$ & $0.37^{*}$ & 0.84\\
& 500 & $0.09^{*}$ & 1.00 & 1.00 & 1.00 & $0.80^{*}$ & 1.00\\
\hline W & 100 & $0.54^{*}$ & $0.42^{*}$ & $0.55^{*}$ & $0.75^{*}$ & $0.34^{*}$ & 0.92\\
& 200 & $0.84^{*}$ & $0.83^{*}$ & $0.92^{*}$ & 1.00 & $0.70^{*}$ & 1.00\\
& 500 & 1.00 & 1.00 & 1.00 & 1.00 & $0.98 $ & 1.00\\
\hline Circle & 100 & $0.00^{*}$ & $0.00^{*}$ & 0.96 & 0.99 & $0.20^{*}$ & 0.99\\
& 200 & $0.00^{*}$ & $0.22^{*}$ & 1.00 & 1.00 & $0.37^{*}$ & 1.00\\
& 500 & $0.00^{*}$ & 1.00 & 1.00 & 1.00 & $0.95^{*}$ & 1.00\\ \hline
\end{tabular}
}
\end{center}
\end{table}
\section{Real data analysis}\label{s:real}
~~~~We compare the six methods in the simulation study on the gene
expression data set from \cite{geneexpression}. Studies of
associations between genes can be found in \cite{gene1} and
\cite{gene2}. The number of observations is $n=300$ for each gene,
and we select 94 genes on chromosome 1 after removing samples with
missing values. The objective is to test
the pairwise associations within these 94 genes. A total of $\bigl(\begin{smallmatrix} 94\\
2
\end{smallmatrix} \bigr)=4371$ hypotheses tests of independence are
performed. Because of the large number of tests, we control false
discovery rate (FDR) at the 0.05 level rather than Type I error. The
Bayesian FDR (BFDR) control procedure is applied
\citep{BayesFDR,BayesFDR3,FDRcontrol,BayesFDR2} for the DPM test,
and the Benjamini--Hochberg procedure \citep{BH} is applied for
the other methods.\\
\indent The Cohen's $\kappa$ statistic \citep{COHEN} is used to
measure agreement between tests. The $\kappa$ statistic is
\beq\kappa=\frac{P_a-P_e}{1-P_e},\nonumber\eeq where $P_a$ is the
proportion of agreements between the two methods among the $N=4371$
tests, and $P_e$ is the theoretical proportion of agreements under
independence. Larger values of $\kappa$ represents more agreement
between the tests. The number of rejections among $N=4371$ tests and
the $\kappa$ statistics of pairwise methods are presented in Table
\ref{realresult2}.
\begin{table}[h]
\caption{Numbers of rejections (of the $N=4371$ tests), and Cohen's
$\kappa$ statistics for each pair of methods.}\label{realresult2}
\begin{center}
\begin{tabular}
{|c|c|c|c|c|c|c|c|} \hline Methods & Number of rejections & LR & ES & HHG & DDP & MIC & DPM\\
\hline LR & 2404 & 1.000 & 0.472 & 0.301 & 0.404 & 0.082 & 0.452\\
ES & 3352 & -- & 1.000 & 0.686 & 0.830 & 0.036 & 0.779\\
HHG & 3442 & -- & -- & 1.000 & 0.751 & 0.032 & 0.720\\
DDP & 3350 & -- & -- & -- & 1.000 & 0.036 & 0.814\\
MIC & 249 & -- & -- & -- & -- & 1.000 & 0.042\\
DPM & 3231 & -- & -- & -- & -- & -- & 1.000\\ \hline
\end{tabular}
\end{center}
\end{table}
\noindent The $\kappa$ statistics show that the ES, HHG, DDP, and
the DPM tests have similar testing powers in this gene expression
data sets, and the number of rejections among these tests are
similar (3231 to 3442). The LR test only captures the linear
associations between genes, and the MIC has
the lowest power as in the simulation study.\\
\indent In Figure \ref{4cases}, we plot six pairs of genes where
there are disagreements among the tests. In the upper two plots
(gene 94 versus gene 8, and gene 88 versus gene 15), the
associations between these pairs of genes are detected by the DPM
test, but not the other tests. The figure shows that between gene 94
and gene 8, there is a horseshoe pattern of dependence, and a
nonlinear relationship between gene 88 and gene 15. In the middle
two plots (gene 17 versus gene 1, and gene 89 versus gene 24), the
ES, HHG, DDP, and DPM tests all flag associations between genes, but
not the LR and the MIC tests. The figure shows that gene 17 and gene
1 have a cone-shape association, and genes 89 and 24 have a
clustering relationship. The bottom two plots (gene 92 versus gene 2
and gene 30 versus gene 6) are the cases where only the LR, ES and
DPM tests flag associations between genes. These three tests are
powerful in testing the linear associations, and the figure shows
linear relationships between genes in these two pairs.\\
\begin{figure}[h]
\setlength{\abovecaptionskip}{1ex} \centering
\includegraphics[height=10cm,width=9cm]{4cases.eps}
\caption{Six pairs of genes where there are disagreements among the
tests. The red lines are the linear regression fitted
lines.}\label{4cases}
\end{figure}
\section{Conclusion}\label{s:conclusion}
~~~~We propose a nonparametric Bayesian test of dependence by
calculating the Bayes factor using the Dirichlet process mixture
model and the reversible jump MCMC algorithm. We compare our method
with the linear model, distance correlation method, HHG, DDP, and
MIC in the simulation study and also in the gene expression data
sets. The simulation results show that the proposed test is
competitive in testing both linear and nonlinear relationships.\\
\indent In the gene expression data analysis, we performed 4371
multiple testing on the gene expression data in comparing pairwise
genes. The proposed test shows similar performance with the distance
correlation, DDP, and HHG methods, and detects some cases that other
methods do not detect. It also shows that the proposed method is
powerful on both linear and nonlinear relationships in the pairwise
gene comparisons.
\newpage
|
\section{Introduction}
The field-antifield formalism \cite{BV,BV1}, summarizing numerous
attempts to find correct quantization rules for various types of
gauge models \cite{dWvH,FT,Nie1,Kal,Nie2}, is a powerful covariant
quantization method which can be applied to arbitrary gauge
invariant systems. This method is based on the fundamental principle
of BRST invariance \cite{brs1,t} and has a rich new geometry
\cite{Sch}. One of the most important objects of the field-antifield
formalism is an odd symplectic structure called antibracket and
known to mathematicians as the Buttin bracket \cite{Butt}. In terms
of the antibracket the master equation and the Ward identity for
generating functional of the vertex functions (effective action) are
formulated. It is an important property that the antibracket is
preserved under the anticanonical transformations which are dual to
canonical transformations for a Poisson bracket. An important role
and rich geometric possibilities of general anticanonical
transformations in the field-antifield formalism have been realized
in the procedure of gauge fixing \cite{VLT} (see, also \cite{BBD}).
The original procedure of gauge fixing \cite{BV,BV1} corresponds in
fact to a special type of anticanonical transformation in an action
being a proper solution to the quantum master equation. That type of
transformations is capable to yield admissible gauge-fixing
conditions in the form of equations of arbitrary Lagrangian surfaces
(constraints in the antibracket involution) in the field-antifield
phase space. Thereby, the necessary class of admissible gauges was
involved actually. The latter made it possible to describe in
\cite{VLT} the structure and renormalization of general gauge
theories in terms of anticanonical transformations. As the authors
\cite{VLT} assumed the use of regularizations in which $\delta(0)
= 0$ in local field theories, they based themselves on the use of
general anticanonical transformations in an action being a proper
solution to the classical master equation. In turn, the gauge
dependence and the structure of renormalization of the effective
action have been analyzed by using infinitesimal anticanonical
transformations only.
In the present article, we extend the use of anticanonical
transformations in the field-antifield formalism from the
infinitesimal level to the finite one, and explore a gauge fixing
procedure for general gauge theories, based on arbitrary
anticanonical transformations in an action being a proper solution
to the quantum master equation with fixed boundary condition. Now it
is worthy to notice the difference between the properties of the
classical and quantum master equations under anticanonical
transformations. The classical master equation is covariant under
anticanonical transformations, as its left-hand side is the
antibracket of the action with itself. In contrast to that, the form
of the quantum master equation is not maintained under anticanonical
transformations. One should accompany the anticanonical
transformation by multiplying the exponential of $i/\hbar$ times the
transformed action with the square root of the superjacobian of that
anticanonical transformation. We will call such an operation an
{\it anticanonical master transformation} and the corresponding
action a {\it master transformed action}. Thus, one can say that
the form of the quantum master equation is maintained under the
anticanonical master transformation.
We consider in all details the relationship between
the two descriptions (in terms of the generating functions and the generators)
for arbitrary finite anticanonical transformations.
Finally, let us notice the study \cite{BLT-BV}, among the other
recent developments, where a procedure was found to connect
generating functionals of the Green functions for a gauge system
formulated in any two admissible gauges with the help of finite
field-dependent BRST transformations.
\\
\section{Field-Antifield Formalism}
\noindent
The starting point of the field-antifield formalism \cite{BV} is a
theory of fields $\{{\cal A}\}$ for which the initial classical
action $S_0({\cal A})$ is assumed to be invariant under the gauge
transformations $\delta {\cal A}=R({\cal A}) \xi$. Here $\xi$ are
arbitrary functions of space-time coordinates , and $\{R({\cal
A})\}$ are generators of gauge transformations. The set of
generators is complete but, in general, maybe reducible and forms an
open gauge algebra so that one works with general gauge theories.
Here we do not discuss these points, referring to original papers
\cite{BV,BV1}. The structure of the gauge algebra determines
necessary content of total configuration space of fields
$\{\varphi^i\;(\varepsilon(\varphi^i)=\varepsilon_i)\}$ involving
fields
$\{{\cal A}\}$ of initial
classical system, ghost and antighost fields, auxiliary fields and ,
in case of reducible generators, pyramids of extra ghost and
antighost fields as well as pyramids of extra auxiliary fields.
To each field $\varphi^i$ one introduces an antifield
$\varphi^*_i$, whose statistics is opposite to that of the
corresponding fields $\varphi^i$, $\varepsilon (\varphi^*_i) =
\varepsilon_i + 1 $. On the space of the fields $\varphi^i$ and
antifields $\varphi^*_i$ one defines an odd symplectic structure
$(\;,\;)$ called the antibracket
\begin{eqnarray}
\label{defAB} (F, G)\equiv
F\left(\overleftarrow{\pa}_{\varphi^i}\overrightarrow{\pa}_{\varphi^*_i}-
\overleftarrow{\pa}_{\varphi^*_i}\overrightarrow{\pa}_{\varphi^i}\right)G
\end{eqnarray}
and the nilpotent fermionic operator $\Delta$,
\begin{eqnarray}
\label{DeltaBV} \Delta=(-1)^{\varepsilon_i} \partial_{\varphi^{i}}\partial
_{\varphi_{i}^{\ast}},\quad\Delta^{2}=0,
\quad \varepsilon (\Delta)=1.
\end{eqnarray}
Here the notation
\beq
\pa_{\varphi^i}=\frac{\pa }{\pa\varphi^i},\quad
\pa_{\varphi^*_i}=\frac{\pa}{\pa\varphi^*_i}
\eeq
is introduced. In terms of the antibracket and
$\Delta$-operator the quantum master equation is formulated
as
\begin{eqnarray}
\label{MastEBV} \frac {1}{2}
({\cal S},{\cal S})=i\hbar{\Delta}{\cal S}\qquad\Leftrightarrow\qquad
\Delta\;\exp\Big\{\frac{i}{\hbar}{\cal S}\Big\}=0
\end{eqnarray}
for a bosonic functional
${\cal S}={\cal S}(\varphi,\varphi^*)$ satisfying the
boundary condition
\begin{eqnarray}
\label{BoundCon} {\cal S}|_{\varphi^* = \hbar = 0}= S_0({\cal A})
\end{eqnarray}
and being the basic object of the field-antifield quantization
scheme \cite{BV,BV1}. Among the properties of the antibracket and
$\Delta$-operator we mention the Leibniz rule,
\beq
(F, GH)=(F,G)H+(F, H)G(-1)^{\varepsilon(G)\varepsilon(H)},
\eeq
the Jacobi identity,
\beq
\label{JI}
((F, G),
H)(-1)^{(\varepsilon(F)+1)(\varepsilon(H)+1)} +{\sf cycle} (F, G,
H)\equiv 0,
\eeq
and the $\Delta$-operator being a derivative to the
antibracket,
\beq
\label{DAnt}
\Delta (F,G)=(\Delta F,G)-(F,\Delta
G)(-1)^{\varepsilon(F)}.
\eeq
There exists a generating functional
$Y=Y(\varphi,\Phi^*),\;\varepsilon(Y)=1$ of the anticanonical
transformation, \beq \label{antican}
\Phi^i=\pa_{\Phi^*_i}Y(\varphi,\Phi^*),\quad \varphi^*_i=
Y(\varphi,\Phi^*)\overleftarrow{\pa}_{\varphi^i}. \eeq The
invariance property of the odd symplectic structure (\ref{defAB}) on
the phase space of $(\varphi,\varphi^*)$ is dual to the invariance
property of an even symplectic structure (a Poisson bracket) under
a canonical transformation of canonical variables $(p,q)$ (for
further discussions of relations between Poisson bracket and
antibracket, see \cite{BH,BM}).
The generating functional of the Green functions $Z(J)$ is defined in
terms of the functional integral as \cite{BV,BV1}
\begin{eqnarray}
\label{ZBV} {\cal Z}(J)=\int {\cal D}\varphi\;
\exp\left\{\frac{i}{\hbar}\left[S_{e}(\varphi)+ J_i\varphi^i\right]\right\}
=\exp\left\{\frac{i}{\hbar}W(J)\right\},
\end{eqnarray}
where
\beq
S_{e}(\varphi) = {\cal S}\big(\varphi,\varphi^*=\pa_{\varphi}\psi(\varphi)\big),
\eeq
$\psi(\varphi)$ is a fermionic gauge
functional, $J_i$
$(\varepsilon(J_i) = \varepsilon_i)$ are usual external sources
to the fields $\varphi^i$ and $W(J)$ is the generating functional
of the connected Green functions.
To discuss the quantum properties of general
gauge theories, it is useful to consider, instead of the generating functional
(\ref{ZBV}), the extended generating functionals
${\cal Z}(J,\varphi^*)$ and $W(J,\varphi^*)$
defined by the relations
\beq \label{ZBVEx}
{\cal Z}(J,\varphi^*)=\int {\cal D}\varphi\;
\exp\Big\{\frac{i}{\hbar}\big[S(\varphi,\varphi^*)+J_i\varphi^i\big]\Big\}=
\exp\left\{\frac{i}{\hbar}W(J,\varphi^*)\right\}
\eeq
where
\beq
\label{AcBV}
S(\varphi,\varphi^*)={\cal S}\big(\varphi,\varphi^*+
\pa_{\varphi}\psi(\varphi)\big).
\eeq
Obviously, we have
\beq {\cal Z}(J)={\cal Z}(J,\varphi^*)\mid_{\varphi^*=0},
\quad W(J)=W(J,\varphi^*)\mid_{\varphi^*=0}.
\eeq
The action $S=S(\varphi,\varphi^*)$
(\ref{AcBV}) satisfies the quantum master equation
\beq
\label{MEext} \frac{1}{2}(S,S)=i\hbar \Delta
S\qquad\Leftrightarrow\qquad \Delta
\exp\Big\{\frac{i}{\hbar}S\Big\}=0.
\eeq
It follows from (\ref{MEext}) that the Ward identities hold for the extended
generating functionals ${\cal Z}(J,\varphi^*)$ and $W(J,\varphi^*)$
\beq \label{WIZ}
J_i\pa_{\varphi^*_i}{\cal Z}(J,\varphi^*)=0\;,\quad
J_i\pa_{\varphi^*_i}W(J,\varphi^*)=0.
\eeq
Indeed, we have
\beq
\nonumber
&&0=\int d\varphi \exp\left\{\frac{i}{\hbar}J\varphi\right\}
\left(\Delta \exp\left\{\frac{i}
{\hbar}S\right\}\right)=\int d\varphi(-1)^{\varepsilon_i}
\partial_{\varphi^{i}}\left[
\exp\left\{\frac{i}{\hbar}J\varphi\right\}\partial_{\varphi_{i}^{\ast}}
\exp\left\{\frac{i}{\hbar}
S\right\}\right] -\\
\nonumber
&&-\frac{i}{\hbar}J_{i}\partial_{\varphi_{i}^{\ast}}\int d\varphi
\exp\left\{\frac{i}{\hbar}(S+J\varphi)\right\}=
-\frac{i}{\hbar}J_{i}\partial_{\varphi
_{i}^{\ast}}{\cal Z}(J,\varphi^*)=\\
\label{WIW}
&&=-\frac{i}{\hbar}J_{i}\partial_{\varphi_{i}^{\ast}}
\exp\left\{\frac{i}{\hbar}W(\varphi^{\ast},J)\right\}
\Longrightarrow
J_{i}\partial_{\varphi_{i}^{\ast}}W(\varphi^{\ast},J)=0.
\eeq
The generating functional of vertex function
(effective action) is defined via the Legendre transformation
\beq
\Gamma(\varphi,\varphi^*)=W(J,\varphi^*)-J\varphi,\;
\varphi^i=\pa_{J_i} W(J,\varphi^*),\;
\partial_{\varphi_{i}^{\ast}}W(J,\varphi^{\ast})=
\partial_{\varphi_{i}^{\ast}}\Gamma(\varphi,\varphi
^{\ast}),\; \pa_{J_i}=\frac{\pa}{\pa J_i}
\eeq
with the properties
\beq
\label{prG}
J_{i}=-\Gamma(\varphi,\varphi^{\ast})\overleftarrow{\partial}_{\varphi^{i}
}=-(-1)^{\varepsilon_i}\Gamma_{i},\;\Gamma_{i}=\Gamma_{i}(\varphi,\varphi^{\ast}
)=\partial_{\varphi^{i}}\Gamma(\varphi,\varphi^{\ast})\:.
\eeq
It follows from (\ref{WIW}) and (\ref{prG}) that the Ward
identity for the effective action holds,
\beq
\label{WIG}
\Gamma\overleftarrow{\partial}_{\varphi^{i}}\partial_{\varphi_{i}^{\ast}%
}\Gamma=0\;\Longrightarrow\frac{1}{2}(\Gamma,\Gamma)=0,
\eeq
which has the form of classical master equation in the field-antifield formalism.
As it was pointed out for the first time in \cite{VLT}, the gauge
fixing procedure in the field-antifield formalism
(\ref{AcBV}) can be described in terms of a special type
of anticanonical transformation (\ref{antican}).
Indeed, let us consider anticanonical transformations of the variables
($\varphi,\varphi^*$) with specific generating function
\beq
\label{santican}
Y=Y(\varphi,\Phi^*)=\Phi^*_i\varphi^i-\psi(\varphi).
\eeq
We have
\beq
\Phi^i=\varphi^i,\quad \varphi^*_i=\Phi^*_i-\pa_{\varphi^i}\psi(\varphi),
\eeq
so that the transformed action ${\tilde S}={\tilde S}(\varphi,\varphi^*)$
\beq
{\tilde S}(\varphi,\varphi^*)=S(\Phi,\Phi^*)=S(\varphi,\varphi^*+
\pa_{\varphi}\psi(\varphi))
\eeq
coincides with (\ref{AcBV}). In particular, this fact made it possible
to study effectively the
gauge dependence and structure of renormalization of general
gauge theories \cite{VLT}.
In what follows we explore a gauge fixing procedure in the field-antifield
formalism as an
anticanonical transformation of general type with the only
requirement for anticanonically
generalized action:
the supermatrix of the second field derivatives of this action must be non-degenerate.
An essential difference in this point with the approach used in \cite{VLT} is
that we work with a general setting for an action (\ref{AcBV}) which
satisfies the quantum master equation (not the classical master
equation as in \cite{VLT}).
\\
\section{Infinitesimal anticanonical transformations}
As the first step in our study of anticanonical transformations in
the field-antifield formalism, we consider the properties of the main
objects subjected to infinitesimal anticanonical transformations. In
the latter case, the generating functional $Y$ reads
\beq
\label{infantican}
Y=Y(\varphi,\Phi^*)=\Phi^*_i\varphi^i+X(\varphi,\Phi^*), \quad
\varepsilon(X)=1.
\eeq
The functional $X$ is considered as the
infinitesimal one. Then the anticanonical transformations of the variables,
\beq \Phi^i=\varphi^i+\pa_{\Phi^*_i}X(\varphi,\Phi^*),\quad
\varphi^*_i=\Phi^*_i+\pa_{\varphi^i}X(\varphi,\Phi^*),
\eeq
can be written down to the first order in $X$ as
\beq
\Phi^i=\varphi^i+\pa_{\varphi^*_i}X(\varphi,\varphi^*)+O(X^2),\quad
\Phi^*_i=\varphi^*_i-\pa_{\varphi^i}X(\varphi,\varphi^*)+O(X^2)
\eeq
or, in terms of the antibracket (\ref{defAB}),
\beq
\Phi^i=\varphi^i+(\varphi^i,X)+O(X^2),\quad
\Phi^*_i=\varphi^*_i+(\varphi^*_i,X)+O(X^2). \eeq The
anticanonically transformed action ${\tilde S}$,
\beq
\label{trnS}
{\tilde S}={\tilde
S}(\varphi,\varphi^*)=S(\Phi,\Phi^*)=S+(S,X)+O(X^2)
\eeq
does not
satisfy the quantum master equation to the first approximation in $X$,
\beq
\frac{1}{2}({\tilde S},{\tilde S})-i\hbar \Delta {\tilde
S}=i\hbar(S,\Delta X)+O(X^2)\neq 0.
\eeq
Consider now the
superdeterminant of the anticanonical transformation
\beq
\label{Jacob} {\cal J}(\varphi,\varphi^*)={\cal J}(Z)=\sDet
\left[{\bar Z}^A(Z) \overleftarrow{\pa}_B\right],
\eeq
where
\beq
{\bar Z}^A=(\Phi^i,\Phi^*_i),\quad Z^A=(\varphi^i,\varphi^*_i),\quad
\pa_A=\frac{\pa}{\pa Z^A}.
\eeq
To the first-order approximation in
$X$, the ${\cal J}$ reads
\beq
{\cal J}=\exp\{2\Delta X\}+O(X^2)=
\exp\left\{\frac{i}{\hbar}\big(-2i\hbar\Delta X\big)\right\}+O(X^2).
\eeq
In contrast to the notation used in \cite{LT,BV2}, now we refer to
$S'=S'(\varphi,\varphi^*)$ constructed from $S=S(\varphi,\varphi^*)$
via the {\it anticanonical master transformation},
\beq
\label{E3.10}
S'&=&S'(\varphi,\varphi^*)=
S(\Phi(\varphi,\varphi^*),\Phi^*(\varphi,\varphi^*))-
i\hbar\frac{1}{2}\ln {\cal J}(\varphi,\varphi^*), \label{trAc}
\eeq
as the {\it master-transformed action}.
Note that, by itself, the anticanonical master transformation can be defined
without reference on
solutions of the quantum master equation. Namely, let us define a transformation
of the form\footnote{In the present article, we only consider the case
in which the generator $F$ is a function; the case of
operator-valued $F$ it was studied in \cite{BB1}.}
\beq
\exp\left\{\frac{i}{\hbar}G'\right\}=
\exp\{-[F,\Delta]\}\exp\left\{\frac{i}{\hbar}G\right\},
\eeq
where $G, F$ ($\varepsilon(G)=0,\;\varepsilon(F)=1$)
are arbitrary functions of $\varphi,\varphi^*$, and $[\;,\;]$
stands for the supercommutator.
Then we can prove
(see Appendices C and D) the relation
\beq
\label{E3.12}
G'=\exp\{{\rm ad}(F)\}G+i\hbar f({\rm ad}(F))\Delta F,
\; f({\rm ad}(F))\Delta F=-\frac{1}{2}\ln {\cal J},\; {\rm ad}(F)(...)=(F, (...)),
\eeq
which repeats the relation (\ref{E3.10}). In (\ref{E3.12}) the notation
$f(x)=(\exp x -1)x^{-1}$ is used.
The action $S'$ (\ref{trAc})
to the first order in $X$
\beq
S'=S+(S,X)-i\hbar\Delta X+O(X^2)
\eeq
does satisfy the quantum master equation
\beq \label{QMEStr}
\frac{1}{2}(S',S')-i\hbar\Delta S'=O(X^2).
\eeq
Note that, due to the
results of \cite{LT,BV2}, the action (\ref{trAc}) by itself
satisfies the quantum master equation in the case of arbitrary
anticanonical transformation, as well (see, also \cite{BLT4,BBD}).
Let us consider the generating functionals constructed
with the help of master-transformed action $S^{\prime}$ to the first order
in $X$. We have
\beq
\nonumber &&{\cal Z}'
={\cal Z}'(J,\varphi^{\ast})=\int d\varphi
\exp\left\{\frac{i}{\hbar}(S'+J\varphi)\right\}=
\exp\left\{\frac{i}{\hbar}W^{\prime}(J,\varphi^{\ast})\right\}=\\
&&=\exp\left\{\frac{i}{\hbar}W(J,\varphi^{\ast})
\right\}\left(1+\frac{i}{\hbar}\delta
W(J,\varphi^{\ast})\right)\!,\\
&&\Gamma'(\varphi,\varphi^{\ast}) =W'(J,\varphi^{\ast})
-J\varphi=\Gamma(\varphi,\varphi^{\ast})+\delta\Gamma(\varphi
,\varphi^{\ast}),\\
\nonumber
&&\delta\Gamma(\varphi,\varphi^{\ast})=
\delta W(J(\varphi,\varphi^{\ast}),\varphi^*).
\eeq
Therefore
\beq
\nonumber
&&{\cal Z}'-{\cal Z}=\delta {\cal Z}=
\frac{i}{\hbar}\exp\left\{\frac{i}{\hbar}W(J,\varphi^{\ast})\right\}
\delta
W(J,\varphi^{\ast})=\\
\nonumber
&&=\frac{i}{\hbar}\exp\left\{\frac{i}{\hbar}W(J,\varphi^{\ast})\right\}
\delta\Gamma(\varphi,\varphi^{\ast})=\nonumber\\
\nonumber
&&=\frac{i}{\hbar}\int d\varphi\left[ (S,X)-i\hbar\Delta X\right]
\exp\left\{\frac{i}{\hbar}(S+J\varphi)\right\}=\\
\nonumber
&&=\int d\varphi \exp\left\{\frac{i}{\hbar}J\varphi
\right\}\Delta\left( X\exp\left\{\frac{i}{\hbar}S\right\}\right) =\\
\nonumber
&& =-\frac{i}{\hbar}J_{i}\partial_{\varphi_{i}^{\ast}}\left[ \tilde
{X}(J,\varphi^{\ast})\exp\left\{\frac{i}{\hbar}W(J,\varphi^{\ast})\right\}\right]=\\
\label{delZ}
&&=\exp\left\{\frac{i}{\hbar}W(J,\varphi^{\ast})\right\}\left[-\frac{i}{\hbar}J_{i}
\partial_{\varphi_{i}^{\ast}}\tilde{X}(J,\varphi^{\ast})\right],
\eeq
where
\beq
\tilde{X}(J,\varphi^{\ast})=
\exp\left\{-\frac{i}{\hbar}W(J,\varphi^{\ast})\right\}\int
d\varphi X\exp\left\{\frac{i}{\hbar}(S+J\varphi)\right\}.
\eeq
When deriving (\ref{delZ}), the Ward identity for $W(J,\varphi^*)$
(\ref{WIW}), the quantum master equation for $S(\varphi,\varphi^*)$
(\ref{MEext}) and the following identities:
\beq
\label{usID}
i\hbar\exp\left\{\frac{i}{\hbar}S\right\}\Delta X=
i\hbar\Delta\left( X\exp\left\{\frac{i}{\hbar} S\right\}\right)
+(S,X)\exp\left\{\frac{i}{\hbar}S\right\}, \eeq \beq \nonumber
&&\exp\left\{\frac{i}{\hbar}J\varphi\right\}
\Delta\left(X\exp\left\{\frac{i}{\hbar}S\right\}\right)
=(-1)^{\varepsilon_i}\partial_{\varphi^{i}}
\left[\exp\left\{\frac{i}{\hbar}J\varphi\right\}
\partial_{\varphi_{i}^{\ast}}\left( Xe^{\frac{i}{\hbar}S}\right) \right]-\\
\label{usID1}
&&\qquad-\frac{i}{\hbar}J_{i}\partial_{\varphi_{i}^{\ast}}
\left( X\exp\left\{\frac{i}{\hbar
}(S+J\varphi)\right\}\right)
\eeq
are used. Rewriting (\ref{delZ}) for a variation of the effective action
$\Gamma=\Gamma(\varphi,\varphi^*)$, we obtain
\beq
\nonumber
&&\delta\Gamma(\varphi,\varphi^{\ast})=-J_{i}\partial_{\varphi_{i}^{\ast}}
\tilde{X}(J,\varphi^{\ast})=(-1)^{\varepsilon_i}
\Gamma_{i}\partial_{\varphi_{i}^{\ast}}
\mathcal{X}(\varphi,\varphi^{\ast})-\\
\label{delGamma}
&&\qquad -(-1)^{\varepsilon_i}\Gamma_{i}\left[\partial_{\varphi_{i}^{\ast}}J_{j}
(\varphi,\varphi^{\ast})\right] \left. \partial_{J_{j}}\tilde{X}
(J,\varphi^{\ast})\right| _{J=J(\varphi,\varphi^{\ast})},
\eeq
where
\beq
\mathcal{X}(\varphi,\varphi^{\ast})=
\left.\tilde{X}(J,\varphi^{\ast})\right|_{J=J(\varphi,\varphi^{\ast})}.
\eeq
One can rewrite Eq. (\ref{delGamma}) in terms of
$\Gamma=\Gamma(\varphi,\varphi^*)$ as
\begin{equation}
\delta\Gamma(\varphi,\varphi^{\ast})=\Gamma\overleftarrow{\partial}
_{\varphi^{i}}\partial_{\varphi_{i}^{\ast}}\mathcal{X}-\Gamma\overleftarrow
{\partial}_{\varphi_{i}^{\ast}}\partial_{\varphi^{i}}\mathcal{X}
=(\Gamma,\mathcal{X})=-(\mathcal{X},\Gamma). \label{anticGamma}
\end{equation}
This result is proved in Appendix A.
Equation (\ref{anticGamma}) means that any infinitesimal anticanonical master
transformation of the action $S$ (\ref{trnS})
with a generating functional $X$ induces an infinitesimal anticanonical
transformation in
the effective action $\Gamma$ (\ref{anticGamma}) with a generating functional
$\mathcal{X}$, provided
the generating functional of the Green functions is constructed via
the master-transformed action. An important goal of our present study
is a generalization
of this fact (for the first time known among results of paper \cite{VLT})
to the case of arbitrary (finite) anticanonical transformation.
\\
\section{Finite anticanonical transformation}
Consider an arbitrary (finite) anticanonical transformation described
by a generating functional $Y=Y(\varphi,\Phi^*),\;\varepsilon(Y)=1$,
\footnote{Note that any anticanonical transformation can be described
in terms of a generating functional.}
\beq
\label{acY}
\varphi_{i}^{\ast}=Y(\varphi,\Phi^{\ast})\overleftarrow{\partial}_{\varphi^i}
,\quad\Phi^{A}=\pa_{\Phi^{\ast}_i}Y(\varphi,\Phi^{\ast}).
\eeq
Let $Y$ have the form
\beq
\label{Ya}
Y(\varphi,\Phi^*)=\Phi_{i}^{\ast}\varphi^{i}+a f(\varphi
,\Phi^{\ast}), \quad \varepsilon(f(\varphi,\Phi^{\ast}))=1,
\eeq
where $a$ is a parameter. Then solution of equations (\ref{acY})
up to second order in $a$ can be written as
\beq
\label{Phia}
&&\Phi^{i}=\varphi^{i}+a f\overleftarrow{\partial}_{\varphi^{\ast}_i}-
a^{2}(-1)^{(\varepsilon_{i}+1)(\varepsilon_{j}+1)}f\overleftarrow
{\partial}_{\varphi^{\ast}_j}
\overleftarrow{\partial}_{\varphi^{\ast}_i}\overrightarrow
{\partial}_{\varphi^j}f+O(a^{3}),\\
\label{Phi*a} &&
\Phi_{i}^{\ast}=\varphi_{i}^{\ast}-a\partial_{\varphi^i}f+a^{2}
(-1)^{\varepsilon_{i}(\varepsilon_{j}+1)}
f\overleftarrow{\partial}_{\varphi^{\ast}_j}
\overleftarrow{\partial}_{\varphi^i}
\overrightarrow{\partial}_{\varphi^j}f+O(a^{3}),
\eeq
where $f\equiv f(\varphi,\varphi^{\ast})$. Let us denote
\beq
Z^{A}=\{\varphi^{i},\varphi_{i}^{\ast}\},\quad
{\bar Z}^{A}=\{\Phi^{i},\Phi_{i}^{\ast}\},\quad
\varepsilon({\bar Z}^A)=\varepsilon(Z^A)= \varepsilon_A,
\eeq
and
\beq
\label{Fsc2}
F=F(\varphi,\varphi^{\ast};a)=-f(\varphi,\varphi^{\ast})+\frac{a}
{2}f(\varphi,\varphi^{\ast})\overleftarrow{\partial}_{\varphi^{\ast}_j}
\overrightarrow{\partial}_{\varphi^j}f(\varphi,\varphi^{\ast}).
\eeq
Then we have
\beq
\label{ZFa}
{\bar Z}^{A}={\bar Z}^{A}(Z;a)=\exp\{a{\rm ad}(F)\}Z^{A}+O(a^{3}),
\eeq
where ${\rm ad}(F)$ means the left adjoint of the antibracket
\beq \label{Fhat}
{\rm ad}(F)(...)=(F(Z;a),(...)).
\eeq
We call $F$ in (\ref{Fsc2}) a
generator of the anticanonical transformation to the second order. It
should be noticed that
the generator of an anticanonical transformation does
not coincide with the generating functional of this transformation
already to the second order.
A natural question arises: Does a generator exist for a given anticanonical
transformation, actually? To answer this question,
we begin with the claim
that an operator $\exp\{{\rm ad} (F)\}$ generates
anticanonical transformation. Indeed, let $Z^A$ be anticanonical
variables so that the antibracket (\ref{defAB}) can be presented in
the form
\beq
\label{ABz}
(H(Z),G(Z))=H(Z)\overleftarrow{\partial}_{A}E^{AB}\overrightarrow{\partial}
_{B}G(Z),\quad
(Z^{A},Z^{B})=E^{AB},\quad\partial_{A}=\frac{\partial}{\partial Z^{A}},
\eeq
where $E^{AB}$ is a constant supermatrix with the properties
\beq
E^{BA}=-(-1)^{(\varepsilon_A+1)(\varepsilon_B+1)}E^{AB},\quad
\varepsilon(E^{AB})=\varepsilon_A+\varepsilon_B+1.
\eeq
Then the transformation
\beq
\label{trzZ}
Z^{A}\rightarrow
{\bar Z}^{A}(Z)=\exp\{{\rm ad}F(Z)\}Z^{A}
\eeq
is anticanonical,
\beq
({\bar Z}^{A}(Z),{\bar Z}^{B}(Z))={\bar Z}^{A}(Z)\overleftarrow{\partial}_{C}
E^{CD}\overrightarrow
{\partial}_{D}{\bar Z}^{B}(Z)=E^{AB}.
\eeq
To prove this fact we introduce
an one-parameter family of transformations
\beq
\label{Zi}
{\bar Z}^A(Z,a)=\exp\{a{\rm ad}( F)\}Z^A, \quad {\bar Z}^A(Z,0)=Z^A,
\eeq
and quantities ${\bar Z}^{AB}(Z,a)$,
\beq
\label{Zij}
{\bar Z}^{AB}(Z,a)=({\bar Z}^A(Z,a),{\bar Z}^B(Z,a)),
\quad {\bar Z}^{AB}(Z,0)=E^{AB}.
\eeq
It follows from the definitions (\ref{Zi}) and (\ref{Zij}), that the
relations
\beq
&&\frac{d}{da}{\bar Z}^A(Z,a)=(F(Z),{\bar Z}^A(Z,a)),\\
\nonumber
&&\frac{d}{da}{\bar Z}^{AB}(Z,a)=((F(Z),{\bar Z}^A(Z,a)),{\bar Z}^B(Z,a))
+({\bar Z}^A(Z,a),(F(Z),{\bar Z}^B(Z,a))=\\
\label{eqZij}
&&=(F(Z),({\bar Z}^A(Z,a),{\bar Z}^B(Z,a))=
(F(Z),{\bar Z}^{AB}(Z,a))={\rm ad}( F(Z)){\bar Z}^{AB}(Z,a),
\eeq
hold,
where the Jacobi identity (\ref{JI}) for antibrackets is used.
A solution to Eq. (\ref{eqZij}) has the form
\beq
{\bar Z}^{AB}(Z,a)=\exp\{a{\rm ad}( F(Z))\}{\bar Z}^{AB}(Z,0)=
\exp\{a{\rm ad}( F(Z))\}E^{AB}=E^{AB},
\eeq
and the transformation (\ref{trzZ}) is really anticanonical.
The inverse to this statement
is valid as well: an arbitrary set of anticanonical variables ${\bar Z}^A(Z)$
can be presented in the form
\beq
\label{genF}
{\bar Z}^A(Z)=\exp\{{\rm ad}( F(Z))\}Z^A
\eeq
with some generator functional $F(Z),\;\varepsilon(F(Z))=1 $. In Appendix B,
a proof of this fact is given.
Consider now a master-transformed action
$S'=S'(\varphi,\varphi^*)$ (\ref{trAc}). It was pointed out in
\cite{BLT-BV} that there are presentations of $S'$ in the following
forms
\beq
\label{Sprime1}
\exp\left\{\frac{i}{\hbar}S'\right\}=
\exp\{-[F,\Delta]\}\exp\left\{\frac{i}{\hbar}S\right\},
\eeq or
\beq
\label{Sprime2} S'=\exp\{{\rm ad}(
F)\}S+ i\hbar f({\rm ad}(F))\Delta F,
\eeq
where $S=S(\varphi,\varphi^{*})$, and $F=F(\varphi,\varphi^*)$
is a generator functional of an anticanonical
transformation, $f(x)=(\exp(x)-1)x^{-1}$. In accordance with
(\ref{trAc}), the first term in the right-hand side in
(\ref{Sprime2}) describes an anticanonical transformation of $S$ with an
odd generator functional $F$, while the second term is a half of a
logarithm of the Jacobian (\ref{Jacob}) of that transformation, up to
$(-i\hbar)$. In Appendix D, we give a proof of the latter statement.
Now we are in a position to study the properties of generating
functionals of Green functions
subjected to an arbitrary anticanonical transformation. We start with the
generating functionals of Green and connected Green functions
\beq
\label{ZWprime}
{\cal Z}'={\cal Z}'(J,\varphi^{\ast})=\int d\varphi \exp\left\{\frac{i}{\hbar
}(S^{\prime}(\varphi,\varphi^{\ast})+J\varphi)\right\}=
\exp\left\{\frac{i}{\hbar}W^{\prime
}(J,\varphi^{\ast})\right\},
\eeq
where $S'$ is defined in (\ref{Sprime1}). The constructed generating functionals
(\ref{ZWprime}) obey the very important property of independence of $F$ for physical
quantities on-shell.\footnote{Note that in gauge theories the "on-shell"
includes a definition of the physical state space.} Indeed, for infinitesimal
$\delta F$ the variation of ${\cal Z}'$
(\ref{ZWprime}),
\beq
\nonumber
&&\delta {\cal Z}'=-\frac{i}{\hbar}\int d\varphi
[(S,\delta F)-i\hbar(\Delta\delta F)]
\exp\left\{\frac{i}{\hbar
}(S(\varphi,\varphi^{\ast})+J\varphi)\right\}=\\
\label{infZprime}
&&=\frac{i}{\hbar}J_{A}\partial_{\varphi_{A}^{\ast}}\left[
\delta{\tilde F}(J,\varphi^{\ast})
\exp\left\{\frac{i}{\hbar}W(J,\varphi^{\ast})\right\}\right],
\eeq
is proportional to the external sources $J$. Due to the equivalence theorem
\cite{KT}, it means that the Green functions calculated with the help of
the generating functionals ${\cal Z}(J,\varphi^*)$ and ${\cal Z}'(J,\varphi^*)$
give the same physical answers on-shell.
In deriving (\ref{infZprime}), the result
of calculation (\ref{delZ}) is used and the notation
\beq
\delta
{\tilde F}(J,\varphi^{\ast})=Z^{-1}(J,\varphi^*)\int d\varphi\;
\delta F(\varphi,\varphi^*) \exp\left\{\frac{i}{\hbar
}(S(\varphi,\varphi^{\ast})+J\varphi)\right\}
\eeq
is introduced.
In the case of finite anticanonical transformations, we consider the
following anticanonically generalized action $S''$
\beq
\label{Sprime3}
\exp\left\{\frac{i}{\hbar}S''(\varphi,\varphi^{\ast})\right\}=
\exp\{-[F(\varphi ,\varphi^{\ast})+\delta
F(\varphi,\varphi^{\ast}),\Delta]\}
\exp\left\{\frac{i}{\hbar}S(\varphi,\varphi^{\ast})\right\},
\eeq
where $\delta F=\delta F(\varphi,\varphi^*)$ is an infinitesimal
functional. The following representation holds:
\beq
\label{Fd}
\exp\{-[F(\varphi ,\varphi^{\ast})+\delta
F(\varphi,\varphi^{\ast}),\Delta]\}= \exp\{-[\delta{\cal
F}(\varphi,\varphi^{\ast}),\Delta]\}
\exp\{-[F(\varphi,\varphi^{\ast}),\Delta]\},
\eeq
where $\delta{\cal F}(\varphi,\varphi^{\ast})$ is defined by the relation
\beq
\label{dFc} \exp\{-{\rm ad}( F(\varphi,\varphi^{\ast}))-{\rm ad}
(\delta F(\varphi,\varphi^{\ast}))\} \exp\{-{\rm ad}
(F(\varphi,\varphi^{\ast}))\}= \exp\{-{\rm ad}(\delta {\cal
F}(\varphi,\varphi^{\ast}))\}.
\eeq
In Appendix C, a proof of
Eqs. (\ref{Fd}) and (\ref{dFc}) is given. Due to ({\ref{Fd}}), we
can present the action $S''$ in the form
\beq \label{S2S1}
\exp\left\{\frac{i}{\hbar}S''(\varphi,\varphi^{\ast})\right\}=
\exp\{-[\delta{\cal F}(\varphi,\varphi^{\ast}),\Delta]\}
\exp\left\{\frac{i}{\hbar}S'(\varphi,\varphi^{\ast})\right\}.
\eeq
Although we need here the infinitesimal functional
$\delta{\cal F}(\varphi,\varphi^{\ast})$,
the representation (\ref{S2S1}) by itself is valid
for arbitrary functional $\delta{\cal F}$.
In turn, the representation (\ref{S2S1}) allows us the use of the previous
arguments concerning the case of infinitesimal anticanonical
transformations and for the statement that the generating functionals ${\cal Z}''$
and ${\cal Z}'$ constructed with the help of the actions $S''$ and $S'$,
respectively, give the same physical results.
The next point of our study is connected with the behavior
of generating functionals
subjected to an arbitrary anticanonical transformation.
Consider a one-parameter family of functionals ${\cal Z}'(J,\varphi^{\ast};a)$,
\beq
&& \!\!\!\! {\cal Z}'(a)={\cal Z}^{\prime}(J,\varphi^{\ast};a)=\!
\int d\varphi \exp\left\{\frac{i}
{\hbar}(S'(\varphi,\varphi^{\ast};a)+J\varphi)\right\}=
\exp\left\{\frac{i}{\hbar}W^{\prime}(J,\varphi^{\ast};a)\right\}\!\!,\\
\label{Sprimea3}
&& \exp\left\{\frac{i}{\hbar}S'(\varphi,\varphi^{\ast};a)\right\}=
\exp\{-a[F(\varphi,\varphi^{\ast}),\Delta]\}
\exp\left\{\frac{i}{\hbar}S(\varphi,\varphi^{\ast})\right\},
\eeq
so that
\beq
{\cal Z}'(1)={\cal Z}'.
\eeq
Taking into account (\ref{delZ}) and (\ref{Sprimea3}), we derive the relation
\beq
\nonumber
&& \partial_{a}{\cal Z}'(a)=\frac{i}{\hbar}{\cal Z}'(a)\partial
_{a}W'(J,\varphi^{\ast};a)=\frac{i}{\hbar}{\cal Z}'(a)\partial
_{a}\Gamma(\varphi,\varphi^{\ast};a)=\\
\nonumber
&& =-\int d\varphi \exp\left\{\frac{i}{\hbar}J\varphi\right\}
[F(\varphi,\varphi^{\ast}),\Delta]
\exp\left\{\frac{i}{\hbar}S'(\varphi,\varphi^{\ast};a)\right\}=\\
\label{derZa}
&&=-\int d\varphi\exp\left\{\frac{i}{\hbar}J\varphi\right\}\Delta\left(F(\varphi
,\varphi^{\ast})\exp\left\{\frac{i}{\hbar}S'(\varphi,\varphi^{\ast}
;a)\right\}\right).
\eeq
By repeating similar calculations which lead us from (\ref{delZ})
to (\ref{anticGamma}) due to Eqs. (\ref{usID}), (\ref{usID1})
and (\ref{usR1})-(\ref{usRs}), we obtain
\beq
&&\partial_{a}\Gamma(\varphi,\varphi^{\ast};a)=(\mathcal{F}(\varphi
,\varphi^{\ast};a),\Gamma(\varphi,\varphi^{\ast};a)),\label{derfGam}\\
\label{calF}
&&\mathcal{F}(\varphi,\varphi^{\ast};a) =\left. \frac{1}{{\cal Z}'
(J,\varphi^{\ast};a)}\int d{\tilde\varphi}
F({\tilde \varphi},\varphi^{\ast})\exp\left\{\frac{i}{\hbar
}(S'({\tilde \varphi},\varphi^{\ast};a)+
J{\tilde\varphi})\right\}\right|_{J=J(\varphi,\varphi^{\ast};a)}.
\eeq
We will refer to (\ref{derfGam}) as the basic equation describing dependence
of effective action
on an anticanonical transformation in the field-antifield formalism.
In Sect. 5, we present a
solution to this equation.
\\
\section{Solution to the basic equation}
In what follows below, we will use a short notation for all quantities depending
on the variables $\varphi,\varphi^*$,
\beq
\label{Eq.5.1}
\Gamma(\varphi,\varphi^*;a)\equiv \Gamma(a),\quad
\Gamma(\varphi,\varphi^*)\equiv \Gamma,\quad
{\cal F}(\varphi,\varphi^*;a)\equiv {\cal F}(a)
\eeq
and so on. Then the basic equation (\ref{derfGam}) is written as
\beq
\label{Eq.5.2}
\pa_a \Gamma(a)=({\cal F}(a),\Gamma(a))
={\rm ad}({\cal F}(a))\Gamma(a).
\eeq
We will study solutions to (\ref{Eq.5.2}) in the class
of regular functionals in $a$, by using
a power series expansion in this parameter. In the beginning,
let us find a solution to this equation
to the first order in $a$, presenting $\Gamma(a)$ and ${\cal F}(a)$ in the form
\beq
&& \Gamma_{1}(a)\equiv\Gamma(a)=\Gamma+a\Gamma_{1|1}+O(a^{2}),\\
&& \mathcal{F}_{1}(a)\equiv\mathcal{F}(a)=\frac{1}{a}\mathcal{F}_{1|1}(a)+O(a),
\;\mathcal{F}_{1|1}(a)=a\mathcal{F}_{1|1}.
\eeq
A straightforward calculation yields the following result
\beq
\Gamma_{1|1}=(\mathcal{F}_{1|1},\Gamma)={\rm ad}({\cal F}_{1|1})\Gamma.
\eeq
Introduce the notation $U_{1}(a)=\mathcal{F}_{1|1}(a)=
a\mathcal{F}_{1|1}$ and the functional $\Gamma_2(a)$ by the rule
\beq
\label{G2}
\Gamma_{2}(a)=\exp\{-{\rm ad}({U}_{1}(a))\}\Gamma_{1}(a).
\eeq
The dependence of $\Gamma_2(a)$ on $a$ is described by the equation
\beq
\label{EqG2}
\partial_{a}\Gamma_{2}(a)=(\mathcal{F}_{2}
(a),\Gamma_{2}(a))
\eeq
where
\beq
\label{F2}
\mathcal{F}_{2}(a)=\left[\exp\{-a{\rm ad}(\mathcal{F}_{1|1})\}
\mathcal{F}_{1}(a)-\mathcal{F}_{1|1}\right].
\eeq
It follows from (\ref{G2}) that the
functional $\Gamma_2(a)$ coincides with $\Gamma$ up to the second
order in $a$,
\beq
\Gamma_{2}(a)=\Gamma+O(a^{2})=\Gamma+a^{2}\Gamma_{2|2}+O(a^{3}).
\eeq
In turn, the functional ${\cal F}_2(a)$ vanishes to the first
order in $a$
\beq
\mathcal{F}_{2}(a)=O(a)=\frac{2}{a}\mathcal{F}_{2|2}(a)+O(a^{2}),
\;\mathcal{F}_{2|2}(a)=a^{2}\mathcal{F}_{2|2}.
\eeq
To the second
order in $a$, the solution to (\ref{EqG2}) reads
\beq
\Gamma_{2|2}=(\mathcal{F}_{2|2},\Gamma).
\eeq
Then, the functional
${\tilde \Gamma}_3(a)$ constructed by the rule
\beq
\tilde{\Gamma}_{3}(a)=\exp\{-{\rm ad}(\mathcal{F}
_{2|2}(a))\}\Gamma_{2}(a)
\eeq
coincides with $\Gamma$ up to the third order in $a$
\beq
\tilde{\Gamma}_{3}(a)=\Gamma+O(a^{3}).
\eeq
Introduce the functional $\Gamma_3(a)$
\beq
\Gamma_{3}(a)=\exp\{-{\rm ad}(U_{2}(a))\}\Gamma_1(a),\quad
U_{2}(a)=\mathcal{F}_{1|1}(a)+\mathcal{F}_{2|2}(a).
\eeq
Note that
$\Gamma_3(a)$ coincides with ${\tilde \Gamma}_3(a)$ up to the third
order in $a$
\beq \label{G3ap}
\Gamma_{3}(a)=\tilde{\Gamma}_{3}(a)+O(a^{3})=\Gamma+O(a^{3})=
\Gamma+a^{3}\Gamma_{3|3}+O(a^{4}),
\eeq
so that we have
\beq
\nonumber
&&\Gamma_{3}(a)=\exp\{-{\rm ad}(\mathcal{F}_{2|2}(a))\}
\exp\{-{\rm ad}(\mathcal{F}_{1|1}(a))\}\Gamma_{1}(a)+O(a^3)=\\
&&=
\exp\{-{\rm ad}(\mathcal{F}_{2|2}(a))\} \Gamma_{2}(a)+O(a^3)
\eeq
due to the relation (\ref{Fn}). It follows from (\ref{G3ap}) that
\beq
\partial_{a}\Gamma_{3}(a)=3a^{2}\Gamma_{3|3}+O(a^{3}).
\eeq
On the other hand, we have
\beq
\label{eqG3}
\partial_{a}\Gamma_{3}(a)=(\mathcal{F}_{3}(a),\Gamma_{3}(a))=
{\rm ad}({\cal F}_3(a))\Gamma_3(a),
\eeq
where
\beq
\nonumber
\label{F3}
&&{\rm ad}(\mathcal{F}_{3}(a))=-\exp\{-{\rm ad}(U_{2}(a))\}
\partial_{a}\exp\{{\rm ad}(U_{2}(a))\}+\\
\label{F3}
&&-\exp\{-{\rm ad}(U_{2}(a))\}
{\rm ad}(\mathcal{F}_{1}(a))
\exp\{{\rm ad}(U_2(a))\},
\eeq
the operators on the right-hand side of (\ref{F3}) have certainly the form of
the ones of ${\rm ad}$, see Eqs. (\ref{X}), (\ref{Xt}) and (\ref{ABA}),
(\ref{eAB}). By using (\ref{AnA}), (\ref{Xza}), we derive from (\ref{F3})
and (\ref{eqG3})
\beq
\nonumber
&&{\rm ad}(\mathcal{F}_3(a))=-\frac{2}{a}{\rm ad}(\mathcal{F}_{2|2}(a))+
\exp\{-{\rm ad}(\mathcal{F}_{2|2}(a))\}{\rm ad}(\mathcal{F}_{2}(a))
\exp\{{\rm ad}(\mathcal{F}_{2|2}(a))\}+O(a^{2})=\\
&& =\frac{3}{a}{\rm ad}(\mathcal{F}_{3|3}(\varphi,\varphi^{\ast};a))+O(a^{3}%
),\qquad{\rm ad}(\mathcal{F}_{3|3}(a))=a^{3}{\rm ad}(\mathcal{F}_{3|3}),\\
&&\qquad\qquad\qquad\qquad\qquad \Gamma_{3|3}=(\mathcal{F}_{3|3},\Gamma).
\eeq
Suppose that on the {\it n}th step of our procedure we
have obtained the following relations,
\beq
\nonumber
&& \Gamma_{n}(a)=\exp\{-{\rm ad}(U_{n-1}(a))\}\Gamma_{1}(a)=\Gamma+O(a^{n})=\\
&& =\Gamma+a^{n}\Gamma_{n|n}+O(a^{n+1}),\;
U_{n-1}(a)=\sum_{k=1}^{n-1}\mathcal{F}_{k|k}(a)\equiv\mathcal{F}_{[n-1|n-1]}(a),\\
&& \qquad\qquad\qquad\partial_{a}\Gamma_{n}(a)=(\mathcal{F}_{n}
(a),\Gamma_{n}(a)),\\
&& \mathcal{F}_{n}(a)=O(a^{n})=\frac{n}{a}
\mathcal{F}_{n|n}(a)+O(a^{n+1}),\quad \mathcal{F}_{n|n}(a)=
a^{n}\mathcal{F}_{n|n},\\
&& \qquad\qquad\qquad\qquad \Gamma_{n|n}=(\mathcal{F}_{n|n},\Gamma).
\eeq
We set
\beq
U_{n}(a)=\mathcal{F}_{[n|n]}(a).
\eeq
Then we have
\beq
&& \exp\{-{\rm ad}(\mathcal{F}_{n|n}(a)\}\Gamma_{n}(a)=\Gamma+O(a^{n+1}),\\
\nonumber
&& \Gamma_{n+1}(a)=\exp\{-{\rm ad}(U_{n}(a))\}\Gamma_{1}(a)=
\exp\{-{\rm ad}(\mathcal{F}_{n|n}(a))\}\Gamma_{n}(a) +O(a^{n+1})=\\
&&=\Gamma+O(a^{n+1})=\Gamma+a^{n+1}\Gamma_{n+1|n+1}+O(a^{n+2}).
\eeq
In a similar manner, we derive the equation for $\Gamma_{n+1}(a)$
\beq
\partial_{a}\Gamma_{n+1}(a)=(\mathcal{F}_{n+1}(a),\Gamma_{n+1}(a)),
\eeq
where
\beq
\nonumber
&&{\rm ad}(\mathcal{F}_{n+1}(a))=-\exp\{-{\rm ad}(U_{n}(a))\}
\partial_{a}\exp\{{\rm ad}(U_{n}(a))\}+\\
&&+
\exp\{-{\rm ad}(U_{n}(a))\}{\rm ad}(\mathcal{F}_{1}(a))\exp\{{\rm ad}(U_{n}(a))\}.
\eeq
By the same reasons used at the previous stages, we conclude that
\beq
\nonumber
&&{\rm ad}(\mathcal{F}_{n+1}(a))=-\frac{n}{a}{\rm ad}(\mathcal{F}_{n|n}(a))
+\exp\{-{\rm ad}(\mathcal{F}_{n|n}(a))\}{\rm ad}(\mathcal{F}_{n}(a))
\exp\{{\rm ad}(\mathcal{F}_{n|n}(a))\}+O(a^{n})=\\
&&=\frac{n+1}{a}{\rm ad}(\mathcal{F}_{n+1|n+1}(a))+O(a^{n+1}),\quad
\mathcal{F}_{n+1|n+1}(a)=a^{n+1}\mathcal{F}_{n+1|n+1},\\
&&\qquad\qquad\qquad\qquad \Gamma_{n+1|n+1}=(\mathcal{F}_{n+1|n+1},\Gamma),\\
&&
\Gamma(a)=\exp\{{\rm ad}(U_{n}(a))\}\Gamma_{n+1}(a)=\exp\{{\rm ad}({U}_{n}(a))\}
\Gamma+O(a^{n+1}).
\eeq
Finally, by applying the induction method,
we obtain that a solution to the basic equation (\ref{Eq.5.2}) can
be presented in the form
\beq
\label{canGf}
\Gamma(a)=\exp\{{\rm ad}({U}(a))\}\Gamma,
\eeq
which is nothing but an
anticanonical transformation of $\Gamma$ with a generator functional
$U(a)$ defined by functional $\mathcal{F}(a)$ in (\ref{Eq.5.2}) as
\beq
U(a)=\sum_{k=1}^{\infty}\mathcal{F}_{k|k}(a).
\eeq
In this proof, we have found a possibility to express the
relation between $U(a)$ and $\mathcal{F}(a)$ in the form
\beq
\label{calF1}
\mathcal{F}(a)=-\exp\{{\rm ad}({U}(a))\}\partial_{a}\exp\{-{\rm ad}({U}(a))\}.
\eeq
In turn, the relation (\ref{calF1}) can be considered as a new
representation of the functional (\ref{calF}). Let us notice that
the functional $U(a)$ in (\ref{canGf}) depends on the functional
$F(a)$ only and does not depend on the choice of an initial data
for $\Gamma(a)$. The above formulae (\ref{canGf}) - (\ref{calF1})
just represent the important relationship between the ordinary
exponential and the path-ordered one.
Let us state again that the dependence of the effective action on a
finite anticanonical transformation
with a generating functional $Y(\varphi,\Phi^*;a)$ is really described in terms
of anticanonical transformation with a generator functional
$U(\varphi,\varphi^*;a)$. As an anticanonical transformation is a change
of variables in $\Gamma$,
in particular, it means that, on-shell, the effective action does not depend
on gauges introducing with the help of anticanonical transformations.
\\
\section{Discussions}
In the present article, we have explored a conception of a gauge
fixing procedure in the field-antifield formalism \cite{BV,BV1},
based on the use of anticanonical transformations of general
type. The approach includes an action (master-transformed
action) constructed with the help of anticanonical master transformation
and being non-degenerate. The master-transformed action is a sum of two terms:
one is an action subjected
to an anticanonical transformation and the other is a term
connecting with a logarithm of a superdeterminant of this
anticanonical transformation. This action satisfies the quantum
master equation \cite{BLT4,BBD} (see also Appendix D). The generating
functionals of the Green functions constructed via the
master-transformed action obey the important property
of the gauge independence of physical quantities on-shell, and they
satisfy the Ward identity. We have found that any (finite)
anticanonical master transformation of an action leads to the corresponding
anticanonical transformation of effective action (generating
functional of vertex functions) provided the generating
functional of Green functions is constructed with the help of an
anticanonical master action. We have proved the existence of
a generator functional of an anticanonical transformation of the
effective action. This result is essential when proving the
independence of the effective action of anticanonical
transformations on-shell and, on the other hand, it may supplement
in a non-trivial manner the representation of anticanonical
transformations in the form of a path-ordered exponential
\cite{BBD}.
\\
\section*{Acknowledgments}
\noindent
I. A. Batalin would like to thank Klaus Bering of Masaryk
University for interesting discussions. The work of I. A. Batalin is
supported in part by the RFBR grants 14-01-00489 and 14-02-01171.
The work of P. M. Lavrov is supported by the Ministry of Education and Science of
Russian Federation, project No Z.867.2014/K. The work of
I. V. Tyutin is partially supported by the RFBR grant 14-02-01171.
\\
|
\section{Broadcast channel} \label{cqBroadcast}
\subsection{Marton-Gelfand-Pinsker region for private messages}\label{MGPpm}
In this section we will show how to achieve the Marton-Gelfand-Pinsker region, initially without the use of common messages for classical-quantum broadcast channels using polar codes.
Indeed, we will use the technique of alignment as explained in Section \ref{alignment} to achieve the rate region
\begin{align}
\begin{split}\label{marton}
R_1 &\leq I(V,V_1;B_1), \\
R_2 &\leq I(V,V_2;B_2) \\
R_1 + R_2 &\leq I(V,V_1;B_1) + I(V_2;B_2|V) - I(V_1;V_2|V), \\
R_1 + R_2 &\leq I(V,V_2;B_2) + I(V_1;B_1|V) - I(V_1;V_2|V),
\end{split}
\end{align}
for the classical-quantum two-user broadcast channel described by a classical input $X = \varphi(V, V_1, V_2)$ and a quantum output $\rho_x^{B_1B_2}$.
Let $V, V_1, V_2$ be auxiliary binary random variables with $(V, V_1, V_2) \sim p_V p_{V_2|V} p_{V_1|V_2V}$. Now, let $X = \varphi(V, V_1, V_2)$ be a deterministic function.
Without loss of generality we consider a broadcast channel such that $I(V; B_1) \leq I(V;B_2)$.
With $G_n$ the usual polar coding transformation, set
\begin{align}
U_{(0)}^{n} = V^{n}G_n, \\
U_{(1)}^{n} = V_1^{n}G_n, \\
U_{(2)}^{n} = V_2^{n}G_n.
\end{align}
The variable $U_{(1)}^{n}$ carries the message of the first user, while $U_{(0)}^{n}$ and $U_{(2)}^{n}$ carry the message of the second user.
To exploit the technique of superposition, we take $U_0^n$ to be decodable by both receivers, while $V$ carries information only for the second receiver. The variables $V_1, V_2$ correspond to the binning scheme and can only be decoded by one receiver, respectively.
To handle these additional auxiliary variables we use the polarization technique used to achieve the asymmetric capacity of a channel introduced in Section \ref{asy}. Hence we introduce sets to determine the polarization of the probability distribution for the input and the channel independently.
Define for $l\in\{ 1,2\}$, the following sets, with interpretations provided below,
\begin{align*}
\HH_{V} &= \{ i\in [n] : Z(U_{(0),i}\mid U_{(0)}^{n-1} ) \geq \delta_n \}, \\
\LL_{V} &= \{ i\in [n] : Z(U_{(0),i}\mid U_{(0)}^{n-1} ) \leq \delta_n \}, \\
\HH_{V|B_l} &= \{ i\in [n] : Z(U_{(0),i}\mid U_{(0)}^{n-1} B_{(l)}^{n} ) \geq \delta_n \}, \\
\LL_{V|B_l} &= \{ i\in [n] : Z(U_{(0),i}\mid U_{(0)}^{n-1} B_{(l)}^{n} ) \leq \delta_n \}, \\
\HH_{V_l|V} &= \{ i\in [n] : Z(U_{(l),i}\mid U_{(l)}^{n-1} U_{(0)}^{n} ) \geq \delta_n \}, \\
\LL_{V_l|V} &= \{ i\in [n] : Z(U_{(l),i}\mid U_{(l)}^{n-1} U_{(0)}^{n} ) \leq \delta_n \}, \\
\HH_{V_l|V,B_l} &= \{ i\in [n] : Z(U_{(l),i}\mid U_{(l)}^{n-1} U_{(0)}^{n} B_{(l)}^{n} ) \geq \delta_n \}, \\
\LL_{V_l|V,B_l} &= \{ i\in [n] : Z(U_{(l),i}\mid U_{(l)}^{n-1} U_{(0)}^{n} B_{(l)}^{n} ) \leq \delta_n \}, \\
\HH_{V_1|V,V_2} &= \{ i\in [n] : Z(U_{(1),i}\mid U_{(1)}^{n-1} U_{(0)}^{n} U_{(2)}^{n} ) \geq \delta_n \}, \\
\LL_{V_1|V,V_2} &= \{ i\in [n] : Z(U_{(1),i}\mid U_{(1)}^{n-1} U_{(0)}^{n} U_{(2)}^{n} ) \leq \delta_n \},
\end{align*}
which due to the polarization effect satisfy
\begin{align*}
\lim_{n\rightarrow\infty} \frac{1}{n} |\HH_{V}| &= H(V), \\
\lim_{n\rightarrow\infty} \frac{1}{n} |\LL_{V}| &= 1 - H(V), \\
\lim_{n\rightarrow\infty} \frac{1}{n} |\HH_{V|B_l}| &= H(V|B_l), \\
\lim_{n\rightarrow\infty} \frac{1}{n} |\LL_{V|B_l}| &= 1 - H(V|B_l), \\
\lim_{n\rightarrow\infty} \frac{1}{n} |\HH_{V_l|V}| &= H(V_l|V), \\
\lim_{n\rightarrow\infty} \frac{1}{n} |\LL_{V_l|V}| &= 1 - H(V_l|V), \\
\lim_{n\rightarrow\infty} \frac{1}{n} |\HH_{V_l|V,B_l}| &= H(V_l|V,B_l), \\
\lim_{n\rightarrow\infty} \frac{1}{n} |\LL_{V_l|V,B_l}| &= 1 - H(V_l|V,B_l), \\
\lim_{n\rightarrow\infty} \frac{1}{n} |\HH_{V_1|V,V_2}| &= H(V_1|V,V_2), \\
\lim_{n\rightarrow\infty} \frac{1}{n} |\LL_{V_1|V,V_2}| &= 1 - H(V_1|V,V_2).
\end{align*}
Note that the polarization of the classical quantities follows from polar codes for classical source coding \cite{A10}, while the polarization of the quantities with quantum side information is shown in \cite{SRDR13}.
Intuitively $\HH_{V}$ and $\LL_{V}$ describe the polarization of the random variable $V$ and correspond to whether or not the $i$th bit is nearly completely deterministic given the previous bits, respectively.
Similarly $\HH_{V|B_l}$ and $\LL_{V|B_l}$ determine whether the $l$th receiver can decode bits knowing the previous inputs and all outputs.
$\HH_{V_l|V}$, $\LL_{V_l|V}$, $\HH_{V_l|V,B_l}$, $\LL_{V_l|V,B_l}$ have the same interpretation, with the additional side information from first decoding $V$.
$\HH_{V_1|V,V_2}$ and $\LL_{V_1|V,V_2}$ handle the indices decoded by the first user, with assumed knowledge of $V$ and $V_2$, while in our case that user does not have access to the latter.
Recall that $U_{(0)}^{n}$ can be decoded by both users but only contains information for the second one. Define $\II^{(2)}_{sup} = \HH_V \cap \LL_{V|B_2}$ to contain positions decodable for the second user and $\II_v^{(1)} = \HH_V \cap \LL_{V|B_1}$ those decodable for the first user.
Again, $U_{(2)}^{n}$ can only be decoded by the second user and also only contains information for this receiver. Indeed by $\II^{(2)}_{bin} = \HH_{V_2|V} \cap \LL_{V|B_2}$ we denote the set of indices which can be decoded by the second receiver.
$U_{(1)}^{n}$ only needs to be decoded by the first user and also only contains information for that user. The set $\II^{(1)} = \HH_{V_1|V} \cap \LL_{V|B_1}$ denotes the indices which the first receiver can decode reliably.
We also have to take into account that the first user cannot decode $U_{(2)}^{n}$ therefore the indices in the set $\FF^{(1)} = \LL_{V_1|V,V_2} \cap \HH_{V_1|V} \cap \HH_{V_1|V,B_1}$ are critical.
We now use, in total, three different steps of alignment construction as described in Section \ref{alignment} and illustrated in Figure \ref{broadcast}.
First we handle $U_{(0)}^{n}$. By definition these variables should be decoded by both receivers and contain information only for the second one.
We simply use the alignment technique for classical-quantum channels to send the message assigned for the second receiver to both of them. By the assumption that $I(V; B_1) \leq I(V;B_2)$ we conclude that we can reliably send $I(V; B_1)$ of information to both users. We know that whenever $I(V; B_1)$ is not equal to $I(V; B_2)$ there will be unaligned indices remaining. Lets call this set $\BB^{(2)}$.
In the second step we choose a subset $\BB^{(1)}$ of $\II^{(1)}$ such that $|\BB^{(1)}| = |\BB^{(2)}|$. We can then align these two subsets and therefore raise the number of indices from $U_{(0)}^{n}$, which both receivers can decode, to $I(V; B_2)$.
In the third step we need to cope with the fact that the first user cannot decode the informations in $\FF^{(1)}$.
Again choose a subset $\RRR_{bin}$ of $\II^{(1)}$ such that $|\RRR_{bin}| = |\FF^{(1)}|$. We use $\RRR_{bin}$ to repeat the information for the first user in $\FF^{(1)}$ of the following block.
In order to get the correct order for the successive cancellation decoder, we will encode $U_{(0)}^{n}$ and $U_{(2)}^{n}$ forward, while $U_{(1)}^{n}$ will be decoded backwards.
Moreover, the first receiver decodes $U_{(0)}^{n}$ and $U_{(1)}^{n}$ forwards, while the second receiver decodes $U_{(0)}^{n}$ and $U_{(2)}^{n}$ backwards.
\pgfooclass{ssstamp}{
\method ssstamp() {
}
\method cnot(#1,#2,#3,#4) {
\draw (#1,#2) -- (#1-#3,#2) -- (#1-#3,#2-#4) -- (#1,#2-#4);
\draw (#1-#3,#2) circle (0.1);
\draw (#1-#3,#2+0.1) -- (#1-#3,#2) -- (#1-#3-0.25,#2);
\draw[dotted] (#1-#3-0.25,#2) -- (#1-#3-0.75,#2);
}
\method box(#1,#2,#3) {
\draw (#1,#2) -- (#1+4,#2) -- (#1+4,#2+4) -- (#1,#2+4) -- cycle;
}
}
\pgfoonew \myssstamp=new ssstamp()
\begin{figure*}[t]\label{broadcast}
\centering
\begin{tikzpicture}[scale=0.53]
\tikzstyle{every node}=[font=\tiny]
\node[] at (0+2,10+0.5) {$\II_v^{(1)}$};
\node[] at (0+2,10+3.3) {$\II^{(2)}_{sup}$};
\node[] at (5+2,10+0.5) {$\II_v^{(1)}$};
\node[] at (5+2,10+3.3) {$\II^{(2)}_{sup}$};
\node[] at (10+2,10+0.5) {$\II_v^{(1)}$};
\node[] at (10+2,10+3.3) {$\II^{(2)}_{sup}$};
\draw[line width=2pt, dotted] (0,10) -- (0+4,10+4) -- (0+4,10) -- cycle;
\draw[line width=2pt, dotted] (5,10) -- (5+4,10+4) -- (5+4,10) -- cycle;
\draw[line width=2pt, dotted] (10,10) -- (10+4,10+4) -- (10+4,10) -- cycle;
\draw[line width=2pt, densely dotted] (0,14) -- (0+4,10+4) -- (0+4,10) -- cycle;
\draw[line width=2pt, densely dotted] (5,14) -- (5+4,10+4) -- (5+4,10) -- cycle;
\draw[line width=2pt, densely dotted] (10,14) -- (10+4,10+4) -- (10+4,10) -- cycle;
\node at (0+2,0+0.5) {$\II^{(2)}_{bin}$};
\node at (5+2,0+0.5) {$\II^{(2)}_{bin}$};
\node at (10+2,0+0.5) {$\II^{(2)}_{bin}$};
\draw[line width=2pt, densely dotted] (0,0) -- (0+4,0+0) -- (0+4,2) -- (0,2) -- cycle;
\draw[line width=2pt, densely dotted] (5,0) -- (5+4,0+0) -- (5+4,2) -- (5,2) -- cycle;
\draw[line width=2pt, densely dotted] (10,0) -- (10+4,0+0) -- (10+4,2) -- (10,2) -- cycle;
\draw[line width=2pt, dotted] (0,5) -- (0+4,5+0) -- (0+4,5+2) -- (0,5+2) -- cycle;
\draw[line width=2pt, dotted] (5,5) -- (5+4,5+0) -- (5+4,5+2) -- (5,5+2) -- cycle;
\draw[line width=2pt, dotted] (10,5) -- (10+4,5+0) -- (10+4,5+2) -- (10,5+2) -- cycle;
\myssstamp.box(0,0,2);
\myssstamp.box(5,0,2);
\myssstamp.box(10,0,2);
\myssstamp.box(0,5,1);
\myssstamp.box(5,5,1);
\myssstamp.box(10,5,1);
\myssstamp.box(0,10,0);
\myssstamp.box(5,10,0);
\myssstamp.box(10,10,0);
\draw[->,solid] (2.5,10.5) -- (6,13.5);
\draw[->,solid] (7.5,10.5) -- (11,13.5);
\draw[] (0.5,7.5) -- (1.5,7.5) -- (1.5,8.5) -- (0.5,8.5) -- cycle;
\draw[] (5.5,7.5) -- (6.5,7.5) -- (6.5,8.5) -- (5.5,8.5) -- cycle;
\draw[] (10.5,7.5) -- (11.5,7.5) -- (11.5,8.5) -- (10.5,8.5) -- cycle;
\draw[] (2.5,5.5) -- (3.5,5.5) -- (3.5,6.5) -- (2.5,6.5) -- cycle;
\draw[] (7.5,5.5) -- (8.5,5.5) -- (8.5,6.5) -- (7.5,6.5) -- cycle;
\draw[] (12.5,5.5) -- (13.5,5.5) -- (13.5,6.5) -- (12.5,6.5) -- cycle;
\node at (1,8) {$\FF^{(1)}$};
\node at (6,8) {$\FF^{(1)}$};
\node at (11,8) {$\FF^{(1)}$};
\node at (3,6) {$\RRR_{bin}$};
\node at (8,6) {$\RRR_{bin}$};
\node at (13,6) {$\RRR_{bin}$};
\draw[->,solid] (11.6,12.6) -- (7.15,6.25);
\draw[->,solid] (6.6,12.6) -- (2.15,6.25);
\draw[->,solid] (10.75,7.75) -- (8.25,6.25);
\draw[->,solid] (5.75,7.75) -- (3.25,6.25);
\draw[] (1.3,5.5) -- (2.3,5.5) -- (2.3,6.5) -- (1.3,6.5) -- cycle;
\draw[] (6.3,5.5) -- (7.3,5.5) -- (7.3,6.5) -- (6.3,6.5) -- cycle;
\draw[] (11.3,5.5) -- (12.3,5.5) -- (12.3,6.5) -- (11.3,6.5) -- cycle;
\node at (1.8,6) {$\BB^{(1)}$};
\node at (6.8,6) {$\BB^{(1)}$};
\node at (11.8,6) {$\BB^{(1)}$};
\draw[densely dotted] (1.0,13) -- (3,13);
\draw[densely dotted] (6.0,13) -- (8,13);
\draw[densely dotted] (11.0,13) -- (13,13);
\node at (2.1,12.7) {$\BB^{(2)}$};
\node at (7.1,12.7) {$\BB^{(2)}$};
\node at (12.1,12.7) {$\BB^{(2)}$};
\node[] at (0.7,6) {$\II^{(1)}$};
\node[] at (5.7,6) {$\II^{(1)}$};
\node[] at (10.7,6) {$\II^{(1)}$};
\node at (-1,12) {$U_{(0)}$};
\node at (-1,7) {$U_{(1)}$};
\node at (-1,2) {$U_{(2)}$};
\end{tikzpicture}
\caption[Coding for the broadcast channel.]{Coding for the broadcast channel. Indices in dotted subsets are considered to be good for a specific receiver denoted by the associated set. Arrows indicate the alignment process. For a colored version of this figure see \cite{H14}.}
\end{figure*}
Now if we let the number of blocks approach infinity, we can calculate the rate for the first receiver as follows
\begin{align*}
R_1 &= \frac{1}{n} (|\II^{(1)}| - |\BB^{(1)}| - |\RRR_{bin}|) \\
&= I(V_1; B_1 \mid V) - I(V_1; V_2\mid V) - (I(V;B_2) - I(V;B_2)) \\
&= I(V,V_1; B_1) - I(V_1; V_2\mid V) - I(V;B_2).
\end{align*}
We can also calculate the rate for the second receiver
\begin{align*}
R_2 &= \frac{1}{n}(|\II^{(2)}_{sup}| + |\II^{(2)}_{bin}|) \\
&= I(V; B_2) + I(V_2; B_2\mid V) \\
&= I(V,V_2;B_2).
\end{align*}
Finally note that if we swap the role of the two receivers, the set $\BB^{(2)}$ will be empty due to the assumption that $I(V; B_1) \leq I(V;B_2)$, therefore we can achieve the rates
\begin{align}
R_1 &= I(V,V_1; B_1) \\
R_2 &= I(V_2; B_2\mid V) - I(V_1; V_2\mid V).
\end{align}
For the classical case it is known \cite{MHSU14} that these two rate pairs coincide with the Marton-Gelfand-Pinsker rate region. The proof can be directly translated to the setting of classical quantum communication and therefore our scheme achieves the rate region stated in Equation \ref{marton}.
\subsection{Marton-Gelfand-Pinsker region with common messages}
We can simply extend our coding scheme in Section \ref{MGPpm} to include the transmission of a common message for both receivers, by noting that the information sent via $U_{(0)}^{n}$ can be reliably decoded by both users. Therefore we can use these indices to send an amount of information equal to $\min\{ I(V;B_1), I(V;B_2)\}$ to both users.
This leads to the rate region
\begin{align}
\begin{split}\label{macm}
R_0 &\leq \min\{ I(V;B_1), I(V;B_2)\}, \\
R_0 + R_1 &\leq I(V,V_1;B_1), \\
R_0 + R_2 &\leq I(V,V_2;B_2) \\
R_0 + R_1 + R_2 &\leq I(V,V_1;B_1) + I(V_2;B_2|V) - I(V_1;V_2|V), \\
R_0 + R_1 + R_2 &\leq I(V,V_2;B_2) + I(V_1;B_1|V) - I(V_1;V_2|V).
\end{split}
\end{align}
\section{Introduction}
One of the fundamental tasks in quantum information theory is to determine the maximum possible rate at which information can be sent reliably from one party to another over a noisy communication channel. Indeed to show the \emph{achievability} of a certain rate, one needs to prove the existence of a code, that is, an encoding and decoding scheme, that achieves this rate with vanishing error in the limit of many channel uses.
Network information theory centers on the study and analysis of communication rates in the multi-user setting, generalizing the single sender and receiver case.
The broadcast channel is one the most fundamental channels in this field and, in the two-user case, it models the simultaneous communication between a single sender and two receivers. In the classical setting, there exist several schemes to prove that certain rate regions are achievable for the broadcast channel. Two of these schemes, which are of particular interest to us, are known as superposition coding \cite{B73, C72} and binning \cite{M79}, also called multicoding. For the binning scheme, independent messages are sent simultaneously to both receivers and can be decoded by the respective receiver, according to an argument based on joint typicality. For the superposition scheme we exploit the fact that certain inputs are decodable by both receivers. Marton \cite{M79} combined both techniques in order to send private messages over the two-user broadcast channel, achieving the region now known as Marton's region. Later this result was extended by Gelfand and Pinsker \cite{GP80} where a common message can be sent to both receivers, resulting in the so-called Marton-Gelfand-Pinsker region with common messages. Interestingly, it can be shown, in the classical setting, that even when we set the common rate in the Marton-Gelfand-Pinsker region to zero the resulting rate region is, in some cases, larger than Marton's region \cite{GK11}.
Broadcast channels have also been generalized to the setting of classical-quantum communication \cite{YHD11, SW12, RSW14}.
To date the best known rate region for the classical-quantum channel has been established in \cite{SW12, RSW14} and it is a generalization of Marton's region in the classical case.
Arikan \cite{A09} recently introduced the now celebrated polar coding scheme for classical channels. Indeed Arikan showed that these codes can achieve the symmetric capacity of any classical single-sender single-receiver channel in the limit of many channel uses and, remarkably, that this can be done with a complexity $O(N\log N)$ for encoding and decoding, where $N$ is the number of channels used for communication.
Moreover, polar codes make use of the effect of channel polarization, where a recursive construction is used to divide the instances of a channel into a fraction that can be used for reliable communication and a fraction that is nearly useless. The crucial feature is that the fraction of \emph{good} channels is approximately equal to the symmetric capacity of the channel.
Polar codes have attracted great deal of attention and were generalized to many additional communication settings, such as the task of source coding \cite{A10, A12} and universal coding for compound channels \cite{HU13, MELK132}.
Polar codes have also been generalized to the setting of sending classical information over quantum channels \cite{RDR12,WG13,WG12}, in addition to sending quantum information \cite{WR12, WG13QDeg, SRDR13}. For the task of sending classical information, in addition to asymmetric channels \cite{H14}, the quantum setting has also been generalized to certain multi-user channels, namely the multiple access and interference channels and compound multiple access channel ~\cite{HMW14, H14}.
Recently polar codes have also been applied to the classical broadcast channel \cite{GAG13, MHSU14}.
In \cite{MHSU14} the authors show how polar codes can be used for the broadcast channel to achieve the Marton-Gelfand-Pinsker region with and without common messages.
In this work we will show that the approach of \cite{MHSU14} can be used to achieve the Marton-Gelfand-Pinsker region with and without common messages for classical-quantum broadcast channels, giving rise to the largest known rate region for the classical-quantum broadcast channel.
The remainder of this work is organized as follows. In Section \ref{notation} we will state the necessary preliminaries and in Section \ref{cqBroadcast} we show how to achieve the Marton-Gelfand-Pinsker region for the broadcast channel using polar codes, before we conclude in Section \ref{Conclusion}.
\section{Conclusion}\label{Conclusion}
We have shown that, using polar coding, it is possible to achieve a new rate region for classical-quantum broadcast channel, which coincides with the classical Marton-Gelfand-Pinsker region and is larger than the previously known rate regions for this channel. Hence this work gives an example where polar coding can be used to prove the achievability of new rate regions.
\section*{Acknowledgments}
We would like to thank Mark M. Wilde for enjoyable and fruitful discussions.
This work was supported by the EU grants SIQS and QFTCMPS and by the cluster of excellence EXC 201 Quantum Engineering and Space-Time Research.
\section{Notation and definitions}\label{notation}
We begin by introducing certain notation which will be used throughout the article, before defining necessary entropic quantities and measures.
In the remaining work $u_1^N\equiv u^N$ will denote a row vector $(u_1, \dots, u_N)$ and correspondingly $u_i^j$ will denote, for $1\leq i, j\leq N$, a subvector $(u_i, \dots, u_j)$.
Note that if $j<i$ then $u_i^j$ is empty. Similarly, for a vector $u_1^N$ and
a set $A \subset \{ 1,\dots, N\}$, we write $u_A$ to denote the subvector $%
(u_i : i\in A)$.
A discrete classical-quantum channel $W$ takes realizations $x \in \mathcal{X%
}$ of a random variable $X$ to a quantum state, denoted $\rho_x^B$, on a
finite-dimensional Hilbert space $\mathcal{H}^B$,
\begin{equation}
W : x \rightarrow \rho_{x}^{B},
\end{equation}
where each quantum state $\rho_x$ is described by a positive semi-definite
operator with unit trace. We will take the input alphabet $\mathcal{X} =
\{0,1\}$ unless otherwise stated, and the tensor product $W^{\otimes N}$ of $%
N$ channels is denoted by $W^N$.
To characterize the behavior of symmetric classical-quantum channels, we will make use of the
symmetric Holevo capacity, defined as follows:
\begin{equation}
I(W) \equiv I(X;B)_\rho,
\end{equation}
where the quantum mutual information with respect to a classical-quantum
state $\rho^{XB}$ is given by
\begin{equation}
I(X;B) \equiv H(X)_\rho + H(B)_\rho - H(XB)_\rho,
\end{equation}
with $\rho^{XB} = \frac{1}{2} | 0\rangle\!\langle 0 | \otimes \rho_0^B +
\frac{1}{2} | 1\rangle\!\langle 1 | \otimes \rho_1^B.$
In the above, the von Neumann entropy $H(\rho)$ is defined as
$H(\rho) \equiv -\tr\{\rho \log_2 \rho\}.$
We will also make use of the conditional entropy defined as
$H(X| B)_\rho = H(X)_\rho - H(XB)_\rho$
and the quantum conditional mutual information defined
for a tripartite state $\rho^{XYB}$ as
$I(X;B| Y)_\rho \equiv H(XY)_\rho + H(YB)_\rho - H(Y)_\rho - H(XYB)_\rho.$
We characterize the reliability of a channel $W$ as the fidelity between the
output states
\begin{equation}
F(W) \equiv F(\rho_0, \rho_1),
\end{equation}
with $F(\rho_0, \rho_1) \equiv \normTr{\sqrt{\rho_0}\sqrt{\rho_1}}^2$
and $\normTr{A} \equiv \tr{\sqrt{A^\dagger A}}.$
Note that in the case of two commuting density matrices, $\rho = \sum_i p_i \ketbra{}{i}{i}$ and $\sigma = \sum_i q_i \ketbra{}{i}{i}$ the fidelity can be written as $F(\rho, \sigma) = \left(\sum_i \sqrt{p_i q_i}\right)^2.$
Note that, the Holevo capacity and the fidelity can be seen as quantum generalizations
of the mutual information and the Bhattacharya parameter from the classical
setting, respectively (see, e.g., \cite{A09}).
We will also use the quantity
\begin{equation}\label{Z}
Z(X| B)_{\rho} \equiv 2\sqrt{p_0p_1}\, F(\rho_0, \rho_1),
\end{equation}
introduced in \cite{SRDR13}, for a classical-quantum state $\rho$ which can, again, be seen as quantum generalization of the Bhattacharya parameter, for a classical variable, now with quantum side information.
We will now define the two-user classical-quantum broadcast channel.
The broadcast channel can
be modeled mathematically as the triple $\left( \mathcal{X},W,\mathcal{H}^{B_{1}}\otimes \mathcal{H}^{B_{2}}\right),$ with
\begin{equation}
W : x \rightarrow \rho_{x}^{B_1B_2}.
\end{equation}
The information processing task for the two-user classical-quantum broadcast channel is described as follows. The sender
would like to communicate messages to both
receivers. These messages are independent, but can also contain some common part for both receivers. The model is such that the first receiver only has access to the output system $B_1$ and therefore receives $\rho^{B_1}_x = \tr_{B_2} \rho^{B_1B_2}_x$, similarly, the second receiver has $\rho^{B_2}_x = \tr_{B_1} \rho^{B_1B_2}_x$. The sender chooses a message $m_k$ for each receiver from a message set $\mathcal{M}_k = \{1, \cdots, 2^{nR_k} \}$, and encodes her messages with the resulting the codeword $x^n(m_1,m_2) \in \mathcal{X}^n$.
The receivers' corresponding decoding POVMs are denoted by $\{\Lambda _{m_{1}}\}$ and $\{\Gamma _{m_{2}}\}$.
The code is said to be a $(n,R_{1},R_{2},\epsilon )$-code, if the average
probability of error is bounded as follows
\begin{equation}
\bar{p}_{e}=\frac{1}{|\mathcal{M}_{1}||\mathcal{M}_{2}|}%
\sum_{m_{1},m_{2}}p_{e}(m_{1},m_{2})\leq \epsilon ,
\end{equation}%
where the probability of error $p_{e}(m_1, m_2)$ for a pair of messages $(m_{1},m_{2})$
is given by
\begin{equation}
p_{e}(m_{1},m_{2})=\mathrm{Tr}\left\{\left( I-\Lambda _{m_{1}}\otimes \Gamma
_{m_{2}}\right) \rho _{x^n(m_1,m_2)}^{B^n_{1}B^n_{2}}\right\},
\end{equation}
with $\rho _{x^n(m_1,m_2)}^{B^n_{1}B^n_{2}}$ the state resulting
when the sender transmits the codeword $x^n(m_1,m_2)$
through $n$ instances of the channel. A rate pair $(R_1, R_2)$ is said to be \emph{achievable} for the two-user
classical-quantum broadcast channel described above if there exists an $(n, R_1,R_2, \epsilon)$-code $\forall \epsilon >0$ and large enough $n$.
\subsection{Polar codes for asymmetric channels}\label{asy}
In this section we will review polar codes for achieving the capacity of asymmetric channels, for more details we refer to \cite{H14}.
Essentially polar codes are described by a linear transformation given by $x^N=u^NG_N$, where $u^N$ is the input sequence and
\begin{equation}
G_N = B_NF^{\otimes n}
\end{equation}
with
\begin{equation}
F \equiv \left[
\begin{matrix}
1 & 0 \\
1 & 1
\end{matrix}
\right],
\end{equation}
and $B_N$ is a permutation matrix known as a ``bit reversal'' operation \cite{A09}.
This is called channel combining and transforms $N$ single copies of a channel $W$ to a channel $W_N$.
In the second step, called channel splitting, $W_N$ is used to define $W^{(i)}_N$ as follows:
\begin{equation}
W^{(i)}_N : u_i \rightarrow \rho_{(i),u_i}^{U_1^{i-1}B^N},
\end{equation}
where
\begin{equation}
\rho^{U^{i-1}_1 B^N}_{(i),u_i} = \sum_{u_1^{i-1}} \frac{1}{2^{i-1}} %
\ketbra{}{u_1^{i-1}}{u_1^{i-1}} \otimes \sum_{u_{i+1}^N} \frac{1}{2^{N-i}}
\rho^{B^N}_{u^N}.
\end{equation}
This is equivalent to a decoder which estimates, by the $i$-th measurement, the bit $u_i$, with the following
assumptions: the entire output is available to the decoder, the previous
bits $u_1^{i-1}$ are correctly decoded and the distribution over the bits $u^N_{i+1}$ is uniform.
The assumptions that all previous bits are correctly decoded is called ``genie-aided'' and can be ensured by a limited amount of classical communication prior to the information transmission.
The decoder described above is thus a ``genie-aided'' successive cancellation decoder.
These two steps give rise to the effect of channel polarization, which ensures that the fraction of channels $%
W_N^{(i)}$ which have the property $I(W_N^{(i)}) \in (1-\delta, 1]$ goes to
the symmetric Holevo information $I(W)$ and the fraction with $I(W_N^{(i)})
\in [0, 1-\delta)$ goes to $1-I(W)$ for any $\delta \in (0,1)$, as $N$ goes to
infinity through powers of two. This is one of the main insights of the work by Arikan \cite{A09} and the generalization in \cite{WG13} to the classical-quantum setting (see \cite{WG13} for a more detailed statement).
To achieve the symmetric capacity we can now simply send information bits over the channels $I(W_N^{(i)}) \in (1-\delta, 1]$ and send prearranged ``frozen'' bits over the remaining channels.
It turns out that, the above approach essentially works for asymmetric channels as well, the crucial point to see this, is that the polar coding transform $G_N$ is its own inverse for binary inputs.
We consider the reverse protocol of classical lossless compression. Hence we can use a uniformly distributed input sequence and transform it to a distribution suitable to achieve the asymmetric capacity of the channel. Due to a polarization effect we can use a fraction of size $H(X)$ for the following channel coding. Note that in the special case of a symmetric channel the uniformity of the required input distribution simply gives a fraction $H(X)=1$.
It is shown in \cite{H14} that this approach achieves the asymmetric capacity
\begin{equation}
C(W) = \max_{p(x)} I(X;B) .
\end{equation}
This is a generalization of a result in \cite{HY13} to the classical-quantum setting.
\subsection{Alignment of polarized sets}\label{alignment}
From the definition of polar codes, it is clear that the set of channels which can be used for information transmission depend on the particular communication channel to be used.
Polarizing, in a scenario with multiple possible channels, such as the case of compound channels, where one must code at rates which are achievable for all channels in a particular known a set of channels, will hence, lead to the situation where some synthesized channels are good for one channel but not for another, and vice versa. This problem can be solved by the technique of \emph{alignment} \cite{HU13}, which is described in detail for the classical-quantum compound channel in \cite{H14}. The main idea is to combine the channels which are good in one case with the channels good in the other by additional CNOT gates. Doing this recursively, we can halve the number of incompatible indices in every step. With the number of channels uses approaching infinity, we can minimize the number of incompatible indices.
|
\section{\label{sec::intro}Introduction}
The symbolic manipulation of complicated formulae has a long tradition in
particle physics. Computer algebra systems (CAS) have been used already
quite early in order to evaluate, e.g., traces over $\gamma$ matrices.
Among the first CAS there are {\tt REDUCE}~\cite{Hearn:1971zza} by
A.~Hearn, {\tt SCHOONSCHIP}~\cite{Schoonschip,Schoonschip2,Veltman:1991xb},
designed by M.~Veltman, {\tt ASHMEDAI}~\cite{ashmedai} by M.~Levine,
and {\tt Macsyma}~\cite{macsyma} developed at MIT.
Afterwards {\tt Mathematica}~\cite{mathematica}, {\tt Maple}~\cite{maple}
and others have been developed which are still in use nowadays. However,
their field of application is limited to small and medium sized problems
since it is not possible to work with very large intermediate expressions.
On the other hand, there are quite a number of problems which produce
intermediate expressions of the order of a few hundred giga bytes up to
tera bytes to be manipulated by the CAS. The only CAS currently available
in order to cope with such tasks is {\tt FORM}{}~\cite{Vermaseren:2000nd,
Kuipers:2012rf}.
{\tt FORM}{} is a program for the
symbolic manipulation of algebraic expressions. It is specialized to handle
very large algebraic expressions of billions of terms in an efficient and
reliable way. That is why it is widely used, in particular in the framework
of perturbative Quantum Field Theory, where often several thousands of
Feynman diagrams have to be computed. However, the abilities of {\tt FORM}{} are
also quite useful in other fields of science where the manipulation of huge
expressions is necessary.
{\tt FORM}{} is constructed in such a way that the size of the expressions is
not restricted by the main memory of the computer but only by the space
available on hard disk. In addition its data representation is very dense
when compared to other general purpose systems. Actually in modern
applications in particle physics it happens quite often that the size of
intermediate expressions for each Feynman diagram may become huge. As a
consequence, even with {\tt FORM}{} such calculations require a CPU time of
several years despite the steady advancement of the hardware and the
continuous improvement of the algorithms.
Furthermore the resources as far as CPU speed, memory and disk
space are concerned are often not sufficient.
One of the most efficient ways to increase the performance is based on
parallelization which makes simultaneously available the
resources of several computers and thereby significantly reduces the wall
clock time. In fact, the project to obtain a parallel version of
{\tt FORM}{} has been started at the end of the nineties. In the recent years
{\tt ParFORM}{}~\cite{Tentyukov:2004hz} and {\tt TFORM}~\cite{Tentyukov:2007mu} have
become reliable tools which shall be described in this contribution.
There is a number of calculations performed within project A1 of the
CRC/TR~9 where {\tt ParFORM}{} and {\tt TFORM}{} were essential for the successful
completion~\cite{Baikov:2002uw,Baikov:2002va,Baikov:2003zg,Baikov:2003gu,Baikov:2004tk,Chetyrkin:2005kn,Baikov:2005rw,Baikov:2006ch,Baikov:2008jh,Baikov:2009bg,Baikov:2010je,Baikov:2012er,Baikov:2012zm,Baikov:2012zn,Baikov:2014qja}.
In all these cases the single-core CPU time was estimated to several years.
Parallelization could reduce the wall clock time to weeks and months at
most.
As a further application we want to mention Ref.~\cite{Blumlein:2009cf}
where {\tt FORM}{} was used to solved exceptionally large systems of equations
to create mathematical tables for general use in mathematics and physics.
The calculation of three-loop helicity-dependent splitting functions in
QCD~\cite{Vogt:2014pha,Moch:2014sna} also could only be completed thanks to {\tt FORM}{}
because expressions of one tera byte or more were no exception and at one
point more than 6 tera bytes of diskspace was needed for a single diagram.
Within the CRC/TR~9 two concepts for parallel versions of {\tt FORM}{} have
been successfully developed and implemented: {\tt ParFORM}{}, essentially
based on MPI (message passing interface), and {\tt TFORM}{} which uses
threads for the parallelization. Both programs run stable, show a good
speedup and are complete in the sense that all programs written for the
serial version of {\tt FORM}{} can now be used with {\tt ParFORM}{} and {\tt TFORM}{}.
In Sections~\ref{sec::parform} and~\ref{sec::tform} details to the
parallel versions are provided.
In this project of the CRC/TR~9 also programs concerned with the reduction
of families of Feynman integrals to a small set of basis elements (master
integrals) and their numerical evaluation have been developed. These two
topics are covered in two program packages, {\tt FIRE} and {\tt FIESTA},
which are discussed in Section~\ref{sec::other}
We continue this review in Section~\ref{sec::hist} with some historical
remarks concerning the first steps towards parallelization of {\tt FORM}{}
and describe in Section~\ref{sec::form} the basic features of {\tt FORM}{}.
\section{\label{sec::hist}Historical remarks}
The first initiatives of parallizing {\tt FORM}{} go back to early 1991, when
version 1 of {\tt FORM}{} was made to run on a computer at the Fermi National
Accelerator Laboratory (FNAL) which was
designed for lattice calculations and had 257 processors. Due to
limitations in accessibility this project was discontinued, but the further
development of {\tt FORM}{} took this experience into account.
The first systematic study of a parallel version of {\tt FORM}{} has been
performed within the DFG-funded Research Unit ``Quantenfeldtheorie,
Computeralgebra und Monte-Carlo Simulation'' which ran from 1996 to 2002
and thus can be considered as a precursor to the CRC/TR~9. In
Ref.~\cite{Fliegner:1999jq} a first parallel prototype of {\tt FORM}{} has
been presented and results for several studies like the runtime for the
parallel sorting on different architectures are shown.
One year later, in July 2000, the first ``working parallel {\tt FORM}{}
prototype, {\tt ParFORM}{}'', has been introduced in
Ref.~\cite{Fliegner:2000uy}. It was based on the syntax of a preliminary
version of {\tt FORM}{}~3 which at that time was not published yet.
In~\cite{Fliegner:2000uy} the parallelization on clusters has been
discussed based on the following hardware:
\begin{itemize}
\item
Digital workstation cluster (TTP Karlsruhe) running DEC UNIX 4.0D
8 nodes with 600~MHz Alpha 21164A (EV56) processors and 512~MB RAM,
\item
PC cluster (TTP Karlsruhe) running Linux 2.2.13
4 nodes with 500~MHz Intel Pentium III processors and 256~MB RAM,
\item
IBM SP2 (Computing Center Karlsruhe) running AIX 4.2.1
160 thin P2SC nodes with 120~MHz processors and 512~MB RAM (256 nodes in total).
\end{itemize}
Next to several feasibility studies also results for the speedup of a {\tt
MINCER}~\cite{Larin:1991fz} job is shown. A reasonable speedup of 2.5 with
four nodes on the PC cluster, a factor of 4.5 with eight nodes on the Alpha
cluster and a factor of 6 with twelve nodes on the IBM SP2 has been
reported. As a first physical application of {\tt ParFORM}{} higher moments
of deep inelastic structure functions at next-to-next-to-leading order of
perturbative QCD have been computed in Ref.~\cite{Retey:2000nq}.
\begin{figure}[t]
\begin{center}
\includegraphics[width=\linewidth]{speedup_qcmsmp_2002}
\caption{\label{fig::speedup_qcmsmp_2002}Speedup for the program {\tt BAICER} on
Compaq-AlphaServer with 8 Alpha (EV67) processors with 700 MHz.}
\end{center}
\end{figure}
At a later stage of the Research Unit {\tt ParFORM}{} was further developed
and one could run parallel {\tt FORM}{} jobs on symmetric multiprocessing
(SMP) computers (not only on clusters). In
Fig.~\ref{fig::speedup_qcmsmp_2002} the speedup is shown for the test
program {\tt BAICER}, a {\tt FORM}{} program developed to compute massless
four-loop two-point integrals within the project A1 of the CRC/TR~9,
running on a
\begin{itemize}
\item
Compaq-AlphaServer GS60e, 8 Alpha (EV67) processors (700 MHz).
\end{itemize}
A speedup of about 4.5 could be achieved using eight processors.
Two years after the start of the CRC/TR~9 a first version of {\tt ParFORM}{}
operating on Cluster- and SMP-architectures was discussed in
Ref.~\cite{Tentyukov:2004hz}. It could run arbitrary {\tt FORM}{} programs in
parallel and was based on {\tt FORM}{}~3 version 3.1~\cite{Vermaseren:2000nd}. At
that time there were already a number of applications which would not have
been possible
without {\tt ParFORM}{}~\cite{Retey:2000nq,Baikov:2002uw,Baikov:2002va,Baikov:2003gu}.
\begin{figure}[t]
\begin{center}
\includegraphics[width=\linewidth]{speedup_sgi_2004}
\caption{\label{fig::speedup_sgi_2004}Computing time and speedup for the
test program BAICER on the SGI Altix 3700 server with 32x Itanium-2
processors (1.3 GHz).}
\end{center}
\end{figure}
For the calculations and for the development of {\tt ParFORM}{} a 32-core
computer was available
\begin{itemize}
\item
SGI Altix 3700 Server 32x 1.3 GHz/3~MB-SC Itanium-2 CPUs
64 GB DDR/116 MHz mem, 2.4 TB SCSI hard disks.
\end{itemize}
The results for the test program {\tt BAICER} are shown in
Fig.~\ref{fig::speedup_sgi_2004}. The speedup is almost linear up to twelve
processors. Afterwards it flattens but is still considerable. An achieved
speedup of 12 means that a {\tt FORM}{} job that would need one year of
computing time can be run as {\tt ParFORM}{} job in about one month. This
leads to a qualitatively new level, because it would practically be
impossible to run jobs for years whereas months are feasible nowadays.
Fig.~\ref{fig::speedup_sgi_2004} shows that with 16 processors a speedup of
10 could be reached. This means that one can run on a 32-processor computer
two jobs simultaneously, having the speedup of 10 for each of them.
In the paper~\cite{Tentyukov:2006ys} the functionality of {\tt FORM}{} and
{\tt ParFORM}{} was extended and facilities were introduced to communicate
with external resources. This mechanism enables the user to include into
the {\tt FORM}{} programs other pieces of software which are used as black
box in order to take over certain tasks. As a typical example we want to
mention is {\tt fermat}~\cite{fermat}, which can compute the greatest
common divisor of multi-variable polynomials efficiently.
In February 2007 {\tt TFORM}{}~\cite{Tentyukov:2007mu} based on {\tt POSIX}
threads has been released, a further major step in the development of
parallel {\tt FORM}{} versions.
For later developments and further comparisons between {\tt ParFORM}{} and {\tt TFORM}{}
we refer to the proceedings
contributions~\cite{Vermaseren:2006ag,Tentyukov:2008zz,Tentyukov:2010qf,Tentyukov:2006pr}
and to Sections~\ref{sec::parform} and~\ref{sec::tform}.
The more recent developments concern the release of
{\tt FORM}{}~4.0~\cite{Kuipers:2012rf} and the inclusion of tools to generate
optimized
code~\cite{Kuipers:2013pba} which is used as input in {\tt FORTRAN} or {\tt
C} programs for numerical integrations.
\section{\label{sec::form}Sequential version of {\tt FORM}{}}
This article is not intended as an introduction to {\tt FORM}{} or even a
reference manual. Nevertheless we want to describe the basic features which
are important in the context of parallelization.
A {\tt FORM}{} program is in general divided into so-called modules which are
terminated by a ``dot''-instruction. During the execution of the program,
which is only possible in batch-mode, each module is processed
separately one after the other which essentially occurs in three steps
\begin{itemize}
\item{Compilation:} The input is translated into an internal
representation.
\item{Generating:} For each term of the input expressions the statements of
the module are executed. This in general generates a lot of terms.
\item{Sorting:} All the output terms that have been generated are sorted
and equivalent terms are summed up.
\end{itemize}
This is illustrated in Fig.~\ref{fig::form_stream}.
\begin{figure}[t]
\begin{center}
\includegraphics[width=.45\textwidth]{form_stream}
\caption{\label{fig::form_stream}
Graphical representation of the processing of an input expression in {\tt FORM}{}.
}
\end{center}
\end{figure}
The fundamental objects which are manipulated by {\tt FORM}{} commands are
expressions which are viewed as sums of individual terms (see also
Fig.~\ref{fig::form_stream}). Next to a sophisticated pattern matcher, it
is the strength of {\tt FORM}{} that only local operations on single terms are
allowed, like replacing parts of a term by some other expressions.
Non-local operations like replacing a sum of two terms are not allowed. For
example, the command \verb|identify| (short: \verb|id|) identifies the
left-hand side with the right-hand side and can be used as
\begin{verbatim}
id a = b + c;
\end{verbatim}
On the other hand, the usage
\begin{verbatim}
id a + b = c;
\end{verbatim}
would lead to an error message.
Non-local operations are allowed only implicitly, e.g., in the sorting
procedure at the end of the modules, where equivalent terms are combined.
At first sight this seems to be a strong limitation for the formulation of
general and efficient algorithms. It is usually possible to get around this
limitation by designing algorithms in clever and non-standard ways.
Due to the locality of the operations it is possible to handle expressions
as ``streams'' of terms that can be read sequentially from the memory or a
file and processed independently. This enables {\tt FORM}{} to deal with
expressions that are larger than the available main memory.
\begin{figure}[t]
\begin{center}
\includegraphics[width=.45\textwidth]{form_example}
\caption{\label{fig::form_example}
Example for the generation and sorting of data in {\tt FORM}{}.
}
\end{center}
\end{figure}
An example illustrating the principle operating mode of a {\tt FORM}{} program
is shown in Fig.~\ref{fig::form_example}. It corresponds to the simple
program
\begin{verbatim}
l expr = a*x + x^2;
id x = a + b;
.sort
if (count(b,1)==1);
multiply 4*a/b;
endif;
print;
.end
\end{verbatim}
\section{\label{sec::parform}{\tt ParFORM}{}}
\subsection{The concept of {\tt ParFORM}{}}
As mentioned above, the locality principle enables {\tt FORM}{} on the one hand
to deal with expressions that are larger than the available main memory, on
the other hand it also allows for parallelization. The concept implemented
in {\tt ParFORM}{} is straightforward and indicated in
Fig.~\ref{fig::parexample}: in a first step the master process splits the
expression into pieces, so-called chunks. Each chunk is sent to one of the
workers where an independent {\tt FORM}{} process runs, i.e. the module to be
executed is compiled, the terms are generated, sorted and sent back to the
master. Once all worker processes have finished their jobs the master
performs the final sorting.
\begin{figure}[t]
\begin{center}
\includegraphics[width=.45\textwidth]{parexample}
\caption{
\label{fig::parexample}
General conception of {\tt ParFORM}{}.
}
\end{center}
\end{figure}
The communication between master and workers is based on the message
passing interface (MPI) standard~\cite{MPI} which provides a library for
the data transfer between processes. Message passing permits to parallelize
{\tt FORM}{} on computer architectures both with shared memory, i.e.
SMP computers and on computer clusters. The way the
master communicates with the workers is sketched in
Fig.~\ref{fig::parform_mpi}.
\begin{figure}[t]
\begin{center}
\includegraphics[width=.45\textwidth]{parform_mpi}
\caption{
\label{fig::parform_mpi}
Visualization of the mode of operation of {\tt ParFORM}{} based on MPI.
}
\end{center}
\end{figure}
It is worth mentioning that the parallelization does not require any
additional efforts from the user. It is possible to run the programs
written for the sequential version using {\tt ParFORM}{} and adding a
specification concerning the number of processors. It is clear that
different codes show a different performance and efficiency in the parallel
version. In particular, modules in which the outcome depends on the order
in which the terms are processed cannot be parallelized and are executed in
sequential mode. This concerns mostly the use of the dollar variables which
were introduced in version 3. In the case that {\tt FORM}{} would switch to
sequential mode, while actually this is not needed, the user can add an
extra statement to overrule such a decision and tell {\tt FORM}{} how to deal
with the `dubious' case.
\subsection{\label{sub::numa}{\tt ParFORM}{} on a NUMA architecture}
The SGI Altix computer is realized with a so-called NUMA architecture where
NUMA stands for non-uniform memory access. This means that the individual
processors have a faster access to some parts of the main memory than to
others. A specialized version of {\tt ParFORM}{} has been developed which
exploits the feature and, at the same time, does not use MPI and the
overhead connected to it. The corresponding scheme of operation is
illustrated in Fig.~\ref{fig::mpi_numa}.
\begin{figure}[t]
\begin{center}
\includegraphics[width=.45\textwidth]{mpi_numa}
\caption{
\label{fig::mpi_numa}
Mode of operation implemented into {\tt ParFORM}{} for a NUMA architecture.
}
\end{center}
\end{figure}
Using the specialized version of {\tt ParFORM}{} in connection with the
32-core SGI Altix a considerable improvement of the speedup could be
obtained, as can be seen in Fig.~\ref{fig::mpi_numa_speedup}. In fact, for
16 processors the speedup improved from 8 to 10, for 32 processors from 10
to 13 (see also the discusion in the next subsection).
\begin{figure}[t]
\begin{center}
\includegraphics[width=.45\textwidth]{mpi_numa_speedup}
\caption{
\label{fig::mpi_numa_speedup}
Runtime and
speedup for the test program {\tt BAICER} running on a
SGI Altix 3700 server with 32 Itanium-2 processors (1.3 GHz).
The lower curve corresponds to the MPI version and the upper
one to the shared memory version of {\tt ParFORM}{}.
}
\end{center}
\end{figure}
\subsection{{\tt ParFORM}{} on clusters and multi-core nodes}
At present, there are a number of calculations of physical quantities which
would not have been possible without the gain in performance and speedup
provided by {\tt ParFORM}{} (see, e.g.,
Refs.~\cite{Retey:2000nq,Baikov:2002va}). Most of the applications are
connected to the evaluation of four-loop Feynman integrals which occur in the
context of perturbative quantum field theory. In particular, there are
algorithms which transform the mathematical complexity of the original problem
to the need of simple manipulations of rather large polynomial expressions
which have billions or even more terms. Manipulations of this type constitute
the basis of the speedup curves which are discussed in the following.
The results for the test program running on a SGI Altix 3700 server with 32
Itanium-2 processors are shown in Fig.~\ref{fig::mpi_numa_speedup} where
both the runtime and the speedup (as compared to the sequential version) is
shown as a function of the number of processors, $p$, involved in the
calculation. The almost horizontal line between $p=1$ and $p=2$ is due to
the fact that for $p=2$ one of the processors takes over the role of the
master and the other one of the worker. Thus a real reduction of the CPU
time only starts from $p=3$. It is interesting to note that the speedup is
almost linear up to twelve processors. Furthermore, for 16 processors the
program is faster by an order of magnitude. As a consequence instead of
years one only has to wait a few months in order to obtain the results of a
calculation. This provides the possibility to consider qualitatively new
kinds of problems, since in practice it is impossible to run a job for
years whereas a few months are feasible nowadays. Beyond $p=16$ the curve
becomes more flat, however, the speedup is still considerable up to 32
processors.
The latest speedup plot for (the MPI version of) {\tt ParFORM}{} is shown in
Fig.~\ref{fig::speedup_parform} where {\tt BAICER} is running on the cluster
{\tt ttpmoon} which has the following configuration:
\begin{itemize}
\item
Computer cluster (TTP Karlsruhe) running Linux,
8 nodes with
2 Hexa-Core Intel Xeon X5675 (3.07 GHz),
96 GB RAM,
and 3.6 TB local hard disk (Raid 0 with 6 stripes),
interconnected by QDR InfiniBand.
\end{itemize}
The top plot shows the used time in minutes as a function of the involved CPUs
(including the master) and on the bottom the speedup as compared to the serial
version is plotted.\footnote{Note, that there is no data point for two CPUs;
otherwise one would observe a flat behaviour between one and two CPUs and
only then the curve starts to raise.} It is interesting to note that a
speedup of about 10 is reached in case 16 CPUs are used, a value obtained in
Fig.~\ref{fig::mpi_numa_speedup} for the shared memory version which avoids
the use of MPI, cf. Subsection~\ref{sub::numa}. For higher number of CPUs the
curve flattens but nevertheless reaches a speeup above 20 for 96 CPUs.
\begin{figure}[t]
\begin{center}
\includegraphics[width=.4\textwidth]{ttpmoon-N16a}
\\
\includegraphics[width=.4\textwidth]{ttpmoon-N16b}
\caption{
\label{fig::speedup_parform}
Timing and speedup plot for the {\tt ParFORM}{} benchmark job
{\tt BAICER} running of {\tt ttpmoon}.
}
\end{center}
\end{figure}
\subsection{{\tt ParFORM}{} on ``low-level'' clusters}
\begin{figure}[t]
\begin{center}
\includegraphics[width=.4\textwidth]{a2_newcluster}
\caption{
\label{fig::clusters}
The speedup for the test program on different clusters in comparison
to the SMP computer (cf. Fig.~\ref{fig::mpi_numa_speedup}).
}
\end{center}
\end{figure}
{\tt ParFORM}{} has been successfully installed on several clusters. In
Fig.~\ref{fig::clusters} the corresponding speedup curves are shown and
compared to the curve from Fig.~\ref{fig::mpi_numa_speedup} obtained on the
SMP computer. The cluster {\tt XC6000} is a Hewlett Packard Itanium-2 QsNet
interconnected cluster. This is the only tested cluster which demonstrates
a better behaviour than the SMP computer, however, it is also significantly
more expensive. {\tt Fphctl} is a cluster consisting of 32-bit Xeon nodes.
This cluster has been tested both with an Infiniband ({\tt FphctlIB}) and a
simple Fast Ethernet ({\tt FphctlEN}) interconnection. Whereas the latter
is not of interest in practice the former shows a quite reasonable
behaviour following closely the SMP curve for a smaller number of
processors. {\tt Plejade} and {\tt Empire} are both dual Opteron clusters.
However, {\tt Plejade} is interconnected using InfiniBand whereas {\tt
Empire} uses Gigabit Ethernet. Both clusters show a reasonable behaviour
leading to a speedup of about six for ten processors.
We want to mention that the SMP curves shown in Fig.~\ref{fig::clusters}
are based on the shared-memory model mentioned above. On the other hand,
for the clusters one has to rely on the MPI library which for our
applications has a significant overhead.
\section{\label{sec::tform}{\tt TFORM}{}}
In the last decade multi-core processing has become a key technology in the
computing industry as system performance improvement through increasing
clock rates of single-core processors is hindered by physical limits. From
laptops to supercomputers multi-core processing is prevalently used and the
modern operating systems allow one to easily use them as SMP computers.
Although {\tt ParFORM}{} works on such SMP computers, interprocess
communications among the master and the workers via MPI can have a
significant overhead when gigantic expressions are transfered.
This overhead problem can be overcome on SMP architectures with the help of
another model for the communication. In this approach the master explicitly
allocates shared memory buffers which can be accessed both by the master and
the workers. In these memory segments the master prepares the chunks for the
workers, they are doing their job and the master collects the results again
from the shared buffers. Thus, copying huge amounts of data is not necessary
any more. The use of the shared-memory model on SMP machines led to an
increase in the speedup of 20-25\%
(cf. Fig.~\ref{fig::mpi_numa_speedup}). This concept is taken even further in
{\tt TFORM}{}~\cite{Tentyukov:2007mu}, a multithreaded version of {\tt FORM}{}.
In {\tt TFORM}{} the implementation uses the POSIX threads library, which is
available on all modern UNIX systems and therefore portable. The way the
master communicates with the workers is sketched in
Fig.~\ref{fig::tform_thr}. {\tt TFORM}{} starts with one master thread and $N$
worker threads in a so-called thread pool. The workers sleep until the
master assigns tasks, and hence do not spend any CPU time. When the master
has some task to be distributed over the workers, the master wakes up one
of the sleeping workers and assigns the task to it. Terms in expressions,
grouped as chunks for reducing the overheads, are distributed in this way.
After distributing all terms, the master waits for all the workers to
finish the tasks, and then the master merges the results of the workers in
a final sorting operation. The data transfer among the threads is done via
the shared memory buffers and by using memory locks for synchronization
between the master and the workers (see Fig.~\ref{fig::tform_thr}).
\begin{figure}[t]
\begin{center}
\includegraphics[width=.45\textwidth]{tfrom_thr}
\caption{
\label{fig::tform_thr}
Mode of operation for {\tt TFORM}{}.
}
\end{center}
\end{figure}
Due to the model for the communications, some features improving the
performance are relatively easy to implement in {\tt TFORM}{}, whereas their
implementations are difficult in {\tt ParFORM}{}. One of them is a load
balancing system. If there is a single worker that is assigned terms
requiring much CPU time, for the final sorting the master may have to wait
for this worker even after the other workers finish their tasks and become
idle. To avoid such inefficiency, after distributing all terms to be
processed, the master looks for idle workers. If such workers are found,
terms are stolen back from the chuncks of workers that are still busy and
redistributed over idle workers. Experiments with an even more fine-grained
load balancing were unsuccessful, because they resulted in too much
overhead.
Another feature in {\tt TFORM}{} concerns the parallel sorting. In the final
sorting, {\tt TFORM}{} used to adopt the simple model in which the master merges
the outputs from all the workers simultaneously. Therefore it often happens
that the master is busy while the workers are waiting for the master to
accept their next chunks of the results. It becomes a bottleneck,
especially when the number of the workers is large. To alleviate this
bottleneck, an improved model of the final sorting has been implemented in
{\tt TFORM}{}. In this model, each two workers send their results to a special
worker thread, called a sortbot, which merges the results. Then each two
sortbots send their results to another sortbot. This continues until the
last two sortbots send their results to the master, which merges the final
two results and writes the result to disk. This is illustrated in
Fig.~\ref{fig::sortbots}.
Because also this method still involves much waiting, a run with $N$
workers will rarely use more than the CPU time provided by $N$ cores, even
when the computer has many more cores. The total wall clock execution time
improves measurably by this method, although it does go at the cost of
extra memory needed for the buffers of the sortbots.
\begin{figure}[t]
\begin{center}
\includegraphics[width=.45\textwidth]{sortbots}
\caption{
\label{fig::sortbots}
Illustration of the mode of operation of sortbots.
}
\end{center}
\end{figure}
Fig.~\ref{fig::speedup_tform_parform} shows up-to-date timing and speedup
plots for {\tt ParFORM}{} and {\tt TFORM}{} running on {\tt
ttpmoon}.\footnote{The {\tt ParFORM}{} curves are already shown in
Fig.~\ref{fig::speedup_parform}.} Note that the cluster {\tt ttpmoon}
consists of 12-core nodes which explains the end point of the {\tt TFORM}{} curves
where a speedup better than 9 is reached. {\tt ParFORM}{} reaches for 12 CPUs,
which means 1 master and 11 workers, a speedup of~8.
\begin{figure}[t]
\begin{center}
\includegraphics[width=.45\textwidth]{ttpmoon-baicer16-a}
\\
\includegraphics[width=.45\textwidth]{ttpmoon-baicer16-b}
\caption{
\label{fig::speedup_tform_parform}
Comparison of timing and speedup for {\tt ParFORM}{} and {\tt TFORM}{} running on
{\tt ttpmoon}.}
\end{center}
\end{figure}
\section{\label{sec::other}Further developments within CRC/TR~9}
\subsection{Reduction to master integrals with {\tt FIRE}}
Nowadays the vast majority of calculations of higher order quantum corrections
involve a huge number (sometimes exceeding several millions) of different
contributing integrals. The standard way to reduce their number to a
manageable amount is based on the so-called ``Laporta algorithm'' which is
described in Ref.~\cite{Laporta:2001dd}. There are many different
implementations of this algorithm, some of them are publicly available like
{\tt AIR}~\cite{Anastasiou:2004vj} or {\tt
Reduze}~\cite{Studerus:2009ye,vonManteuffel:2012np}, others are private like
{\tt crusher}~\cite{crusher} which has been developed in the context of
project~A1 of the CRC/TR~9. Within project~A2 the program {\tt
FIRE}~\cite{Smirnov:2008iw,Smirnov:2013dia,Smirnov:2014hma,fire} has been
developed.
{\tt FIRE} stands for Feynman Integral REduction and implements a special
version of the Gauss elimination method to solve the system of linear equations,
which is generated by the application of the integration-by-parts
relations~\cite{Chetyrkin:1981qh}, for the master integrals. It uses several
external programs like {\tt Snappy}~\cite{snappy} for data compression,
{\tt KyotoCabinet}~\cite{kyotocabinet} as database to store data on disk, {\tt
Fermat}~\cite{fermat} for algebraic simplifications, and {\tt
LiteRed}~\cite{litered} to retrieve additional rules among integrals.
The operation of {\tt FIRE} is divided into two parts: in a first step the
input for the reduction step is prepared within {\tt Mathematica}. This
includes the generation of all integration-by-parts relations, the generation
of symmetry relations, the identification of the sectors of indices where
integrals vanish, i.e. the so-called boundary conditions, and the preparation
of a list of integrals which shall be reduced. The second step is
significantly more time consuming. In the latest version, {\tt
FIRE5}~\cite{Smirnov:2014hma}, this part is written in {\tt C++}. Here the
systematic reduction to master integrals is performed. The output is a table
for the list of integrals provided in part one.
\begin{figure}[t]
\begin{center}
\includegraphics[width=.3\textwidth]{top2l2h1a}
\caption{
\label{fig::fire_ex}
Integral family with three massive (double lines), three massless lines
and an irreducible numerator (not shown).
It it evaluated in forward-scattering kinematics, i.e., there is
one external momentum, $p_1$, flowing through the upper massless line
and another, $p_2$, through the lower massive lines.}
\end{center}
\end{figure}
To demonstrate the use of {\tt FIRE} let us, for example, consider the
integral family of Fig.~\ref{fig::fire_ex} which has three massive internal
lines (with mass $m$). For the external momenta we have $p_4=p_1$, $p_3=p_2$
with $p_1^2=p_2^2=0$. Integrals of that type contribute to the
next-to-leading order corrections to double-Higgs boson production. In fact,
the imaginary part, which is a function of $x=m^2/s$ (with $s=(p_1+p_2)^2$),
is related to the total cross section via the optical theorem. The input for
the {\tt Mathematica} part of {\tt FIRE} contains the following elements (for
a detailed description of the commands we refer to
Ref.~\cite{Smirnov:2014hma}):
\begin{verbatim}
(* load FIRE: *)
Get["FIRE5.m"];
(* define integral family: *)
Propagators = {m^2 - (v1-v2)^2,
m^2 - (p2-v1)^2, m^2 - (p2-v2)^2,
-v2^2, -v1^2, -(p1+v1)^2, -(p1+v2)^2};
Internal = {v1,v2};
External = {p1,p2};
(* IBP relations: *)
PrepareIBP[];
kinset = {p1^2 -> 0, p2^2 -> 0,
p1*p2 -> s/2};
set1 = Internal;
set2 = Join[Internal,External];
ncount = 0;
startinglist = {};
For[ii=1,ii<=Length[set1],ii++,
For[jj=1,jj<=Length[set2],jj++,
ncount = ncount + 1;
ff[ncount] =
IBP[set1[[ii]], set2[[jj]]
] /. kinset;
startinglist =
Join[startinglist,{ff[ncount]}];
];
];
(* boundary conditions: only contributions
with cuts through at least 2 Higgs
lines are kept: *)
(RESTRICTIONS = { {0,-1,0,0,0,-1,0},
{0,0,-1,0,0,0,-1},{0,0,0,0,0,-1,-1},
{-1,0,0,0,0,0,0},{-1,-1,0,0,0,0,0},
{-1,0,-1,0,0,0,0},{0,-1,-1,0,0,0,0} });
SYMMETRIES = { {1,3,2,5,4,7,6} };
Prepare[];
(* save data to top2l2h1a.start: *)
SaveStart["top2l2h1a"];
\end{verbatim}
The last command writes all generated information into the so-called ``start'' file
which, together with the list of integrals, serves as input for the reduction
step. The steering file, {\tt top2l2h1a.config},
for the latter has the following form
\begin{verbatim}
#threads 4
#variables d,s,m
#start
#problem 1|7|top2l2h1a.start
#integrals top2l2h1a.ind
#output top2l2h1a.tab
\end{verbatim}
where we refer to Ref.~\cite{Smirnov:2014hma} for the precise meaning of the individual
commands. The integrals which shall be reduced can be found in
the file {\tt top2l2h1a.ind} which might have the form
\begin{verbatim}
{{1, {1, 1, 1, 2, 2, 2, 2}},
{1, {1, 1, 1, 1, 1, 1, 2}},
{1, {1, 1, 1, 1, 1, 1, -1}}}
\end{verbatim}
Here the individual entries are lists where the integer in the first
entry numbers the family and the second entry contains seven integers
specifying the indices of the propagators as specified above (see
``\verb|Propagators|'').
The reduction is initiated with the help
of {\verb|./FIRE5 -c top2l2h1a|}. After the job is completed the reduction table
can be found in the file {\tt top2l2h1a.tab}
which can be read using again a {\tt Mathematica} session of {\tt FIRE}.
There are several benchmark calculations which have been performed with the
help of {\tt FIRE}. Among them is the reduction of all three-loop integrals
needed for the static
potential~\cite{Smirnov:2008pn,Smirnov:2009fh,Anzai:2009tm} which
involves eight indices for massless relativistic propagators and in addition
three indices for static propagators of the form $1/k_0$. A particular
challenge poses the case for general QCD gauge parameter $\xi$ which involves
about 20 million integrals, 60 times as much as the $\xi=0$ case. A further reduction
problem involves four-loop on-shell integrals needed for the relation between
the $\overline{\rm MS}$ and on-shell quark mass relation or the electron
anomalous magnetic moment (see, e.g., Ref.~\cite{Lee:2013sx}).
\subsection{Numerical evaluation of master integrals with {\tt FIESTA}}
{\tt FIESTA}~\cite{Smirnov:2008py,Smirnov:2009pb,Smirnov:2013eza} stands for
Feynman Integral Evaluation by a Sector decomposiTion Approach and is a
convenient tool to numerically evaluate Feynman integrals using the method of
sector decomposition. The latter is an algorithmic procedure to extract the
$\epsilon$ poles of a given Feynman integral in the so-called
alpha-representation and provide an integral representation for the
coefficients. After the pioneering work of Binoth and
Heinrich~\cite{Binoth:2003ak,Binoth:2004jv} several programs have been
published where different strategies have been implemented. Among them are
{\tt sector\_decomposition}~\cite{Bogner:2007cr}, {\tt
secdec}~\cite{Binoth:2004jv,Borowka:2012yc,Borowka:2013cma}, and {\tt
FIESTA}~\cite{Smirnov:2008py,Smirnov:2009pb,Smirnov:2013eza}.
The basic philosophy of {\tt FIESTA} is that all kinematic variables are
specified at an early stage which is different from other approaches like,
e.g., {\tt secdec}, where generic manipulations are performed up to a certain
point and only then numerical values for masses and momenta are specified.
The use of {\tt FIESTA} splits into the following two steps:
In a first step the momentum integrals are transformed into the
alpha-representation and the sector decomposition algorithm is applied. The
corresponding manipulations are performed in {\tt Mathematica} and can be done
in parallel mode.
For many applications this step is quite fast, however, quite often, in
particular at higher loop order, huge expressions are generated which require
main memory in the range of hundred Gigabyte. In such cases it is
convenient to store the results into a database~\cite{kyotocabinet} since in
general this step has to be performed only once.
The second step is concerned with the numerical integration. In principle
this can also be performed within {\tt Mathematica}, which is advantageous for
small problems or during the developing phase of the program. Complicated problems
have to be integrated with the help of a {\tt C++} integrator which is based
on the {\tt Cuba} library~\cite{Hahn:2004fe,cuba}. It uses the expressions
stored in the database during step one which provides several advantages. For
example, it is possible to perform various runs choosing different values for
the number of points used for the integrations. Furthermore, it is possible to
copy the output of step one to a platform which is suitable for the numerical
integration in massive parallel mode.
\begin{figure}[t]
\begin{center}
\includegraphics[width=.8\linewidth]{diag_4los}
\caption{\label{fig::diag_4los}
Sample on-shell Feynman diagram where solid and dashed lines denote massive
and massless lines, respectively.
}
\end{center}
\end{figure}
Let us as an example consider the Feynman diagram in Fig.~\ref{fig::diag_4los}
which enters the four-loop relation between the $\overline{\rm MS}$-on-shell
quark mass. Executing the {\tt Mathematica} file
\begin{verbatim}
Get["FIESTA3.m"];
NumberOfSubkernels=8;
NumberOfLinks=8;
UsingC=True;
UsingQLink=True;
ComplexMode=False;
SDEvaluate[ UF[ {k1,k2,k3,k4},
{-(k1+q1)^2+m^2,
-(k3+q1)^2+m^2,
-(k1-k2)^2+m^2,
-(k2-k3)^2+m^2,
-(k1-k4)^2+m^2,
-k4^2+m^2,
-k3^2},
{m->1,q1^2->1}],
{1,1,1,1,1,1,1},6]
\end{verbatim}
prepares both the integrand and performs the numerical integration using
the corresponding {\tt C} routines in the background. The result
which is printed on the screen reads
\begin{verbatim}
-276.907674 - 0.625006/ep^4 -
4.937615/ep^3 + (-24.441689 +
0.002*pm69)/ep^2 + (-85.919995 +
0.015937*pm70)/ep + 0.083469*pm71 +
ep*(-864.271585 + 0.468742*pm72) +
ep^2*(-1503.357843 + 2.093833*pm73) +
ep^3*(-6224.681821 + 9.755544*pm74) +
ep^4*(11328.088699 + 40.591518*pm75) +
ep^5*(-18622.607506 + 176.767061*pm76) +
ep^6*(537473.776134 + 713.790523*pm77)
\end{verbatim}
The symbols \verb|pm| indicate the uncertainty
due to the Monte Carlo integration.
In case the option \verb|OnlyPrepare = True;| is added
to the {\tt Mathematica} file the integrand is prepared
and stored to disk. Furthermore the command
is printed on screen which invokes the
numerical integration from the shell without
reference to {\tt Mathematica}.
The result from runs performed at the High Performance Computing Center
Stuttgart (HLRS) are shown in Fig.~\ref{fig::FIESTA_speedup} where the speedup
for the individual $\epsilon^n$ terms ($n=-3,\ldots,6$) is shown. The blue
(dotted), green (dashed), red (dash-dotted) and black (solid) curves (from
bottom to top) corresponds to the use of 64, 128, 256 and 512 cores where the
results have been normalized to the 32-core run. It is interesting to note
that an ideal speedup is obtained for 64 cores. Also for 128 cores the curve
is close to the maximal value of 4. Using 256 instead of 32 cores still shows a
quite flat behaviour with a speedup between 6 and 7. Strong variations in the
speedup are observed for the use of 512 cores. The relatively low value for
$1/\epsilon^{-3}$ can be explained with the fact that probably the expression,
which shall be integrated, is too simple. On the other hand, for the
(complicated) expression of the $\epsilon^{6}$ coefficient it might be that
the disk access becomes the bottle neck.
\begin{figure}[t]
\begin{center}
\includegraphics[width=1.\linewidth]{fiesta_speedup}
\caption{\label{fig::FIESTA_speedup}
Speedup of the calculation of the various $\epsilon$ orders
of Feynman diagram given in Fig.~\ref{fig::diag_4los} using
64 (dotted), 128 (dashed), 256 (dash-dotted) and 512 (solid) cores
normalized to the 32-core run.
}
\end{center}
\end{figure}
The main purpose of {\tt FIESTA} is the fast and convenient cross check of
analytic calculations. Within CRC/TR~9 is has been applied in this way
to several problems. An early version of {\tt FIESTA} has been used to cross
check the master integrals which contribute to the three-loop static
potential~\cite{Smirnov:2008pn,Smirnov:2009fh,Anzai:2009tm}. Furthermore
thirteen four-loop on-shell integrals
contributing to the $\overline{\rm MS}$-on-shell quark mass relation and to
the muon anomalous magnetic moment, which have been computed analytically in
Ref.~\cite{Lee:2013sx}, have been cross-checked numerically with {\tt FIESTA}.
Recently also analytic results for master integrals
of double-box topologies in the physical region have been cross-checked with
the help of {\tt FIESTA}~\cite{Caola:2014lpa}.
There are also several projects where {\tt FIESTA} has been used to evaluate
the most complicated or even the major part of the master integrals
numerically. For example, in the first calculation of the three-loop
corrections of the quark and gluon form factor~\cite{Baikov:2009bg} (see also
Ref.~\cite{Gehrmann:2010ue}) one coefficient in the $\epsilon$ expansion of
the three most complicated integrals could not be evaluated
analytically. Thus, the numerical results of {\tt FIESTA} have been used
which, for all practical purposes, leads to final results with sufficient
precision. The analytic calculation of the missing master integrals has been
performed in Ref.~\cite{Lee:2010ik} and perfect agreement with the numerical
result has been found.
For the calculation of the three-loop matching coefficient between QCD and
non-relativistic QCD (NRQCD) of the vector current~\cite{Marquard:2014pea}
even the majority of the about 100 master integrals have been computed
numerically with the help of {\tt FIESTA}. In such cases it is important to
perform strong cross checks. Among them are the change of the parametrization
of the individual integrals. Thus, in intermediate steps different expressions
are generated which are then integrated numerically. Furthermore, it is
possible to choose a different integrals basis and evaluate the new integrals
again with the help of {\tt FIESTA}. The agreement of the final expression
within the numerical uncertainty among the two set of master integrals serves
as a strong checked for the applicability of {\tt FIESTA}.
\section{\label{sec::sum}Summary}
The computer algebra program {\tt FORM}{} is designed to handle huge expressions
in a quite effective way. Still, for some physical applications even
{\tt FORM}{} would take several years which make a practical calculation
impossible.
In the recent years parallel versions of {\tt FORM}{}, {\tt ParFORM}{} and {\tt TFORM}{},
have been developed and in the meantime they have become a reliable tools to
perform computer algebra in parallel. {\tt ParFORM}{} has demonstrated a good
speedup behaviour both on SMP computers and on different cluster
architectures. Furthermore, for the current version of {\tt ParFORM}{} the {\tt FORM}{}
programs written for the sequential version need not to be modified.
{\tt TFORM}{} is a parallel version of {\tt FORM}{} based on {\tt POSIX} threads
and thus is bound to run on a single node. However, there is less overhead
connected to the parallelization and thus {\tt TFORM}{} shows a slightly better
performance than {\tt ParFORM}{}.
The main advantage of using a parallel version of {\tt FORM}{} is the reduction of
the wall clock time. In fact, there are a number of calculations where it has
been exploited that a speedup of about 10 can be reached with 16 cores
and thus the result was available after about a month instead of a year.
A further advantage of using {\tt TFORM}{} or {\tt ParFORM}{} is the fact that
the size of the intermediate results, which have to be handled by the
individual CPU, is smaller since the workload is
distributed among several workers. This advantage becomes particularly
evident when using {\tt ParFORM}{} on a cluster. In that case the intermediate
expressions are stored into files which are located on different nodes.
To obtain an even better speedup behaviour it would be necessary
to improve the slope of the speedup curves and to push the
flattening to higher number of processors. One starting point which could help to
improve the situation is the sorting procedure.
Another idea might be the combination of {\tt ParFORM}{} and {\tt TFORM}{}
which could be an ideal tool for a cluster with multi-core nodes.
In this article we also describe the programs {\tt FIRE} and {\tt FIESTA}.
{\tt FIRE} can be used for the reduction of integrals belonging to a given
integral family to master integrals. {\tt FIESTA}, on the other hand
is a user-friendly tool to numerically compute the coefficients of the
$\epsilon$ expansion of multi-loop integrals.
\section*{Acknowledgements}
This work is supported by the Deutsche
Forschungsgemeinschaft in the Sonderforschungsbereich Transregio~9
``Computational Particle Physics''.
We acknowledge the use of the High Performance Computing Center Stuttgart
(HLRS) where part of the calculations connected to {\tt FIESTA} have
been carried out. In this context we also acknowledge the help of Peter
Marquard.
|
\section{INTRODUCTION}\label{sec:intro}
The simplest and probably the most reasonable explanation for the current acceleration of the universe is the anti-gravitational effect of
the cosmological constant ($\Lambda$) that is supposed to be dominant at the present epoch. The fundamental assumption that underlies
this explanation is that the general relativity (GR) is the universal law of gravity, the validity of which has been confirmed by recent
cosmological observations \citep{reyes-etal10,wojtak-etal11,rapetti-etal13}. Besides, the latest news from the Planck CMB (Cosmic
Microwave Background) mission has drawn the following bottom line \citep{planck_mg15}:
The risk that one has to take by accepting the $\Lambda$CDM ($\Lambda$+cold dark matter) cosmology founded upon GR
is lower than that by accepting any other alternative.
Nevertheless, the conceptual difficulty to acquiesce in the extremely fine-tuned initial conditions of the universe associated
with $\Lambda$ still dares the cosmologists to take a high risk of suggesting modified gravity (MG) models and developing independent
tests of GR. One of the most plausible tests of GR on the cosmological scale is to compare the lensing mass of a cluster with its dynamic
mass \citep[e.g., see][]{schmidt-etal09,schmidt10,zhao-etal11}. The former is estimated from the weak gravitational lensing effect
generated by the mass of a cluster on the shapes of the foreground galaxies, while the latter is conventionally derived from the velocity
dispersions of the central satellites of a cluster. A substantial discrepancy between the two mass measurements for a cluster,
if found and accurately measured, would challenge the key tenet of GR.
However, it is a formidable task to measure the dynamical mass of galaxy clusters with high accuracy. The success of the conventional
method of using the satellite velocity dispersions is contingent upon the completion of relaxation in the dynamical state and the spherical
symmetry in the shape. There is another method for the dynamical mass measurement of the clusters that uses the peculiar velocities
of the neighbor galaxies located beyond the virial radius of a cluster. This method can be applied even to those clusters which have yet
to be completely relaxed without having spherical symmetry \citep{corbett-etal14}. However, large uncertainties inherently involved in the
measurements of the peculiar velocity field remain as a technical bottleneck that stymies the accurate measurement of the dynamic
mass of a cluster from the peculiar velocities of the neighbor galaxies
\citep[see also][]{corbett-etal14,gronke-etal14b,hellwing-etal14,zu-etal14}
\citet{falco-etal14} devised a novel method to measure the dynamic mass of a cluster by using the {\it universal} radial velocity profile of
the filament galaxies around the cluster. The sole condition for the practical success of their method is that the neighbor galaxies
should be distributed along a narrow filament. Since it has been theoretically and observationally shown that the formation of galaxy
clusters preferentially occur at the junction of the cosmic filaments \citep[e.g., see][]{web96,dietrich-etal12}, the majority of the galaxy
clusters should meet this condition, being eligible for the application of this method.
The method of \citet{falco-etal14} to measure the virial mass of a cluster from the radial velocity profile of the filament galaxies is applicable
even to those clusters which have yet to complete its relaxation process. In this respect their method fits especially well for the Virgo cluster
that still undergoes merging events. Furthermore, the measurements of the redshifts and equatorial coordinates of the galaxies in the
vicinity of the Virgo cluster were done with high accuracy thanks to their proximity, which will help reduce uncertainties in the dynamical
mass measurement with the method of \citet{falco-etal14} for the Virgo cluster.
The goal of this paper is to test this new method against the Virgo cluster around which the narrow filamentary structures were recently
identified by \citet{kim-etal15}.
The upcoming sections will present the following. A review of the method of \citet{falco-etal14} in Section \ref{sec:f14}.
A reconstruction of the radial velocity profile of the filament galaxies around the Virgo cluster and a new dynamical mass
estimate for the Virgo cluster by comparing the reconstructed profile with the numerical formula from \citet{falco-etal14} in Section
\ref{sec:mv}. Discussions on the final results and on the future works for the improvements in Section \ref{sec:con}.
Throughout this paper, we assume a GR+$\Lambda$CDM cosmology with $\Omega_{m}=0.26,\ \Omega_{\Lambda}=0.74,\ h=0.73$
and zero curvature, to be consistent with \citet{falco-etal14}.
\section{METHODOLOGY : A BRIEF REVIEW}
\label{sec:f14}
The galaxies located in the vicinity of a cluster must feel the gravity of the cluster even in the case that they are not
the satellites bound to the cluster, and thus their motions relative to the cluster should deviate from the pure Hubble flows.
To quantify the degree of the deviation of the galaxy motions from the Hubble flows and its dependence on the cluster
mass, \citet{falco-etal14} determined the mean radial velocity profile of the galaxies in the vicinity of the clusters
from a high-resolution N-body simulation for the GR+$\Lambda$CDM cosmology. Their numerical experiment has discovered that
the mean radial velocity profile of the neighbor galaxies located in the zone where the effect of the Hubble expansion exceeds
that of the gravitational attraction of a cluster is well approximated by the following universal formula:
\begin{equation}
\label{eqn:vr}
v_{r}(d; z, M_{\rm v}) = H(z)\, d - 0.8\,V_{\rm v}\,\left(\frac{r_{\rm v}}{d}\right)^{n_{\rm v}}\, ,
\end{equation}
where $v_{r}$ is the mean radial velocity at (comoving) separation distance $d$ from the cluster center, $M_{\rm v}$, $r_{\rm v}$ and
$V_{\rm v}$ are the virial mass, radius and velocity of the cluster, respectively, and $H(z)$ is the Hubble parameter at redshift $z$.
The first term in the right-hand side represents a pure Hubble flow from the cluster center while the second term corresponds to
the mean peculiar velocity induced by the gravitational attraction of the cluster.
According to \citet{falco-etal14}, the radial velocity profile is universal in the respect that the power-law index, $n_{\rm v}$,
in Equation (\ref{eqn:vr}) has a constant value of $0.42$, being independent of $M_{\rm v}$ and $z$.
They also showed that the zone where the radial velocities of the galaxies are well approximated by Equation (\ref{eqn:vr})
with $n_{\rm v}=0.42$ corresponds to the range of $3\,r_{\rm v}\le d\le 8\,r_{\rm v}$. Adopting the conventional definition that $r_{\rm v}$
is the radius of a sphere within which the mass density reaches $\Delta_{c}$ times the critical density of the Universe
$\rho_{c}\equiv 3H^{2}/(8\pi\,G)$, \citet{falco-etal14} used the following simple relations to calculate the virial mass $M_{\rm v}$ and
the virial velocity $V_{\rm v}$ for Equation (\ref{eqn:vr}):
\begin{equation}
\label{eqn:mv_rv}
M_{\rm v} = \frac{4\pi}{3}\Delta\,\rho_{c}\,r^{3}_{\rm v},\ \quad V^{2}_{\rm v} = \frac{G\, M_{\rm v}}{r_{\rm v}}\, .
\end{equation}
The value of $\Delta_{c}$ was set at $93.7$ in accordance with the formula of
$\Delta_{c}=18\pi^{2}+82[\Omega_{m}(z)-1]-39[\Omega_{m}(z)-1]^{2}$ given by \citet{BN98}.
\citet{falco-etal14} suggested that the virial mass of a massive cluster $M_{v}$ be estimated by comparing the observed radial velocity
profile of its neighbor galaxies to Equation (\ref{eqn:vr}). However, since the real values of $v_{r}$ and $d$ are not directly measurable,
it is difficult to determine the radial velocity profiles of the galaxies from observational data. They claimed that this difficulty should be
overcome by using the anisotropic spatial distribution of the galaxies belonging to a one-dimensional filament in the vicinity of a cluster.
Three dimensional separation distance $d$ and the radial velocity $v_{r}$ of the filament galaxies can be estimated from
two direct observables, the projected distance $R$ in the plane of the sky and the line-of-sight velocity $v_{\rm los}$, as
$R=\sin\beta\, r$ and $v_{r}=v_{\rm los}/\cos\beta$, respectively, where $\beta$ is the angle at which the filament is inclined to the line of
sight direction to a cluster (see Figure \ref{fig:illustration}). Finally, Equation (\ref{eqn:vr}) was rewritten in terms of the observables as
\begin{equation}
\label{eqn:2dvr}
v_{\rm los}(R,\beta,M_{\rm v}) =
\cos\beta\left[ H(z)\frac{R}{\sin\beta} - 0.8V_{\rm v}\left(\frac{R}{\sin\beta\,r_{\rm v}} \right)^{n_{\rm v}}\right]\, .
\end{equation}
Since the galaxy clusters usually form at the junction of the cosmic filaments, this new scheme of \citet{falco-etal14} seems quite plausible.
However, the validity of Equation (\ref{eqn:2dvr}) depends strongly on the geometrical shape of the filament.
The more straight and narrow a filament is, the better Equation (\ref{eqn:vr}) is approximated by Equation (\ref{eqn:2dvr}).
Figure \ref{fig:illustration} illustrates a three dimensional configuration of a filamentary structure around a cluster inclined at an angle
$\beta$ to the line of sight direction to the cluster, defining the projected distance $R$ between a filament galaxy and the cluster center
in the plane of sky.
\citet{falco-etal14} performed a numerical test of Equation (\ref{eqn:2dvr}) by applying it to three cluster-size halos around which filamentary
structures were found. The estimated values of $M_{\rm v}$ turned out to agree very well with the true virial masses for all of the three
cases, which numerically verified the validity of Equations (\ref{eqn:2dvr}).
\citet{falco-etal14} also performed an observational test of their algorithm by applying it to the sheet structures identified around the Coma
cluster.
Despite their claim that the dynamic mass of the Coma cluster determined by their method turned out to be consistent with the previous
estimates, their method, strictly speaking, is valid only for the galaxies distributed along a one-dimensional filamentary
structure that interacts gravitationally with the cluster. In fact, it is obvious from Equation (\ref{eqn:2dvr}) that large uncertainties should
contaminate the measurement of the cluster mass if the radial velocity profile is reconstructed from the sheet galaxies.
\section{ESTIMATING THE DYNAMIC MASS OF THE VIRGO CLUSTER}
\label{sec:mv}
\subsection{The Zero-th Order Approximation}\label{sec:mv0}
\citet{kim-etal15} developed an efficient technique to identify a filamentary structure whose galaxies are under the
gravitational influence of a neighbor cluster. The merit of their filament-finding technique is that it requires information only on
observable redshifts and two dimensional positions of the galaxies projected onto the plane of sky. Applying this filament-finding
technique to the NASA-Sloan-Atlas catalog of the local galaxies\footnote{given in http://www.nsatlas.org/data},
\citet{kim-etal15} have identified two narrow filamentary structures to the east and to the west of the Virgo cluster (say, the Filament A and
B, respectively). Both of the Filament A and B were found to consist mainly of the dwarf galaxies with absolute $r$-band magnitudes
$M_{r}> -20$. For a detailed description about how the Filament A and B were identified from the NASA-Sloan-Atlas catalog, see \citet{kim-etal15}.
Figure \ref{fig:dec_ra} plots the right asensions (R.A.) and declinations (decl.) of the galaxies belonging to the filament
A and B, as filled blue and red dots, respectively. The certain and possible member galaxies of the Virgo cluster from the
Extended Virgo Cluster Catalog (EVCC) \citep{kim-etal14} are shown as the black and gray dots, respectively, inside an open
rectangular box. The gray dots outside the box represent the field galaxies that do not belong to the EVCC
nor to the two filaments but in the redshift range of $500\le cz/[{\rm km}\,s^{-1}] \le 2500$.
Note that the Filament A appears to be more straight than the Filament B in the equatorial plane.
Showing that the mean radial velocities of the galaxies belonging to each filament are similar to the system velocity of the Virgo cluster,
\citet{kim-etal15} have confirmed that the filament galaxies are in gravitational interaction with the Virgo cluster.
Figure \ref{fig:cz_dis} plots the redshifts of the galaxies belonging to the Filament A and B as a function of their projected distances
in unit of degree from the M87 (i.e., the Virgo center) as filled blue and red dots, respectively. As can be seen, the redshift variation
of the filament galaxies $\Delta\,cz$ at the constant $R$ are less than $1000$ km/s.
Denoting the equatorial positions (i.e., R.A. and decl.) of a filament galaxy and the Virgo center as $(\alpha_{\rm g},\ \delta_{\rm g})$
and $(\alpha_{\rm cl},\ \delta_{\rm cl})$, respectively, we determine the projected distance $R$ between
each filament galaxy from the Virgo center in the plane of sky (see Figure \ref{fig:illustration}) as
$R = c\,z_{\rm cl}\left(\cos\theta_{\rm g}\cos\theta_{\rm cl} + \sin\theta_{\rm g}\sin\theta_{\rm cl}\cos\Delta\phi_{\rm g}\right)$ where
$\theta_{\rm g}=\pi/2-\delta_{\rm g}$, $\theta_{\rm cl}=\pi/2-\delta_{\rm cl}$, $\Delta\phi_{\rm g}=\alpha_{\rm g}-\alpha_{\rm cl}$, and
$z_{\rm g}$ and $z_{\rm cl}$ are the redshifts of the filament galaxy and the Virgo center, respectively.
The line-of-sight velocity magnitude $v^{\rm ob}_{\rm los}$ of each filament galaxy at a projected distance, $R$, from the Virgo center
is approximated as $c\vert\,z_{\rm g}(R) - z_{\rm cl}\,\vert$.
Equation (\ref{eqn:2dvr}) is valid only for those filament galaxies whose separation distances from the cluster center are larger than three
times the virial radius of the Virgo cluster according to \citet{falco-etal14}. Taking the typical cluster size $2\,h^{-1}$Mpc as a conservative
upper limit on the virial radius of the Virgo cluster, we select only those filament galaxies which satisfy the condition of $R\ge 6\,h^{-1}$Mpc.
A total of $130$ and $96$ galaxies are selected in the Filament A and B, respectively, to each of which the method of \citet{falco-etal14}
is applied as follows.
Setting the power-law index $n_{\rm v}$ at the universal value of $0.42$, we compare the observed values of $v^{\rm ob}_{\rm los}$ of the
filament galaxies with Equation (\ref{eqn:2dvr}). Assuming a flat prior, we search for the best-fit values of $M_{\rm v}$ and $\beta$ which
maximize the following posterior density function $p(M_{\rm v}, \beta | {\rm data})$:
\begin{equation}
\label{eqn:pos_mv_beta}
p(M_{\rm v}, \beta | v^{\rm o}_{\rm los},R_{i}) \propto \exp\Bigg{\{}-\sum_{i=1}^{N_{g}}
\frac{[v^{\rm o}_{\rm los}(R_{i}) - v_{\rm los}(R_{i}; M_{\rm v},\ \beta)]^{2}}{2\sigma^{2}_{\rm v}(N_{\rm g}-2)}\Bigg{\}}\, ,
\end{equation}
where $N_{\rm g}$ is the number of the selected filament galaxies, $R_{i}$ is the projected distance of the $i$-th filament galaxy,
$\sigma_{\rm v}$ represents errors involved in the measurement of $v^{\rm ob}_{los}(R_{i})$ from the redshift space.
Given that $\sigma_{\rm v}$ include all systematic errors that must be existent but unknown, we set $\sigma_{\rm v}$ at unity for all of the
selected filament galaxies. Here, the posterior probability density function $p(M_{\rm v}, \beta | v^{\rm o}_{\rm los},R_{i})$ is normalized to
satisfy the condition of $\int dM_{\rm v}\int d\beta\, p(M_{\rm v}, \beta | v^{\rm ob}_{\rm los},\ R_{i}) = 1$.
The $68\%,\ 95\%$ and $99\%$ confidence level contours of $\log M_{\rm v}$ and $\beta$ obtained from the above fitting procedure
for the case of the Filament A (B) are shown as solid, dashed and dot-dashed lines, respectively, in the left (right) panel of Figure
\ref{fig:mv_beta}, which reveals the existence of a strong degeneracy between $M_{v}$ and $\beta$. To break this degeneracy, we would
like to put some prior on the range of $\beta$ by making the zeroth-order approximation to the three dimensional positions of the
filament galaxies relative to the Virgo cluster.
In the zero-th order approximation of $d\approx d^{0}=c\,z_{\rm g}$ under the assumption of no peculiar velocity, the three dimensional
separation vector of each selected filament galaxy, ${\bf d}^{0}=(d_{1}^{0},\ d_{2}^{0},\ d_{3}^{0})$, can be expressed as
$d_{1}^{0} = c\,z_{g}\sin\delta_{\rm g}\cos\alpha_{\rm g} - c\,z_{\rm cl}\sin\delta_{\rm cl}\cos\alpha_{\rm cl},\
d_{2}^{0} = c\,z_{g}\sin\delta_{\rm g}\sin\alpha_{\rm g} - c\,z_{\rm cl}\sin\delta_{\rm cl}\sin\alpha_{\rm cl},\
d_{3}^{0} = c\,z_{g}\cos\delta_{\rm g} - c\,z_{\rm cl}\cos\delta_{\rm cl}$. The zeroth-order approximation to the angle, $\beta$,
between ${\bf d}$ and the line of sight direction to the Virgo cluster, say $\hat{\bf h}$, is now calculated as,
$\beta\approx \beta^{0}\equiv \cos^{-1}\langle\hat{\bf d}^{0}\cdot\hat{\bf h}\rangle$ with
$\hat{\bf d}^{0} \equiv {\bf d}^{0}/\vert{\bf d}^{0}\vert$ where the ensemble average is taken over the selected filament galaxies.
The Filament A and B yield $\beta^{0}=36.8^{\circ}\pm 3.9^{\circ}$and $\beta^{0}=36.9^{\circ}\pm 5.6^{\circ}$, respectively.
It is interesting to note that this result is in line with the previous observational evidences for the alignment tendency between
the major axis of the Virgo cluster and the line-of-sight direction \citep{mei-etal07}. It is also consistent with the recent detection
that the Virgo satellites tend to fall into the Virgo along the lie of sight directions aligned with the minor principal axis of the local
velocity shear field \citep{lee-etal14}.
Taking these zero-th order values as a prior, we break the degeneracy between $M_{\rm v}$ and $\beta$
shown in Figure \ref{fig:mv_beta} and finally determine the dynamical mass of the Virgo cluster:
$M_{\rm v}\approx (0.84^{+2.75}_{-0.51})\times 10^{15}\,h^{-1}M_{\odot}$ for the case of the Filament A and
$M_{\rm v}\approx (3.2^{+5.0}_{-1.3})\times 10^{15}\,h^{-1}M_{\odot}$ for the case of the Filament B.
\subsection{Radial Motions of the Filament Galaxies around the Virgo Cluster}\label{sec:vr}
It should be worth comparing our result with the previous measurements of the Virgo mass based on various methods.
For instance, \citet{urban-etal11} converted the X-ray temperature of the hot gas in the Virgo cluster to its mass via the mass-temperature
scaling relation obtained by \citet{arnaud-etal05} and found $M_{200}/M_{\odot}\approx 1.4\times 10^{14}$. Here $M_{200}$ represents the
mass enclosed by a spherical radius $r_{200}$ at which the mass density becomes $\rho(r_{200})=200\,\rho_{crit}$ where
$\rho_{crit}\equiv 3H^{2}/8\pi G$.
\citet{karachentsev-etal14} studied the infall motions of seven nearby galaxies located along the line-of-sight direction to the
Virgo cluster whose distances were measured by means of the Tip of the Red Giant Branch (TRGB) method \citep{baade44,SC97} and
found $M_{\rm 200}/M_{\odot}=(7.0\pm 0.4)\times 10^{14}$. It was also claimed that the lensing mass of the Virgo cluster should be in the
range of $\log\left(M_{200}/M_{\odot}\right)=13.87^{+0.42}_{-1.95}$ by employing the technique developed by \citet{kubo-etal09} to estimate
the lensing mass of the galaxy clusters in the local universe
(D. Nelson 2014, private communication\footnote{see also the presentation given in
https://www.cfa.harvard.edu/~dnelson/doc/dnelson.fermilab.talk.pdf. }).
Noting that the definition of $M_{200}$ is different from that of the virial mass $M_{\rm v}$ considered here, we convert the values of
$M_{200}/M_{\odot}$ reported in the previous works into $M_{\rm v}/(h^{-1}\,M_{\odot})$ for a fair comparison under the assumption that
the mass density profile of the Virgo cluster is well approximated by the NFW formula \citep{nfw97} and list them in Table \ref{tab:mv}.
As can be read, our best-fit value $M_{\rm v}$ obtained from the Filament A is consistent with the previous dynamic mass estimated
by the TRGB-distance method but substantially higher than the other two estimates. Meanwhile, for the case of the Filament B,
the best-fit $M_{\rm v}$ is significantly off not only from the previous estimates but from the best-fit value obtained for the case of the
Filament A. Given that the method of \citet{falco-etal14} is strictly valid only for the galaxies distributed along thin straight filament,
we suspect that the bent shape of the Filament B should be responsible for the significant discrepancy on $M_{\rm v}$
between the two filaments and suggest that the Filament B should not be eligible for the application of the method of \citet{falco-etal14}.
Before excluding the Filament B from our analysis, it may be worth investigating how different the reconstructed velocity profile
of the Virgo cluster is from the expected profile, i.e., the universal formula given by \citet{falco-etal14} when the mass of the Virgo
cluster is fixed at some constant value $M_{\rm v}$.
Treating $n_{\rm v}$ in Equation (\ref{eqn:2dvr}) as a free parameter, we search for the best-fit values of $n_{\rm v}$ and $\beta$ that
maximize the following posterior distribution:
\begin{eqnarray}
p(n_{\rm v}, \beta | v^{\rm ob}_{\rm los}, R_{i}) &\propto& p({\rm data} | n_{\rm v}, \beta) p(\beta) \, \nonumber \\
\label{eqn:pos_nv_beta}
&\propto&
\exp\left\{-\frac{[v^{\rm o}_{\rm los}(R_{i}) - v_{\rm los}(R_{i}; n_{\rm v}, \beta)]^{2}}{2(N_{g}-2)}\right\}
\frac{1}{\sigma_{\beta^{0}}}\exp\left[-\frac{\left(\beta-\bar{\beta^{0}}\right)^{2}}{2\sigma^{2}_{\beta^{0}}}\right]\, .
\end{eqnarray}
Here, instead of setting $\beta$ at the fixed value $\bar{\beta^{0}}$, we assume a Gaussian prior on $\beta$, setting its
mean and standard deviation at $\bar{\beta^{0}}$ and $\sigma_{\beta 0}$, respectively, both of which are determined in Section \ref{sec:mv0}.
Figure \ref{fig:nv_betaA} shows the $68\%,\ 95\%$ and $99\%$ confidence level contours of $\beta$ and $n_{\rm v}$ obtained
from the Filament A as solid, dashed and dot-dashed lines, respectively, for the four different cases of the dynamical mass
of the Virgo cluster: $M_{\rm v} = 5\times 10^{13},\ 10^{14},\ 5\times 10^{14},\ 10^{15}\, h^{-1}M_{\odot}$.
As can be seen, for the cases of $M_{\rm v}< 10^{14}\,h^{-1}M_{\odot}$ that is consistent with the previous estimate of
the lensing mass (see Table 1), the original value of $n_{\rm v}=0.42$ is located beyond the $95\%$ confidence level contours.
For the case of $M_{\rm v}\ge 5\times 10^{14}\,h^{-1}M_{\odot}$ that is consistent with the dynamical mass estimate obtained by the
TRGB method, $n_{\rm v}=0.42$ is included within the $68\%$ confidence level contour. Hence, one can conclude that the numerical
formula of the radial velocity profile suggested by \citet{falco-etal14} matches fairly well with the reconstructed profile from observation
for the case of the Filament A around the Virgo cluster, yielding the best-fit value of $M_{\rm v}$ consistent with the previous
dynamical mass measurement for the Virgo. Figure \ref{fig:nv_betaB} plots the same as Figure \ref{fig:nv_betaA} but from the Filament B.
As can be seen, for all of the four cases of $M_{\rm v}$, the original value of $0.42$ stays outside the $95\%$ contours, indicating that
Equation (\ref{eqn:2dvr}) fails to describe the radial motions of the galaxies in the Filament B.
We also show the radial velocity profile, $v_{r}(d)$, reconstructed from the Filament A as black solid line in Figure \ref{fig:vprofileA}.
To plot $v_{r}$, the Virgo mass $M_{\rm v}$ is set at $10^{14}\,h^{-1}M_{\odot}$ (close to the lensing mass),
the slope $n_{\rm v}$ and the angle $\beta$ are set at the best-fit values determined by the above fitting procedure.
The associated $1\sigma$ and $2\sigma$ uncertainties around the reconstructed profile are also shown as dark and light grey regions,
respectively, and the red dashed line represents the expected profile with the fixed value of $n_{v}=0.42$ claimed by \citet{falco-etal14}.
As can be seen, the expected profile is only marginally consistent with the reconstructed one if $M_{\rm v}=10^{14}\,h^{-1}M_{\odot}$.
Figure \ref{fig:vprofileB} plots the same but for the case of the Filament B. It is clear that the difference between the numerical profile
and the reconstructed one becomes much larger for the case of the Filament B.
\section{DISCUSSION AND CONCLUSION}
\label{sec:con}
We have measured the dynamic mass of the Virgo cluster, $M_{\rm v}$, by applying the method of \citet{falco-etal14} to two narrow
filaments (the Filament A and B) consisting mainly of dwarf galaxies located to the east and to the west of the Virgo cluster in the
equatorial reference frame, which were recently identified by \citet{kim-etal15} from the NASA-Sloan-Atlas catalog.
The method of \citet{falco-etal14} that can be applied even to the unrelaxed clusters like the Virgo is based on the numerical findings
that the radial velocity profile of the galaxies at distances larger than three times the virial radius of a neighbor cluster has a universal
shape and that it can be readily reconstructed from direct observables as far as the galaxies are distributed along a filamentary structure.
To break a strong degeneracy found between the value of $M_{\rm v}$ and the angle $\beta$, we have constrained the ranges of $\beta$ by
making a zero-th order approximation in which the redshift-space positions of the filament galaxies are assumed to equal their real-space ones.
Our estimates based on the method of \citet{falco-etal14} has yielded
$M_{\rm v}=(0.84^{+2.75}_{-0.51})\times10^{15}\,h^{-1}M_{\odot}$ and $M_{\rm v}= (3.2^{+5.0}_{-1.3})\times 10^{15}\,h^{-1}M_{\odot}$
from the Filament A and B, respectively.
Given that the Filament B appears not so straight as the Filament A in the equatorial plane, the substantial difference between the two
results on the value of $M_{\rm v}$ from the Filament A and B has been suspected as an indication that the Filament B is not eligible
for the algorithm of \citet{falco-etal14} whose success depends sensitively on the degree of the filament straightness.
The method of \citet{falco-etal14} itself also suffers from a couple of unjustified assumptions. First of all, it assumes the filaments to be
extremely narrow and straight, which is an obvious over-simplication of the true shapes of the filaments. Second of all, in their original
study, the universal radial velocity profile was obtained by taking an average over all the particles and halos around the clusters but not only
over the filament halos. If only a subsample composed of the filament galaxies is used to reconstruct the radial velocity, then it would not
necessarily agree with the average universal profile. Besides, the zero-th order approximation that we have made for $\beta$ in our
estimation of the Virgo mass with the method of \citet{falco-etal14} is an additional simplification of the reality. If a broader range of
$\beta$ were considered, then the mismatches between the two masses would be likely reduced.
Our estimate of $M_{\rm v}$ from the Filament A has turned out to be consistent with the previous dynamical mass estimates based on
the TRGB-distance method by \citet{karachentsev-etal14} but three to four times higher than the other mass estimates based on the weak
gravitational lensing effect (D. Nelson 2014 in private communication) and the X-ray temperature of hot gas \citep{urban-etal11}.
The usual suspect for this disagreement between the mass estimates for the Virgo cluster is the contamination of the measurements due to the
presence of systematics associated with such over-simplified assumptions as the spherical NFW density profile \citep{nfw97} in the lensing mass
estimates, the hydrostatic equilibrium condition for the X-ray temperature measurement, and etc.
Another more speculative explanation for the difference between the lensing and the dynamic mass measurements for the Virgo cluster
is the failure of GR. We have shown that if the dynamic mass of the Virgo cluster is assumed to be identical to the lensing mass, then the
reconstructed radial velocity profile from observational data appears to less rapidly decrease with the distance than the universal radial
profile from the simulation for the GR+$\Lambda$CDM cosmology. The less rapid decrease of the radial velocity profile with distance
indicates higher peculiar velocities of the filament galaxies at large distances than predicted for the case of GR.
In fact, several numerical works already explored how much the peculiar velocities of galaxies around massive clusters would be enhanced
in the presence of modified gravity (MG) \citep{lam-etal12,corbett-etal14,gronke-etal14b,zu-etal14}.
Before exploring this speculative idea, however, it will be necessary to study comprehensively how the systematics produced by the
over-simplified assumptions mentioned in the above would affect the reconstruction of the radial velocity profile of the filament galaxies
in the vicinity of the clusters. Our future work is in this direction.
\acknowledgments
JL thanks D.Nelson for very helpful discussion.
This work was supported by the research grant from the National Research Foundation of Korea to the Center for
Galaxy Evolution Research (NO. 2010-0027910). JL also acknowledges the financial support by the Basic Science Research
Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (NO. 2013004372).
S.C.R acknowledges the support by the Basic Science Research Program through the National Research Foundation of
Korea (NRF) funded by the Ministry of Education, Science, and Technology (NRF-2012R1A1B4003097).
S.K. acknowledges support from the National Junior Research Fellowship of NRF (No. 2011-0012618).
\clearpage
|
\section{Introduction}
\label{s1}
Granular gases frequently exhibit flows similar to those of normal gases, and for practical purposes these flows are often successfully described by phenomenological hydrodynamic equations \cite{Ha83,Ca90}. The basis for such macroscopic balance equations are in the more fundamental statistical mechanics and kinetic theory descriptions of granular gases. In this context, the idealized model of a granular gas as a monodisperse system of smooth inelastic hard spheres or disks with a constant coefficient of normal restitution has been extensively employed \cite{Du00,Go03}. For this system, hydrodynamic equations to Navier-Stokes order have been derived with expressions for the parameters appearing in them. Starting from the Boltzmann equation for inelastic hard spheres or disks \cite{BDKyS98,ByC01} and also from the revised Enskog theory \cite{GyD99}, the transport coefficients have been evaluated by using an extension of the Chapman-Enskog method. The predictions from the Boltzmann equation have been found to be in good agreement with the values obtained by particle simulation methods in the dilute limit \cite{ByC01}. Using linear response theory, formal Green-Kubo like expressions for the transport coefficients have been derived for low density granular gases \cite{DyB02,ByR04}, and also for arbitrary densities \cite{DByL02,BDyB08}. The latter are not tied to any specific kinetic equation, but their explicit evaluation requires the introduction of some approximations \cite{DByL02}.
In normal fluids, non-equilibrium steady states can be generated by imposing appropriate boundary conditions. Moreover, the control of the boundary conditions permits to keep the gradients of the hydrodynamic fields small, so that the steady state can be studied in the Navier-Stokes domain of the hydrodynamic equations. In granular gases, a new class of steady states shows up. In them, stationarity is reached by an autonomous balance between external constrains and the internal cooling. A typical example is a system under shear flow. There is viscous heating due to the work made on the system at the boundaries. If the system is a granular gas, stationarity is possible when the viscous heating is compensated by the dissipation due to the inelasticity of collisions. The steady state has uniform density and temperature, and a flow velocity with a linear profile. Due to its macroscopic simplicity, it has been extensively studied \cite{LSJyCh84,JyR88,SGyN96,BRyM97,MGSyB99}. This particular steady state exemplifies two features that are characteristic of hydrodynamic steady states of granular gases. First, to compensate the energy dissipation in collisions, the system must develop spatial gradients generating an energy flux, i.e. it must be inhomogeneous. Second, the energy balance leads to a coupling between gradients and inelasticity, so that the limit of small gradients also implies the quasi-elastic limit. Even more, the above coupling is often non-analytical \cite{SGyN96}, implying that the macroscopic description of the steady state can never be brought within the range of validity of the Navier-Stokes hydrodynamics in those cases.
An interesting alternative to the kind of steady states of granular gases described above has being attracting increasing interest in the last years \cite{OyU05,RPGRSCyM11,CMyS12,GSVyP11,PGGSyV12}. A granular gas is confined to a quasi-two-dimensional geometry by placing it between two large parallel plates in the horizontal directions, while the distance between the two plates is smaller than two particle diameters, so the system is actually a monolayer since the particles can not jump on another. The container is vertically vibrated to inject energy through the collisions of the particles with the top and bottom walls. The two-dimensional dynamics of the system when seen from above is considered. It has been observed that it corresponds to that of a two-dimensional granular fluid. Moreover, the system remains homogeneous under a large range of parameters and it eventually reaches a steady homogeneous state. Very recently \cite{BRyS13}, an idealized model has been proposed trying to describe the horizontal dynamics in the above experiment, assuming that the particles are smooth inelastic hard spheres. Then, the projections of the particles on the horizontal plane are described as inelastic hard disks, whose collision rule is modified in order to incorporate a mechanism to transfer the energy injected vertically to the horizontal degrees of freedom. In this sense, it can be classified as a collisional model. The new collision rule has a constant coefficient of normal restitution $\alpha$, and contains a characteristic velocity $\Delta$ that is added to each particle in a collision so that the normal component of the relative velocity is increased by $2 \Delta$ in the collision, independently of the effect of the coefficient of normal restitution. The methods of non-equilibrium statistical mechanics and kinetic theory developed for inelastic hard spheres and disks \cite{BDyS97,vNEyB98} have been applied to the model \cite{BGMyB13}. In this way, the Boltzmann equation and the revised Enskog theory have been formulated. Moreover, the existence of homogeneous hydrodynamics has been analyzed \cite{BMGyB14}. There is a time regime over which the granular temperature of a homogeneous system obeys a closed (hydrodynamic) equation. In the long time limit, the temperature tends to its steady value. Moreover, it has been shown that the homogeneous relaxation of the temperature of the system presents nonlinear memory effects \cite{BGMyB14}, which can be considered as reminiscent of the Kovacs memory effect occurring in the relaxation towards equilibrium of molecular fluids \cite{Ko63}.
The aim of this paper is to derive the hydrodynamic equations to Navier-Stokes order for a dilute confined granular gas as described by the collisional model proposed by Brito {\em et al} \cite{BRyS13}. The starting point will be the Boltzmann kinetic equation and the method to be used a generalization of the Chapman-Enskog procedure. The derivation is based on a special ``normal'' solution of the kinetic equation expanded to low order in the gradients of the hydrodynamic fields. The zeroth order approximation is not a local version, both in space and time, of the distribution function of the homogeneous steady state, but it is based on the distribution describing the homogeneous hydrodynamics. This is an important conceptual and practical difference with the standard application of the Chapman-Enskog method to molecular systems. Of course, it is also possible to consider states close to a stationary one and carry out, for instance, linear response theory around that state to compute transport properties associated to that particular state. The ranges of applicability of the results obtained by both methods are clearly different, although there can be a common limit for the simultaneous validity of both. In particular, the Navier-Stokes shear viscosity of a confined dilute granular gas described by the collisional model in a stationary and uniform Couette flow has been computed by employing linear response theory \cite{SRyB15}. The results obtained here for the shear viscosity will be related with those reported in Ref. \cite{SRyB15}.
The remaining of this paper is organized as follows. In the next section, the Boltzmann equation for the model is given, and the exact balance equations for mass, momentum, and energy are derived from it. The Chapman-Enskog method for obtaining a ``normal'' solution of the kinetic equation as an expansion in the gradients of the hydrodynamic fields is described. Results through Navier-Stokes order for the pressure tensor and the heat flux are given. The associated transport coefficients are shown to obey a complete set of first order differential equations. Some details of the calculations are given in Appendices \ref{ap1} and \ref{ap2}, while the explicit results for the transport coefficients are presented in Sec. \ref{s3}. The theory is not restricted to any range of values of the coefficient of normal restitution nor of the characteristic velocity of the model $\Delta$. The contributions to the transport equations coming from the energy sink term due to the non conservation of kinetic energy in collisions are also discussed. It is shown that there is a first order in the gradients contribution, Euler term, which is proportional to the divergence of the velocity field. The associated transport coefficient is explicitly evaluated.
Appendix \ref{ap3} provides an sketch of the calculation of this coefficient. The Euler term does not exist in molecular systems and it is peculiar of inelastic collisions \cite{DByB08,GMyT13,GCyV13}, although it vanishes in the low density limit of granular gases composed of smooth inelastic hard spheres or disks \cite{BDKyS98}. The expressions of the transport coefficients are particularized for the steady homogeneous state in Sec. \ref{s3}. The peculiarity of this state, in which the temperature is a function of the parameters defining the system \cite{BRyS13,BGMyB13}, leads to much simpler expressions of the coefficients. Further comments and conclusions, as well as the relationship with some previous work for the viscosity transport coefficient, are given in Sec.\ \ref{s4}.
\section{Chapman Enskog solution of the Boltzmann equation}
\label{s2}
The Boltzmann equation obeyed by the one-particle distribution function, $f({\bm r},{\bm v},t)$, of the model reads \cite{BGMyB13}
\begin{equation}
\label{2.1}
\left( \frac{\partial}{\partial t} + {\bm v} \cdot \frac{\partial }{\partial {\bm r}} \right) f({\bm r},{\bm v},t) =
\int d{\bm v}_{1}\, \overline{T}_{0} ({\bm v},{\bm v}_{1}) f({\bm r},{\bm v},t) f({\bm r},{\bm v}_{1},t).
\end{equation}
Here, $\overline{T}_{0}$ is the binary collision operator defined by
\begin{equation}
\label{2.2}
\overline{T}_{0} ({\bm v}_{1}, {\bm v}_{2}) \equiv \sigma^{d-1} \int d \widehat{\bm \sigma}\, \left[ \Theta \left( {\bm v}_{12} \cdot \widehat{\bm \sigma} - 2 \Delta \right) \left( {\bm v}_{12} \cdot \widehat{\bm \sigma} - 2 \Delta \right) \alpha^{-2} b_{\bm \sigma}^{-1}(1,2)- \Theta ({\bm v}_{12} \cdot \widehat{\bm \sigma}) ({\bm v}_{12} \cdot \widehat{\bm \sigma})
\right],
\end{equation}
where $\sigma$ is the diameter of the particles, $d$ the dimension of the system \cite{d}, $\widehat{\bm \sigma}$ is the unit vector joining the center of the two particles at contact, ${\bm v}_{12} \equiv {\bm v}_{1}-{\bm v}_{2}$ is the relative velocity, $\Theta (x)$ is the Heaviside step function, $\alpha$ is the coefficient of normal restitution defined in the interval $0 < \alpha \leq 1$, $\Delta$ is some positive characteristic speed, and $b_{\bm \sigma}^{-1}(i,j)$ is an operator changing all the velocities ${\bm v}_{i}$ and ${\bm v}_{j}$ to its right into the precollisional values corresponding to a collision between them defined by $\widehat{\bm \sigma}$, i.e.,
\begin{equation}
\label{2.3}
b_{\bm \sigma}^{-1}(i,j){\bm v}_{i} = {\bm v}^{*}_{i} = {\bm v}_{i} - \frac{1+ \alpha}{2 \alpha}\, {\bm v}_{ij} \cdot \widehat{\bm \sigma} \widehat{\bm \sigma} + \frac{\Delta \widehat{\bm \sigma}}{\alpha},
\end{equation}
\begin{equation}
\label{2.4}
b_{\bm \sigma}^{-1}(i,j){\bm v}_{j} ={\bm v}^{*}_{j} = {\bm v}_{j} + \frac{1+ \alpha}{2 \alpha}\, {\bm v}_{ij} \cdot \widehat{\bm \sigma} \widehat{\bm \sigma} - \frac{\Delta \widehat{\bm \sigma}}{\alpha}.
\end{equation}
For arbitrary velocity functions, $a({\bm v}_{i},{\bm v}_{j})$ and $b({\bm v}_{i},{\bm v}_{j})$, it is \cite{BGMyB13}
\begin{equation}
\label{2.5}
\int d{\bm v}_{i} \int d {\bm v}_{j}\, b({\bm v}_{i},{\bm v}_{j}) \overline{T}_{0}({\bm v}_{i}, {\bm v}_{j}) a({\bm v}_{i}, {\bm v}_{j})=\int d{\bm v}_{i} \int d {\bm v}_{j}\, a({\bm v}_{i},{\bm v}_{j}) T_{0}({\bm v}_{i}, {\bm v}_{j}) b({\bm v}_{i}, {\bm v}_{j}),
\end{equation}
where
\begin{equation}
\label{2.6}
T_{0}({\bm v}_{i},{\bm v}_{j}) \equiv \sigma^{d-1} \int d \widehat{\bm \sigma}\, \Theta (-{\bm v}_{ij} \cdot \widehat{\bm \sigma} ) |{\bm v}_{ij} \cdot \widehat{\bm \sigma}| \left[ b_{\bm \sigma} (i,j)-1 \right].
\end{equation}
The operator $b_{\bm \sigma} (i,j)$ is the inverse of $b_{\bm \sigma}^{-1} (i,j)$, i.e. it changes ${\bm v}_{i}$ and ${\bm v}_{j}$ into their postcollisional values, ${\bm v}^{\prime}_{i}$ and ${\bm v}^{\prime}_{j}$, given by
\begin{equation}
\label{2.7}
b_{\bm \sigma}(i,j){\bm v}_{i}= {\bm v}^{\prime}_{i} = {\bm v}_{i}- \frac{1+ \alpha}{2} {\bm v}_{ij} \cdot \widehat{\bm \sigma} \widehat{\bm \sigma} + \Delta \widehat{\bm \sigma},
\end{equation}
\begin{equation}
\label{2.8}
b_{\bm \sigma}(i,j){\bm v}_{j} = {\bm v}^{\prime}_{j} = {\bm v}_{j}+ \frac{1+ \alpha}{2} {\bm v}_{ij} \cdot \widehat{\bm \sigma} \widehat{\bm \sigma} - \Delta \widehat{\bm \sigma}.
\end{equation}
The kinetic energy change in a collision is
\begin{equation}
\label{2.9}
e^{\prime}_{ij}-e_{ij} = m\left[ \Delta^{2} - \alpha \Delta {\bm v}_{ij} \cdot \widehat{\bm \sigma} - \frac{1-\alpha^{2}}{4}\, ({\bm v}_{ij} \cdot \widehat{\bm \sigma} )^{2} \right],
\end{equation}
with $m$ being the mass of a particle. Using the identity (\ref{2.5}) it is easily found that
\begin{equation}
\label{2.10}
\int d{\bm v} \int d {\bm v}_{1}\, \overline{T}_{0}({\bm v},{\bm v}_{1} ) f({\bm r},{\bm v},t)f({\bm r},{\bm v}_{1},t)=0,
\end{equation}
\begin{equation}
\label{2.11}
\int d{\bm v} \int d {\bm v}_{1}\, {\bm v} \overline{T}_{0}({\bm v},{\bm v}_{1} ) f({\bm r},{\bm v},t)f({\bm r},{\bm v}_{1},t)=0,
\end{equation}
reflecting the conservation of the number of particles and the momentum, respectively. On the other hand, it is
\begin{equation}
\label{2.12}
\int d{\bm v} \int d {\bm v}_{1}\, \frac{m v^{2}}{2} \overline{T}_{0}({\bm v},{\bm v}_{1} ) f({\bm r},{\bm v},t)f({\bm r},{\bm v}_{1},t)= \omega[f,f].
\end{equation}
The term $\omega[f,f]$ provides the rate of energy change due to the inelasticity of collisions, and the functional $\omega[f,h]$ is
\begin{eqnarray}
\label{2.13}
\omega[f,h] & \equiv & \frac{\pi^{(d-1)/2} m \sigma^{d-1}}{2} \int d {\bm v}_{1} \int d {\bm v}_{2}\, f({\bm r},{\bm v}_{1},t) h ({\bm r}, {\bm v}_{2},t) \nonumber \\
&& \times \left[ \frac{ \Delta^{2} v_{12}}{\Gamma \left( \frac{d+1}{2} \right)}+ \frac{\pi^{1/2} \alpha \Delta v_{12}^{2}}{2 \Gamma \left( \frac{d+2}{2} \right)}- \frac{(1- \alpha^{2})v_{12}^{3}}{4 \Gamma \left( \frac{d+3}{2} \right)} \right].
\end{eqnarray}
The macroscopic number of particles density, $n({\bm r},t)$, flow velocity, ${\bm u}({\bm r},t)$, and granular temperature, $T({\bm r},t)$, are defined from the one-particle distribution function in the usual way,
\begin{equation}
\label{2.14}
n({\bm r},t) \equiv \int d{\bm v}\, f({\bm r},{\bm v},t),
\end{equation}
\begin{equation}
\label{2.15}
n({\bm r},t) {\bm u}({\bm r},t) \equiv \int d{\bm v}\,{\bm v} f({\bm r},{\bm v},t),
\end{equation}
\begin{equation}
\label{2.16}
\frac{d}{2} n({\bm r},t) T({\bm r},t) \equiv \int d{\bm v}\,\frac{m V^{2}}{2}\, f({\bm r},{\bm v},t),
\end{equation}
where ${\bm V} ({\bm r},t)= {\bm v}- {\bm u} ({\bm r},t) $ is the velocity of the particle relative to the flow field. Balance equations for the above fields follow by taking velocity moments in the Boltzmann equation, Eq. (\ref{2.1}),
\begin{equation}
\label{2.17}
\frac{\partial n}{\partial t} + {\bm \nabla} \cdot \left( n {\bm u} \right) =0,
\end{equation}
\begin{equation}
\label{2.18}
\frac{\partial {\bm u}}{\partial t} +{\bm u} \cdot {\bm \nabla} {\bm u} + (mn)^{-1} {\bm \nabla}
\cdot \sf{P}=0,
\end{equation}
\begin{equation}
\label{2.19}
\frac{\partial T}{\partial t} +{\bm u} \cdot {\bm \nabla}T + \frac{2}{nd} \left( {\sf P} : {\bm \nabla} {\bm u} + {\bm \nabla} \cdot {\bm J}_{q} \right) = - \zeta T.
\end{equation}
In the above equations, the pressure tensor, ${\sf P}$, and the heat flux, ${\bm J}_{q}$, are defined by
\begin{equation}
\label{2.20}
{\sf P}({\bm r},t) \equiv m \int d{\bm v}\, {\bm V}({\bm r},t) {\bm V}({\bm r},t) f ({\bm r},{\bm v},t)
\end{equation}
and
\begin{equation}
\label{2.21}
{\bm J}_{q}({\bm r},t) \equiv \frac{m}{2} \int d{\bm v}\, V^{2}({\bm r},t) {\bm V}({\bm r},t) f({\bm r},{\bm v},t),
\end{equation}
respectively. In addition, Eq. (\ref{2.19}) contains the rate of change of the temperature, $\zeta({\bm r},t)$, due to the inelasticity of collisions, whose expression is
\begin{equation}
\label{2.22}
\zeta ({\bm r},t) \equiv -\frac{2}{n({\bm r},t) T({\bm r},t) d}\, \omega[f,f].
\end{equation}
The minus sign has been introduced by analogy with a system of smooth inelastic hard spheres or disks, but in the present context it does not presuppose that $\zeta$ is (semi)defined positive \cite{Ha83}.
To close the balance equations (\ref{2.17})-(\ref{2.19}), it is necessary to express the fluxes and the temperature change rate in terms of the macroscopic fields, by means of some constitutive relations. To accomplish this, the Chapman-Enskog theory \cite{McL89} assumes the existence of a normal solution of the Boltzmann equation, i.e. a solution in which all the dependence of the distribution function on position and time occurs through its functional dependence on the macroscopic fields $n$, ${\bm u}$, and $T$,
\begin{equation}
\label{2.23}
f({\bm r},{\bm v},t) = f[{\bm v}|n, {\bm u}, T].
\end{equation}
Next, it is assumed that the space and time variations of the fields are small, so that the functional dependence of the distribution function on the fields can be localized in space and time by means of an expansion in gradients. Then, the distribution function is expressed as a power series expansion in a formal uniformity parameter $\epsilon$,
\begin{equation}
\label{2.24}
f=f^{(0)}+ \epsilon f^{(1)}+ \epsilon^{2} f^{(2)}+ \cdots.
\end{equation}
Since the aim is to generate a gradient expansion, a factor of $\epsilon$ is assigned to every gradient operator. Moreover, the Chapman-Enskog method uses the multiple-scale perturbation theory \cite{Sch91}. In practice, this is done by using the expansion in Eq. (\ref{2.24}) into the definition of the fluxes and the dissipation rate $\zeta$. Then the resulting expansions are introduced into the macroscopic balance equations to get an identification of the time derivatives of the macroscopic fields in the form of an expansion in the uniformity parameter,
\begin{equation}
\label{2.25}
\frac{\partial}{\partial t}= \partial_{t}^{(0)} + \epsilon \partial_{t}^{(1)} + \epsilon^{2} \partial_{t}^{(2)} + \cdots.
\end{equation}
Details of the application of the method are given in Appendices \ref{ap1}, \ref{ap2}, and \ref{ap3}. To first order in the gradients, the pressure tensor and heat flux are given by
\begin{equation}
\label{2.26}
{\sf P} = n T {\sf I} - \eta \left[ {\bm \nabla} {\bm u} + ( {\bm \nabla} {\bm u})^{+} - \frac{2}{d} {\bm \nabla} \cdot {\bm u} {\sf I} \right],
\end{equation}
\begin{equation}
\label{2.27}
{\bm J}_{q}= - \kappa {\bm \nabla}T- \mu {\bm \nabla }n,
\end{equation}
where ${\sf I}$ is the unit tensor in $d$ dimensions, $({\bm \nabla} {\bm u})^{+}$ is the transposed of ${\bm \nabla} {\bm u}$, $\eta$ is the coefficient of shear viscosity, $\kappa$ the heat conductivity, and $\mu$ a new coefficient coupling the heat flux and the density gradient, which is peculiar of inelastic collisions \cite{BMyD96}. To distinguish between the two energy transport coefficients, sometimes $\kappa$ is referred to as the thermal heat conductivity and $\mu$ as the diffusive heat conductivity. The transport coefficients are determined by the normal solutions of the first order differential equations
\begin{equation}
\label{2.28}
\frac{\overline{\zeta}^{(0)} \Delta^{*}}{2} \frac{\partial \overline{\eta}}{\partial \Delta^{*}} + \left(\overline{\nu}_{\eta} - \frac{\overline{\zeta}^{(0)}}{2} \right) \overline{\eta}= \frac{2^{5/2} \pi^{\frac{d-1}{2}}}{(d+2) \Gamma \left( d/2 \right)}\, ,
\end{equation}
\begin{equation}
\label{2.29}
\frac{\overline{\zeta}^{(0)} \Delta^{*}}{2} \frac{\partial \overline{\kappa}}{\partial \Delta^{*}} + \left( \overline{\nu}_{\kappa}+ \frac{\Delta^{*}}{2} \frac{\partial \overline{\zeta}^{(0)}}{\partial \Delta^{*}} - 2 \overline{\zeta}^{(0)}\right) \overline{\kappa} = \frac{2^{5/2} (d-1) \pi^{\frac{d-1}{2}} }{d(d+2) \Gamma \left( d/2 \right)}\, \left( 1+2a_{2}- \frac{\Delta^{*}}{2} \frac{\partial a_{2}}{\partial \Delta^{*}}\ \right),
\end{equation}
\begin{equation}
\label{2.30}
\frac{\overline{\zeta}^{(0)} \Delta^{*}}{2} \frac{\partial \overline \mu}{\partial \Delta^{*}} + \left( \overline{\nu}_{\mu} - \frac{3 \overline\zeta^{(0)}}{2} \right) \overline{\mu}
- \overline{\zeta}^{(0)} \overline{\kappa} = \frac{2^{5/2} (d-1) \pi^{\frac{d-1}{2}} a_{2}}{d(d+2) \Gamma \left( d/2 \right)}\, .
\end{equation}
In these equations, $\Delta^{*} \equiv \Delta \left( m/2T \right)^{1/2}$ and dimensionless transport coefficients have been introduced. They are defined by
\begin{equation}
\label{2.31}
\overline{\eta} \equiv \frac{\eta}{\eta_{0}}, \quad \overline{\kappa} \equiv \frac{\kappa}{\kappa_{0}}
, \quad \overline{\mu} \equiv \frac{n \mu }{T \kappa_{0}},
\end{equation}
where
\begin{equation}
\label{2.32}
\eta_{0}= \frac{2+d}{8} \Gamma \left( d/2 \right) \pi^{-\frac{d-1}{2}} \left(mT \right)^{1/2} \sigma^{-(d-1)},
\end{equation}
and
\begin{equation}
\label{2.33}
\kappa_{0}= \frac{d(d+2)^{2}}{16(d-1)}\, \Gamma \left( d/2 \right) \pi^{-\frac{d-1}{2}} \left( \frac{T}{m} \right)^{1/2} \sigma^{-(d-1)}
\end{equation}
are the shear viscosity and the (thermal) heat conductivity, respectively, of a molecular gas described by the Boltzmann equation, with the Boltzmann constant set equal to unity. The dimensionless functions introduced in Eqs. (\ref{2.28})-(\ref{2.30}) are
\begin{equation}
\label{2.34}
\overline{\zeta}^{(0)} \equiv \frac{\zeta^{(0)}}{n \sigma^{d-1}} \, \left( \frac{m}{2T} \right)^{1/2}, \quad
\overline{\nu}_{\eta} \equiv \frac{\nu_{\eta}}{n \sigma^{d-1}} \, \left( \frac{m}{2T}\right)^{1/2}, \quad \overline{\nu}_{\kappa} = \overline{\nu}_{\mu}
\equiv \frac{\nu_{\kappa}}{n \sigma^{d-1}} \, \left( \frac{m}{2T}\right)^{1/2}.
\end{equation}
The expression of the zeroth order rate of change of the temperature, $\zeta^{(0)}$, is given in Eq.\ (\ref{ap1.13}), while the frequencies $\nu_{\eta}$ and $\nu_{\kappa}$ are given in Eqs. ({\ref{ap2.5}) and (\ref{ap2.6}), respectively. Some details of the calculations are shown in Appendices \ref{ap1} and {\ref{ap2}. Finally, Eqs. (\ref{2.28})-(\ref{2.30}) have been obtained by considering the distribution function in the called first Sonine approximation, in which the deviation of the one-particle distribution function of the gas from the Gaussian are characterized by the coefficient $a_{2}$, given by the normal solution of the ordinary differential equation (\ref{ap1.14}). Moreover, the same kind of approximation has been considering upon evaluating the hydrodynamic fluxes. This is the usual approximation to obtain explicit expressions for the transport coefficients of a gas with elastic collisions and there is no reason to question its accuracy here as well. Actually, it has been shown to lead to quantitatively right approximations in the case of a system of smooth inelastic hard spheres or disks \cite{ByC01}.
\section{Euler and Navier-Stokes transport coefficients}
\label{s3}
As a consequence of the confinement of the fluid and its description by means of a modified hard collision, there is a contribution to the hydrodynamic equations of the rate of change of the temperature $\zeta ({\bm r},t)$ of first order in the gradients, namely it is (see Eq.\ (\ref{ap1.32}))
\begin{equation}
\label{3.1}
\zeta^{(1)} ({\bm r},t) = \zeta_{1} {\bm \nabla} \cdot {\bm u}.
\end{equation}
The Euler transport coefficient $\zeta_{1}$ represents dissipation due to the inelastic character of collisions proportional to ${\bm \nabla} \cdot {\bm u}$. It has no analogue for elastic fluids, where the Euler hydrodynamics (first order in the gradients of the fields) is referred to as the ``perfect fluid'' equations, since there is not dissipation in that limit. The expression derived here vanishes in the limit $\Delta^{*} \rightarrow 0$, as a consequence of symmetry considerations \cite{BDKyS98}, i.e. there is no dissipation to Euler order in a dilute gas of smooth inelastic hard spheres or disks. On the other hand, a term of the form given in Eq. (\ref{3.1}) is present in the hydrodynamic equations even in systems of smooth particles if density effects are considered \cite{GyD99,DByB08,BDyB08}. The calculation of $\zeta_{1}$ using the Chapman-Enskog procedure, requires to determine $f^{(1)}$. The details of the calculation are given in Appendix \ref{ap3}, where the first Sonine polynomial expansion is again employed. The same approximation was used above to compute the Navier-Stokes transport coefficients. In Fig.\ \ref{fig1}, the dimensionless coefficient $ \zeta_{1}$
is plotted as a function of the speed parameter for two values of the coefficient of normal restitution, namely $\alpha=0.8$ and $\alpha=0.9$. The value of the transport coefficient in the homogenous steady state corresponding to each value of $\alpha$ is indicate in the figure. To find the values of the transport coefficient a coupled of first order differential equations have to be solved, namely Eqs. (\ref{ap1.14}) and (\ref{ap3.3}). The curves reported in the figure correspond to the hydrodynamic solutions of the equations, i.e. they are identified independently of the initial conditions. The method is described below in detail for the shear viscosity coefficient and it will be not be discussed now.
The coefficient $\zeta_{1} $ is the contribution of the energy source in collisions to what would physically constitute the effects of the hydrostatic pressure at the Euler order. If a small element of the confined granular gas is considered, then the pressure that the fluid element exerts on its boundaries is decreased or increased by the energy lost locally in collisions. At the level of linear hydrodynamics, the pressure and the Euler dissipative term are indistinguishable in their physical implications \cite{BDyB08}.
\begin{figure}
\includegraphics[scale=0.6,angle=0]{ByB15f1.eps}
\caption{(Color online) Dimensionless Euler transport coefficient $\zeta_{1}$ as a function of dimensionless characteristic speed $\Delta^{*}$ in a two-dimensional system. The (red) dashed line is for a coefficient of normal restitution $\alpha=0.8$ and the (black) solid line is for $\alpha=0.9$. The (blue) dots indicate the values of the transport coefficient in each of the two steady states.}
\label{fig1}
\end{figure}
To compute the Navier-Stokes transport coefficients, Eqs.\ (\ref{2.28})-(\ref{2.30}) have to be solved. In the equations, the hydrodynamic expression of the second Sonine coefficient, $a_{2}$, has to be used. The latter is obtained by numerically solving Eq. (\ref{ap1.14}) as a function of $\Delta^{*}$ for a fixed value of $\alpha$, and a given initial condition $a_{2}(\alpha, \Delta^{*}_{0})=a_{2,0}$. It is seen that all the trajectories converge quite fast towards a universal curve, identified as the hydrodynamic expression of the second Sonine coefficient \cite{BMGyB14}. A similar method has been employed here to generate the hydrodynamic transport coefficients. Actually, what has been done is to simultaneously solve the equation for $a_{2}$ and those for the transport coefficients. Since the rate of variation of the temperature vanishes in the steady state, the equations for the transport coefficients have a singularity at the steady value $\Delta^{*}= \Delta^{*}_{st}$. Therefore, in the numerical simulations, trajectories have been generated starting from both $\Delta^{*} _{0}> \Delta^{*}_{st}$ and with $\Delta^{*}_{0} < \Delta^{*}_{st}$. The hydrodynamic solution giving the expression of the transport coefficient is the common part of all the numerical solutions. As an example, the numerical results obtained for the dimensionless coefficient $\overline{\eta}$ in a two dimensional system with $\alpha=0.9$ are shown in Fig.\ \ref{fig2}. All the numerical trajectories converge towards the same curve, then forgetting the initial conditions used to generate them. This is consistent with the existence of a hydrodynamic shear viscosity being a function of only the local hydrodynamic fields, but not of the previous history or some initial values. In the particular case shown in Fig.\ \ref{fig2}, several initial values of the viscosity parameter corresponding to $\Delta^{*}= 0.005$ and $\Delta^{*}=10$ have been employed. The curves tend to converge quite fast in the range $0 \leq \Delta^{*} \lesssim 0.2$. For $\Delta^{*} \gtrsim 0.2$, the dependence of the solution of the differential equation on the initial value of $\overline{\eta}$ used for $\Delta^{*}=\Delta^{*}_{0}$ is rather strong and much more intensive numerical simulations would be needed to identify the value of the hydrodynamic shear viscosity. The two particular solutions drawn in Fig.\ \ref{fig2} for $\Delta^{*}>\Delta^{*}_{st}$ correspond to $\overline{\eta}(\Delta^{*}=10)=100 $ and $\overline{\eta}(\Delta^{*}=10)=0$, respectively, while the third plotted particular solution has been obtained with the initial condition $\overline{\eta}(\Delta^{*}=0.005)= 10$. Results obtained with other initial conditions can not be distinguished from the normal curve on the scale of the figure.
\begin{figure}
\includegraphics[scale=0.6,angle=0]{ByB15f2.eps}
\caption{(Color online) Adimensionalized quantity $\overline{\eta}$ as a function of the dimensionless characteristic speed $\Delta^{*}$ in a two-dimensional system with $\alpha=0.9$. The (red) dashed lines correspond to numerical solutions of Eq.\ (\ref{2.28}) obtained by using different initial conditions, i.e. different values for the pair $\Delta^{*}_{0}, \overline{\eta}(\Delta^{*}_{0})$. The (black) solid line is the universal curve to which all the solutions converge. This is precisely the function identified as the dimensionless hydrodynamic shear viscosity. Also indicated in the figure is the steady value of $\Delta^{*}$, denoted by $\Delta^{*}_{st}$ and the shear viscosity of the steady state, $\overline{\eta}_{st}$ .}
\label{fig2}
\end{figure}
In Figs.\ \ref{fig3} - \ref{fig5} the coefficients of shear viscosity, $\overline{\eta}$, (thermal) heat conductivity, $\overline{\kappa}$, and diffusive heat conductivity, $\overline{\mu}$, are plotted as a function of the dimensionless characteristic speed for a two-dimensional system. Two values of the coefficient of normal restitution have been considered, namely $\alpha=0.8$ and $\alpha=0.9$. The reported curves correspond to the hydrodynamic transport coefficients and have been obtained by the same method as described above for the shear viscosity. The values of the several transport coefficients in the steady state are indicated.
\begin{figure}
\includegraphics[scale=0.6,angle=0]{ByB15f3.eps}
\caption{(Color online) Adimensionalized shear viscosity $\overline{\eta}$ of a two-dimensional system as a function of the dimensionless speed parameter $\Delta^{*}$. The (red) dashed curve is for $\alpha=0.8$ and the (black) solid line is for $\alpha=0.9$. The (blue) dots indicate the steady state values in each case.}
\label{fig3}
\end{figure}
\begin{figure}
\includegraphics[scale=0.6,angle=0]{ByB15f4.eps}
\caption{(Color online) Adimensionalized (thermal) heat conductivity $\overline{\kappa}$ of a two-dimensional system as a function of the dimensionless speed parameter $\Delta^{*}$. The (red) dashed curve is for $\alpha=0.8$ and the (black) solid line is for $\alpha=0.9$. The (blue) dots indicate the steady state values in each case. }
\label{fig4}
\end{figure}
\begin{figure}
\includegraphics[scale=0.6,angle=0]{ByB15f5.eps}
\caption{(Color online) Adimensionalized diffusive heat conductivity $\overline{\mu}$ as a function of the dimensionless speed parameter $\Delta^{*}$ for a two dimensional system. The (red) dashed curve is for $\alpha=0.8$ and the (black) solid line is for $\alpha=0.9$. The (blue) dots indicate the steady state values in each case.}
\label{fig5}
\end{figure}
It is observed that the three Navier-Stokes transport coefficients are monotonically decreasing functions of the speed parameter $\Delta^{*}$ for the two values of the restitution coefficient $\alpha$ considered in the figures. A similar behavior was found for other values of $\alpha$. The coefficient of diffusive heat conductivity $\mu$, becomes even negative for large enough values of $\Delta^{*}$. Notice that this does not seem to imply the violation of any fundamental physical law or be incompatible with any physical symmetry. Nevertheless, it is quite possible that the exact value of $\Delta^{*}$ at which the change in sign of $\mu$ occurs be a consequence of the approximations made and, in particular, of the first Sonine approximation. When the prediction for $\mu$ in this approximation is rather small, it is evident that higher order corrections might become relevant.
\section{Transport coefficients in the homogeneous steady state}
\label{s4}
A particularly relevant state of the confined granular gas is the homogeneous steady state. Stationarity of the temperature implies that the rate of change of the temperature vanishes, i.e. it is
\begin{equation}
\label{4.1}
\overline{\zeta}^{(0)}(\Delta^{*}_{st})=0.
\end{equation}
Then, particularization of Eq.\ (\ref{ap3.3}) for the steady state yields
\begin{eqnarray}
\label{4.1a}
\zeta_{1,st} &= & 2 \Delta^{*}_{st} \left( \frac{\partial a_{2}}{\partial \Delta^{*}} \right)_{\Delta^{*}= \Delta^{*}_{st}}\ \overline{\zeta}_{1}(\Delta^{*}_{st}) \nonumber \\&& \times \left\{ \frac{8 \chi(\Delta^{*}_{st})}{d(d+2)} + d \overline{\zeta}_{1}(\Delta^{*}_{st}) \left[ 4(a_{2,st}+1)- \Delta^{*}_{st} \left( \frac{\partial a_{2}}{\partial \Delta^{*}} \right)_{\Delta^{*} = \Delta^{*}_{st}}\ \right] \right\}^{-1}.
\end{eqnarray}
Similarly, particularization of Eqs. (\ref{2.28})-(\ref{2.30}) for the steady state lead to explicit expressions for the Navier-Stokes transport coefficients for this state,
\begin{equation}
\label{4.2}
\overline{\eta}_{st}= \frac{2^{5/2} \pi^{\frac{d-1}{2}}}{(d+2) \Gamma \left( d/2 \right)}\, \overline{\nu}_{\eta}^{-1}(\Delta^{*}_{st},a_{2,st}),
\end{equation}
\begin{eqnarray}
\label{4.3}
\overline{\kappa}_{st} & = & \frac{2^{5/2} (d-1) \pi^{\frac{d-1}{2}} }{d(d+2) \Gamma \left( d/2 \right)}\, \left[ 1+2a_{2,st}- \frac{\Delta^{*}_{st}}{2}
\left( \frac{\partial a_{2}}{\partial \Delta^{*}} \right)_{\Delta^{*}= \Delta^{*}_{st}}\ \right] \nonumber \\
&& \times \left[ \overline{\nu}_{\kappa} (\Delta^{*}_{st},a_{2,st})+ \frac{\Delta^{*}_{st}}{2} \left( \frac{\partial \overline{\zeta}^{(0)}}{\partial \Delta^{*}} \right)_{\Delta^{*}= \Delta^{*}_{st}} \right]^{-1},
\end{eqnarray}
\begin{equation}
\label{4.4}
\overline{\mu}_{st}= \frac{2^{5/2} (d-1) \pi^{\frac{d-1}{2}} a_{2,st}}{d(d+2) \Gamma \left( d/2 \right) \overline{\nu}_{\mu}( \Delta^{*}_{st},a_{2,st})}\, .
\end{equation}
The frequencies $\overline{\nu}_{\eta}$ and $\overline{\nu}_{\kappa}= \overline{\nu}_{\mu}$ are defined in Eqs.\ (\ref{2.34}). Moreover, the calculation of $a_{2}(\Delta^{*})$ for $\Delta^{*}$ in the vicinity of $\Delta^{*}_{st}$ can be carried out in an efficient and quite accurate way by noting that near $\Delta^{*}_{st}$ it is $|\partial a_{2} / \partial \Delta^{*} | \ll 1$ (see, for instance, Fig. 9 in Ref. \cite{BGMyB14}). Then, near the steady state, Eq. (\ref{ap1.14}) yields
\begin{equation}
\label{4.5}
a_{2} \approx - \frac{B_{0}+4A_{0}}{B_{1}+4(A_{0}+A_{1})}.
\end{equation}
The expressions of $A_{0}$, $A_{1}$, $B_{0}$, and $B_{1}$ are given in Eqs. (\ref{ap1.15})-(\ref{ap1.18}).
It is now a simple task to evaluate the Euler and Navier-Stokes transport coefficients in the steady state. They are plotted in Figs. \ref{fig6}-\ref{fig9} as a function of the coefficient of normal restitution $\alpha$. The four coefficients present a clear dependence with the inelasticity. The Euler transport coefficient is a monotonic decreasing function of the coefficient of normal restitution, while the dimensionless shear viscosity monotonically increases with the value of $\alpha$. This latter behavior is consistent with molecular dynamics simulation results reported both for dilute \cite{SRyB15} and moderately dense systems
\cite{BRyS13}. Moreover, the dependence of the viscosity on the coefficient of normal restitution is clearly nonlinear, again in agreement with the simulations for dilute systems. It is worth to remind that in (non-confined) dilute granular gases of smooth inelastic hard spheres \cite{BDKyS98,ByC01} the viscosity decreases as the coefficient of normal restitution increases, and that in a stochastic thermostat model it has been found to be a non-monotonic function of the inelasticity \cite{GMyT13,GCyV13}. On the other hand, the dependence on the restitution coefficients of the two transport coefficients associated with the heat flux is not monotonic in the homogeneous steady state of the model discussed here, exhibiting both a minimum. Moreover, the coefficient of diffusive heat conductivity $\mu_{st}$ is negative in the whole range of values of $\alpha$, while it is always positive in a dilute non-confined gas of inelastic hard spheres or disks. Notice that the dependence of the steady transport coefficients on the inelasticity of the system is quite strong. In particular, the (thermal) heat conductivity for $\alpha=0.8$ is about $25\%$ smaller than its elastic limit value.
In any case, when interpreting the results in Figs. \ref{fig6}-\ref{fig9}, it must be kept in mind that the bare transport coefficients have been scaled with a function of the temperature of the steady state, and that this temperature is in turn a function of the coefficient of normal restitution (and the velocity parameter of the model $\Delta$). As a consequence, it is not possible to deduce the general expressions of the transport coefficients of the model to Navier-Stokes order from their form in the steady homogeneous state.
\begin{figure}
\includegraphics[scale=0.6,angle=0]{byb15f6.eps}
\caption{Dimensionless Euler transport coefficient of the two-dimensional confined granular gas in the homogeneous steady state $\zeta_{1,st}$, as a function of the coefficient of normal restitution $\alpha$.}
\label{fig6}
\end{figure}
\begin{figure}
\includegraphics[scale=0.6,angle=0]{ByB15f7.eps}
\caption{Dimensionless shear viscosity of the two-dimensional confined granular gas in the homogeneous steady state $\overline{\eta}_{1,st}$, as a function of the coefficient of normal restitution $\alpha$.}
\label{fig7}
\end{figure}
\begin{figure}
\includegraphics[scale=0.6,angle=0]{ByB15f8.eps}
\caption{Dimensionless (thermal) heat conductivity of the two-dimensional confined gas in the homogeneous steady state $\overline{\kappa}_{st}$, as a function of the coefficient of normal restitution $\alpha$.}
\label{fig8}
\end{figure}
\begin{figure}
\includegraphics[scale=0.6,angle=0]{ByB15f9.eps}
\caption{ Dimensionless diffusive heat conductivity of the two-dimensional confined granular gas in the homogeneous steady state $\overline{\mu}_{st}$, as a function of the coefficient of normal restitution $\alpha$.}
\label{fig9}
\end{figure}
\section{Discussion and conclusions}
\label{s5}
The objective of this work has been to derive the hydrodynamic equations to Navier-Stokes order for a model of confined granular gas from an underlying kinetic
theory, with all the parameters given explicitly. For clarification and context, the following comment must be taken into account. When applying the Chapman-Enskog procedure, the distribution function has been computed up to first order in the gradients of the hydrodynamic fields density, flow velocity, and granular temperature. Consequently, the heat and momentum fluxes are also determined to the same order. Since they occur as divergences in the balance equations, they lead to terms of second order in the gradients in those equations, what is usually referred to as the Navier-Stokes approximation for the fluxes. Also the rate of change of the temperature $\zeta$ has been computed to first order, but it appears without any gradient operator in front of it in the balance equation for the energy. It follows that consistency of the Navier-Stokes order would require, in principle, computing $\zeta$ up to second order in gradients, i.e. going an order further in the Chapman-Enskog expansion of the distribution function, the Burnett order. Such a calculation is rather involved and lengthy. The only case in which we are aware that second order contributions from $\zeta$ have been analyzed deals with the linear contributions in a low density gas of smooth inelastic hard spheres
\cite{BDKyS98}. There, it is found than the terms are very small as compared with the similar ones arising from the fluxes. A similar behavior is likely to occur here with all the second order in the gradients contributions from $\zeta$.
The derived hydrodynamic equations are general, in the sense of having no restriction with regards to the values of the coefficient of normal restitution $\alpha$ or the velocity parameter of the model $\Delta$. In particular, they hold in principle arbitrarily far from the homogeneous steady state, as long as the system be near an homogeneous hydrodynamic state. In the steady state, both parameters $\alpha$ and $\Delta$ determine the temperature of the system, that is not an arbitrary parameter anymore. Of course, the general hydrodynamic equations can be particularized for the steady state, as it has been actually done here, but it must be emphasized that the general form of the hydrodynamic equations can not be inferred from the equations derived for the steady situation.
The shear viscosity of the model in the steady state has been measured by using event driven molecular dynamics simulations. The transport coefficient was obtained from the decay rate of the transverse current \cite{BRyS13,SRyB15}. In a dense system ($n\sigma^{2}=0.4)$ it was found that a linear fit in $(1-\alpha)$ gives
\begin{equation}
\label{5.1}
\overline{\eta} \simeq 1.0512 \sqrt{\pi} \left[ 1-0.28(1-\alpha) \right].
\end{equation}
An expansion of Eq. (\ref{4.2}) in powers on $\alpha$ to first order gives
\begin{equation}
\label{5.2}
\overline{\eta} \simeq 1- \frac{17}{64} (1-\alpha).
\end{equation}
As expected, the low density theory developed here is not able to predict the prefactor in Eq.\ (\ref{5.1}). In dense systems, the collisional contributions to momentum transfer, and consequently to the shear viscosity play a fundamental role, and those effects are neglected at the level of the Boltzmann equation. However, if the lowest order inelasticity correction is considered, the results obtained in this paper are in good agreement with the simulation results for dense systems. A similar conclusion was reached in Ref.\ \cite{SRyB15}.
A relevant and recurrent question when deriving hydrodynamic equations from kinetic theory is to determine the context in which the equations apply. The small parameter in the Chapman-Enskog expansion is the ratio of the mean free path relative to the wavelength of the variation of the hydrodynamic fields. The mean free path is independent of the time for the homogeneous hydrodynamics. Consequently, it seems sensible to conclude that the conditions for the Navier-Stokes order hydrodynamics of the model are the same as for usual granular gases of inelastic hard spheres or disks and also for elastic collisions, i.e. for sufficiently large space and time scales as compared with the mean free path and the inverse collision frequency.
\section{Acknowledgements}
This research was supported by the Ministerio de Econom\'{\i}a y Competitividad (Spain) through Grant No. FIS2014-53808-P (partially financed by FEDER funds).
|
\section{Introduction}
The CIPM Mutual Recognition Arrangement (MRA) \cite{CIPM-MRA}, which was put into force in October 1999 and meanwhile was signed by 95 National Metrology Institutes (NMIs) and 4 international organisations worldwide describes the foundations and requirements for the mutual recognition of national measurement standards and for calibration and measurement certificates issued by NMIs. In addition the CIPM MRA covers a further 150 institutes designated by the signatory bodies. The MRA, chapter 3 states, that ``\emph{\dots the technical basis of this arrangement is the set of results obtained in the course of time through key comparisons carried out by the Consultative Committees of the CIPM, the BIPM, and the regional metrology organisations (RMOs), and published by the BIPM and maintained in the key comparison database. Key comparisons carried out by Consultative Committees or the BIPM are referred to as CIPM key comparisons; key comparisons carried out by regional metrology organisations are referred to as RMO key comparisons; RMO key comparisons must be linked to the corresponding CIPM key comparisons by means of joint participants. The degree of equivalence derived from an RMO key comparison has the same status as that derived from a CIPM key comparison.}'' To further support mutual confidence in the validity of calibration and measurement certificates issued by participating institutes so-called supplementary comparisons shall be carried out in addition. The metrology community has followed these documented requirements to carry out international comparison measurements \cite{CIPM-MRA-GUIDE} on a regular basis, whose results are continuously registered and made available to the public in the BIPM key comparison database (KCDB) \cite{KCDB}.
The issue of linking by means of joint participants is already addressed in the MRA, however without prescribing a particular mathematical linking procedure. Different approaches for robust statistical linking of the results of comparisons, e.~g. by transferring a key comparison reference value (KCRV) from one comparison to the other, have been discussed and proposed in literature and have been applied for the analysis of comparison results, too \cite{STEELE}, \cite{ELSTER}, \cite{WHITE}, \cite{SUTTON}.
Most of the proposed approaches for robust statistical linking follow the basic assumption of the MRA, namely that the results of an RMO key comparison have to be linked to the results of a CIPM key comparison. The CIPM key comparison in this concept generally is regarded as a primary comparison, both with respect to the chronological order and with respect to the participating laboratories, because ``\emph{\dots participation in a CIPM key comparison is open to laboratories having the highest technical competence and experience, normally the member laboratories of the appropriate Consultative Committee.}'' (see sect. 6.1 of \cite{CIPM-MRA}). In this sense the linking of CIPM key comparisons and RMO key comparisons can be regarded as a hierarchical process. However, in \cite{STEELE} approaches for linking of comparisons through global minimisation by application of generalised least squares methods \cite{NIELSEN} were shortly discussed too, which can be regarded as examples of non-hierarchical linking methods.
In the field of dimensional metrology which is under the responsibility of the CIPM Consultative Committee for Length (CCL), international comparisons were started immediately after the MRA was signed. The results of the first CIPM key comparison on gauge blocks e.~g. were already published in 2002 \cite{THALMANN}.
Based on the experience gained from the international comparisons in the field of dimensional metrology, the CCL working group on dimensional metrology (CCL-WGDM\footnote{In June 2009 the working group structure of CCL has changed: MRA related work is now dealt with by the WG-MRA while linking issues are addressed by the task group on on KC linking (TG-L)}) has published a document which describes the characteristics of dimensional comparisons and which also proposes an alternative comparison scheme, the so-called CCL-RMO key comparison scheme \cite{CCL}. The background and the details of the proposed comparison scheme are not repeated here, however some core statements from this document are cited:
\begin{itemize}
\item The CCL-RMO key comparisons follow the idea of several comparisons mutually linked together, without the necessity of a top level comparison delivering a KCRV.
\item The classical scheme of CCL and RMO key comparisons can be considered to be a special case of the more general CCL-RMO scheme.
\item In terms of linking laboratories within and across regions by calculation of their respective degrees of equivalence, the classical (hierarchical) and the CCL-RMO comparison scheme can be regarded to be equivalent. The linking of the comparisons is guaranteed by common participation of selected laboratories.
\end{itemize}
In \cite{DECKERJ} a generalised formalism was presented for linking a CIPM key comparison with similar regional key comparisons. As an example, this procedure was applied to link the results of a regional (SIM) gauge block comparison \cite{DECKER} to the results of the first CIPM key comparison on gauge blocks \cite{THALMANN}. In the applied procedure, a so-called linking invariant parameter is calculated based on the results of the linking laboratories, i.~e. those laboratories which participated in both comparisons, which in this case used different transfer standards.
This proposed method was discussed and well accepted in the CCL-WGDM, however it was also criticised that it still required to choose one comparison as primary. This condition is not in full alignment of the linking procedure to be applied for the analysis of comparisons according to the proposed CCL-RMO key comparison scheme. It was argued, that an adapted `distributed linking' procedure should be developed which should provide a statistically rigorous linking taking into account all the available information of the comparisons to be linked, however without the necessity to choose one comparison as a primary one.
In this paper a Bayesian approach for implementation of such a `distributed linking' is presented and an example for linking of two gauge block comparisons is discussed, which refers to the data published in \cite{THALMANN},\cite{DECKER}.
\section{The scenario}
Let us assume the following scenario: A group of laboratories (group A) has participated in a key comparison of a travelling standard A and another group of laboratories (group B) has participated in a key comparison of a travelling standard B. The task of each group was to measure one single measurand of the respective travelling standard (for example the length of a gauge block). In order to link the results of the two key comparisons, a certain group of laboratories (group C) has participated in both key comparisons, {i.\,e.} the laboratories of this group have been members of group A as well as of group B and thus have measured both travelling standards. All laboratories have reported the respective measurement results and the standard uncertainties associated with them. In addition, the laboratories of group C have reported the covariances associated with the measurement results of the two different travelling standards. It should further be assumed, that the standards have been stable, {i.\,e.} that they have not changed their properties in terms of time during the key comparisons.
In order to simplify the organisation of the data, we assume a labelling of the laboratories by assigning, independently of the membership in one of the groups, a unique natural number to each of them, starting by the number one in ascending order. In the following we will use the labels as indices for the respective data. It should be understood, although we are using an ordered set of indices, it is totally arbitrary which index is assigned to which laboratory. Any permutation of the indices used would also serve.
After having assigned the indices, we are able to refer to the index sets of the laboratories, rather than to the groups. Those laboratories, which have {e.\,g.} only measured the travelling standard A are within the index set $\set{I}_\text{A}\operatorname{\setminus}\set{I}_\text{B}$.
\section{The information available}
After having assigned labels (indices) to the laboratories, we are able to summarise the available data in a clear way just by using the same indices for the laboratories and the data. In the following, we will denote the measurement results concerning the travelling standard A obtained by the laboratories of group A by $x_{\text{A},i}$ $(i\in\set{I}_\text{A})$ and the standard uncertainties associated with this results by $u(x_{\text{A},i})$ $(i\in\set{I}_\text{A})$. Accordingly, we will denote the measurement results concerning the travelling standard B obtained by the laboratories of group B by $x_{\text{B},i}$ $(i\in\set{I}_\text{B})$ and the standard uncertainties associated with this results by $u(x_{\text{B},i})$ $(i\in\set{I}_\text{B})$. It then only remains to deal with the covariances, associated with the measurement results concerning both travelling standards, as reported by the laboratories of the group C. In the following we will denote these covariances by $u(x_{\text{A},i},x_{\text{B},i})$ $\bigl(i\in(\set{I}_\text{A}\operatorname{\cap}\set{I}_\text{B})\bigr)$. Equivalently we may use the correlation coefficients, defined by
\begin{equation}
r_{\text{A}\text{B},i}=\frac{u(x_{\text{A},i},x_{\text{B},i})}{u(x_{\text{A},i})u(x_{\text{B},i})},
\qquad i\in(\set{I}_\text{A}\operatorname{\cap}\set{I}_\text{B}),
\eqlbl{1}
\end{equation}
instead of the covariances. We will frequently use the correlation coefficients as abbreviations in the subsequent calculations.
Under the prerequisite, that the two travelling standards under consideration have been stable during the key comparison, the assumption is justified, that all measurement results are the experimental realisation of only two different quantities, {i.\,e.} the measurands of the travelling standards A and B, denoted in the following by $Y_\text{A}$ and $Y_\text{B}$, respectively. In addition we assume, that the measurement results have already been corrected for possible systematic deviations in a suitable way and the uncertainties associated with the corrections have been included in the combined standard uncertainties associated with the corrected results.
Note, that we use upper case letters to denote the measurands and \emph{not} the respective measured value of a measurand, which we will denote by lower case letters throughout the text. We will however not differentiate between a measurand and its value (an ideal value to be estimated, sometimes called \emph{true value}) and use the same letter for both, since the respective meaning follows from the context.
\section{Establishing key comparison reference values}
\subsection{Preliminary remarks}
Using the ordinary arithmetic mean of the measured values to calculate a key comparison reference value (KCRV) can not be justified, because in doing so none of the known uncertainty values are taken into account and thus we do not make use of all available information, as required by the \emph{Guide to the Expression of Uncertainty in Measurement} (GUM) \cite{GUM}. Therefore we shall apply the Bayesian theory of measurement uncertainty \cite{WEISE-WOEGER}, which is entirely based on Bayesian statistics and the principle of maximum entropy (PME). The idea of this theory is first of all to establish a joint probability density function (pdf) of the measurands under consideration by taking into account the data obtained by measurement or otherwise given, together with their associated uncertainties, as well as physical relations and prior information about the measurands. This joint pdf expresses the state of incomplete knowledge about the measurands and is subsequently used to calculate the expectations of the measurands and its associated covariance matrix.
\subsection{The joint probability density function}
Since the measurements of all laboratories participating in the key comparison are assumed to be independent, we can assign probability density functions to the data of each laboratory individually. These pdfs belong to two different types only, because the data types of the groups A and B, excluding the data of group C, are similar and will thus lead to the same type of pdf. The data type of group C, however, is dissimilar from the data types of the groups A and B, excluding the data of group C, and hence we have to expect to get a different type of pdf for this group.
In order to assign pdfs\footnote{We will use here and throughout the document the abbreviation
$$
p(X|Y)=p_{X|Y}(x|y)
$$
for the conditional pdf, in order to simplify the notation of the formulae.} to the data, we follow a suggestion of Jaynes \cite{JAYNES1}, \cite{JAYNES2}, \cite{JAYNES3}, to use the principle of maximum entropy (PME) for this purpose. The PME requires to maximise the relative information entropy \cite{SHANNON}, \cite{KULLBACK} by a suitable pdf under the constraints imposed on the pdf by the normalisation condition and the data. In the process we make the usual assumption, that the measured data and the associated variances and covariances, respectively, may be equated with the respective expectations.
The PME yields under the circumstances at hand
\begin{equation}
p\bigl(x_{\text{A},i},u(x_{\text{A},i})\mid Y_\text{A}\bigr)=\frac{1}{u(x_{\text{A},i})\sqrt{2\pi}}
\exp\left(-\frac{(Y_\text{A}-x_{\text{A},i})^2}{2u^2(x_{\text{A},i})}\right)\,,
\qquad i\in\set{I}_\text{A}\operatorname{\setminus}\set{I}_\text{B}\,,
\eqlbl{4}
\end{equation}
\begin{equation}
p\bigl(x_{\text{B},i},u(x_{\text{B},i})\mid Y_\text{B}\bigr)=\frac{1}{u(x_{\text{B},i})\sqrt{2\pi}}
\exp\left(-\frac{(Y_\text{B}-x_{\text{B},i})^2}{2u^2(x_{\text{B},i})}\right)\,,
\qquad i\in\set{I}_\text{B}\operatorname{\setminus}\set{I}_\text{A}\,,
\eqlbl{5}
\end{equation}
and
\begin{multline}
p\bigl(x_{\text{A},i},x_{\text{B},i},u(x_{\text{A},i}),u(x_{\text{B},i}),r_{\text{A}\text{B},i}\mid Y_\text{A},Y_\text{B}\bigr)=
\frac{1}{2\pi\,u(x_{\text{A},i})u(x_{\text{B},i})\sqrt{1-r^2_{\text{A}\text{B},i}}}
\\
\times\exp\left(-\frac{1}{2(1-r^2_{\text{A}\text{B},i})}
\left[\frac{(Y_\text{A}-x_{\text{A},i})^2}{u^2(x_{\text{A},i})}
-2\frac{r_{\text{A}\text{B},i}(Y_\text{A}-x_{\text{A},i})(Y_\text{B}-x_{\text{B},i})}
{u(x_{\text{A},i})u(x_{\text{B},i})}
+\frac{(Y_\text{B}-x_{\text{B},i})^2}{u^2(x_{\text{B},i})}\right]\right)\,, \\
\qquad i\in\set{I}_\text{A}\operatorname{\cap}\set{I}_\text{B}\,.
\eqlbl{6}
\end{multline}
These pdfs can be esteemed as sampling distributions an thus be used to establish the likelihood function in the usual way. Note that the values $x_{\text{A},i}$ ($i\in\set{I}_\text{A}$) and $x_{\text{B},i}$ ($i\in\set{I}_\text{B}$) as well as their associated uncertainties $u(x_{\text{A},i})$ and $u(x_{\text{B},i})$, respectively, are given data, while $Y_\text{A}$ and $Y_\text{B}$ are unknown quantities.
The prior pdf $p(Y_\text{A},Y_\text{B})$ would allow us to express our knowledge about the quantities $Y_\text{A}$ and $Y_\text{B}$. However, in order to be conservative, we use Jeffreys' prior \cite{JEFFREYS} here, {i.\,e.} we assign a constant to this prior pdf. If more knowledge about $Y_\text{A}$ and $Y_\text{B}$ would be available, the prior could be changed accordingly.
Using Bayes' theorem we obtain the posterior pdf
\begin{equation}
p(Y_\text{A},Y_\text{B}\mid\set{D})=C\rme^{-\chi^2/2}\,,
\eqlbl{8}
\end{equation}
with the set $\set{D}$ representing all known data,
\begin{multline}
\chi^2=\sum_{i\in(\set{I}_\text{A}\operatorname{\setminus}\set{I}_\text{B})}
\frac{(Y_\text{A}-x_{\text{A},i})^2}{u^2(x_{\text{A},i})}
+\sum_{i\in(\set{I}_\text{B}\operatorname{\setminus}\set{I}_\text{A})}
\frac{(Y_\text{B}-x_{\text{B},i})^2}{u^2(x_{\text{B},i})}
\\
+\sum_{i\in(\set{I}_\text{A}\operatorname{\cap}\set{I}_\text{B})}\frac{1}{1-r^2_{\text{A}\text{B},i}}
\left[\frac{(Y_\text{A}-x_{\text{A},i})^2}{u^2(x_{\text{A},i})}
-2\frac{r_{\text{A}\text{B},i}(Y_\text{A}-x_{\text{A},i})(Y_\text{B}-x_{\text{B},i})}
{u(x_{\text{A},i})u(x_{\text{B},i})}
+\frac{(Y_\text{B}-x_{\text{B},i})^2}{u^2(x_{\text{B},i})}\right]
\eqlbl{9}
\end{multline}
and a normalisation constant $C$. This is the posterior pdf of the quantities $Y_\text{A}$ and $Y_\text{B}$ given the data $\set{D}$.
After some algebraic transformations, equation \gl{9} can be written as
\begin{equation}
\chi^2=\frac{1}{1-\tilde{r}_{\text{A}\text{B}}^2}\left(
\frac{(Y_\text{A}-\hat{y}_{\text{A}})^2}{u^2(\hat{y}_{\text{A}})}
-2\tilde{r}_{\text{A}\text{B}}
\frac{(Y_\text{A}-\hat{y}_{\text{A}})(Y_\text{B}-\hat{y}_{\text{B}})}{u(\hat{y}_{\text{A}})u(\hat{y}_{\text{B}})}
+\frac{(Y_\text{B}-\hat{y}_{\text{B}})^2}{u^2(\hat{y}_{\text{B}})}\right)+q^2
\eqlbl{10}
\end{equation}
with
\begin{equation}
\hat{y}_{\text{A}}=\frac{bs_1+cs_2}{ab-c^2}
\eqlbl{11}
\end{equation}
\begin{equation}
\hat{y}_{\text{B}}=\frac{cs_1+as_2}{ab-c^2}
\eqlbl{12}
\end{equation}
\begin{equation}
u(\hat{y}_{\text{A}})=\sqrt{\frac{b}{ab-c^2}}
\eqlbl{13}
\end{equation}
\begin{equation}
u(\hat{y}_{\text{B}})=\sqrt{\frac{a}{ab-c^2}}
\eqlbl{14}
\end{equation}
\begin{equation}
\tilde{r}_{\text{A}\text{B}}=\frac{c}{\sqrt{ab}}
\eqlbl{15}
\end{equation}
\begin{multline}
q^2=\sum_{i\in(\set{I}_\text{A}\operatorname{\setminus}\set{I}_\text{B})}
\frac{(x_{\text{A},i}-\hat{y}_{\text{A}})^2}{u^2(x_{\text{A},i})}
+\sum_{i\in(\set{I}_\text{B}\operatorname{\setminus}\set{I}_\text{A})}
\frac{(x_{\text{B},i}-\hat{y}_{\text{B}})^2}{u^2(x_{\text{B},i})}
\\
+\sum_{i\in(\set{I}_\text{A}\operatorname{\cap}\set{I}_\text{B})}\frac{1}{1-r^2_{\text{A}\text{B},i}}
\left(\frac{(x_{\text{A},i}-\hat{y}_{\text{A}})^2}{u^2(x_{\text{A},i})}
-2\frac{r_{\text{A}\text{B},i}(x_{\text{A},i}-\hat{y}_{\text{A}})(x_{\text{B},i}-\hat{y}_{\text{B}})}
{u(x_{\text{A},i})u(x_{\text{B},i})}
+\frac{(x_{\text{B},i}-\hat{y}_{\text{B}})^2}{u^2(x_{\text{B},i})}\right)
\eqlbl{16}
\end{multline}
and the abbreviations
\begin{equation}
a=\sum_{i\in(\set{I}_\text{A}\operatorname{\setminus}\set{I}_\text{B})}\frac{1}{u^2(x_{\text{A},i})}
+\sum_{i\in(\set{I}_\text{A}\operatorname{\cap}\set{I}_\text{B})}
\frac{1}{(1-r^2_{\text{A}\text{B},i})u^2(x_{\text{A},i})}
\eqlbl{17}
\end{equation}
\begin{equation}
b=\sum_{i\in(\set{I}_\text{B}\operatorname{\setminus}\set{I}_\text{A})}\frac{1}{u^2(x_{\text{B},i})}
+\sum_{i\in(\set{I}_\text{A}\operatorname{\cap}\set{I}_\text{B})}
\frac{1}{(1-r^2_{\text{A}\text{B},i})u^2(x_{\text{B},i})}
\eqlbl{18}
\end{equation}
\begin{equation}
c=\sum_{i\in(\set{I}_\text{A}\operatorname{\cap}\set{I}_\text{B})}
\frac{r_{\text{A}\text{B},i}}{(1-r^2_{\text{A}\text{B},i})u(x_{\text{A},i})u(x_{\text{B},i})}
\eqlbl{19}
\end{equation}
\begin{equation}
s_1=\sum_{i\in(\set{I}_\text{A}\operatorname{\setminus}\set{I}_\text{B})}\frac{x_{\text{A},i}}{u^2(x_{\text{A},i})}
+\sum_{i\in(\set{I}_\text{A}\operatorname{\cap}\set{I}_\text{B})}\frac{1}{1-r^2_{\text{A}\text{B},i}}
\left(\frac{x_{\text{A},i}}{u^2(x_{\text{A},i})}
-\frac{r_{\text{A}\text{B},i}x_{\text{B},i}}{u(x_{\text{A},i})u(x_{\text{B},i})}\right)
\eqlbl{20}
\end{equation}
\begin{equation}
s_2=\sum_{i\in(\set{I}_\text{B}\operatorname{\setminus}\set{I}_\text{A})}
\frac{x_{\text{B},i}}{u^2(x_{\text{B},i})}
+\sum_{i\in(\set{I}_\text{A}\operatorname{\cap}\set{I}_\text{B})}\frac{1}{1-r^2_{\text{A}\text{B},i}}
\left(\frac{x_{\text{B},i}}{u^2(x_{\text{B},i})}
-\frac{r_{\text{A}\text{B},i}x_{\text{A},i}}{u(x_{\text{A},i})u(x_{\text{B},i})}\right)
\eqlbl{21}
\end{equation}
Note, that the auxiliary quantities $a$, $b$, $c$, $s_1$, $s_2$, and thereby the quantities $\hat{y}_{\text{A}}$, $\hat{y}_{\text{B}}$, $u(\hat{y}_{\text{A}})$, $u(\hat{y}_{\text{B}})$, $\tilde{r}_{\text{A}\text{B}}$ as well, depend only on the data obtained by the measurement. Thus $q^2$ does not depend on the quantities $Y_\text{A}$ and $Y_\text{B}$ and can hence be absorbed by the normalisation constant. Therefore, after renormalisation, we finally get
\begin{multline}
p(Y_\text{A},Y_\text{B}|\set{D})=
\frac{1}{2\pi u(\hat{y}_{\text{A}})u(\hat{y}_{\text{B}})\sqrt{1-\tilde{r}_{\text{A}\text{B}}^2}} \\
\times\exp\left[-\frac{1}{2(1-\tilde{r}_{\text{A}\text{B}}^2)}\left(
\frac{(Y_\text{A}-\hat{y}_{\text{A}})^2}{u^2(\hat{y}_{\text{A}})}
-2\tilde{r}_{\text{A}\text{B}}
\frac{(Y_\text{A}-\hat{y}_{\text{A}})(Y_\text{B}-\hat{y}_{\text{B}})}{u(\hat{y}_{\text{A}})u(\hat{y}_{\text{B}})}
+\frac{(Y_\text{B}-\hat{y}_{\text{B}})^2}{u^2(\hat{y}_{\text{B}})}\right)
\right]\,,
\eqlbl{22}
\end{multline}
{i.\,e.} the posterior pdf of the quantities $Y_\text{A}$ and $Y_\text{B}$ is a bivariate Gaussian pdf.
\subsection{The key comparison reference values}
Since the posterior pdf of the quantities $Y_\text{A}$ and $Y_\text{B}$ is a bivariate Gaussian pdf, it follows that
\begin{equation}
\operatorname{\mathsf{E}}\left[Y_\text{A}\right]=\hat{y}_{\text{A}}
\qquad\text\qquad
\operatorname{\mathsf{E}}\left[Y_\text{B}\right]=\hat{y}_{\text{B}}
\eqlbl{23}
\end{equation}
are the expectations of the quantities $Y_\text{A}$ and $Y_\text{B}$, respectively, and
\begin{equation}
\matr{U}_\vect{Y}=
\begin{pmatrix}
u^2(\hat{y}_{\text{A}}) & \tilde{r}_{\text{A}\text{B}}u(\hat{y}_{\text{A}})u(\hat{y}_{\text{B}}) \\
\tilde{r}_{\text{A}\text{B}}u(\hat{y}_{\text{A}})u(\hat{y}_{\text{B}}) & u^2(\hat{y}_{\text{B}})
\end{pmatrix}
\eqlbl{24}
\end{equation}
their associated covariance matrix. We now identify the values $\hat{y}_{\text{A}}$ and $\hat{y}_{\text{B}}$ as the key comparison reference values and the matrix $\matr{U}_\vect{Y}$ as their associated covariance matrix.
For convenience we summarise the results, using the covariances rather than the correlation coefficients.
\begin{equation}
a=\sum_{i\in(\set{I}_\text{A}\operatorname{\setminus}\set{I}_\text{B})}\frac{1}{u^2(x_{\text{A},i})}
+\sum_{i\in(\set{I}_\text{A}\operatorname{\cap}\set{I}_\text{B})}
\frac{u^2(x_{\text{B},i})}{u^2(x_{\text{A},i})u^2(x_{\text{B},i})-u^2(x_{\text{A},i},x_{\text{B},i})}\,,
\eqlbl{25}
\end{equation}
\begin{equation}
b=\sum_{i\in(\set{I}_\text{B}\operatorname{\setminus}\set{I}_\text{A})}\frac{1}{u^2(x_{\text{B},i})}
+\sum_{i\in(\set{I}_\text{A}\operatorname{\cap}\set{I}_\text{B})}
\frac{u^2(x_{\text{A},i})}{u^2(x_{\text{A},i})u^2(x_{\text{B},i})-u^2(x_{\text{A},i},x_{\text{B},i})}\,,
\eqlbl{26}
\end{equation}
\begin{equation}
c=\sum_{i\in(\set{I}_\text{A}\operatorname{\cap}\set{I}_\text{B})}
\frac{u(x_{\text{A},i},x_{\text{B},i})}{u^2(x_{\text{A},i})u^2(x_{\text{B},i})-u^2(x_{\text{A},i},x_{\text{B},i})}\,,
\eqlbl{27}
\end{equation}
\begin{equation}
s_1=\sum_{i\in(\set{I}_\text{A}\operatorname{\setminus}\set{I}_\text{B})}\frac{x_{\text{A},i}}{u^2(x_{\text{A},i})}
+\sum_{i\in(\set{I}_\text{A}\operatorname{\cap}\set{I}_\text{B})}
\frac{u^2(x_{\text{B},i})x_{\text{A},i}-u(x_{\text{A},i},x_{\text{B},i})x_{\text{B},i}}
{u^2(x_{\text{A},i})u^2(x_{\text{B},i})-u^2(x_{\text{A},i},x_{\text{B},i})}\,,
\eqlbl{28}
\end{equation}
\begin{equation}
s_2=\sum_{i\in(\set{I}_\text{B}\operatorname{\setminus}\set{I}_\text{A})}\frac{x_{\text{B},i}}{u^2(x_{\text{B},i})}
+\sum_{i\in(\set{I}_\text{A}\operatorname{\cap}\set{I}_\text{B})}
\frac{u^2(x_{\text{A},i})x_{\text{B},i}-u(x_{\text{A},i},x_{\text{B},i})x_{\text{A},i}}
{u^2(x_{\text{A},i})u^2(x_{\text{B},i})-u^2(x_{\text{A},i},x_{\text{B},i})}\,.
\eqlbl{29}
\end{equation}
Key comparison reference values (KCRVs):
\begin{equation}
\hat{y}_{\text{A}}=\frac{bs_1+cs_2}{ab-c^2}\,,
\eqlbl{30}
\end{equation}
\begin{equation}
\hat{y}_{\text{B}}=\frac{cs_1+as_2}{ab-c^2}\,.
\eqlbl{31}
\end{equation}
Standard uncertainties of the KCRVs:
\begin{equation}
u(\hat{y}_{\text{A}})=\sqrt{\frac{b}{ab-c^2}}\,,
\eqlbl{32}
\end{equation}
\begin{equation}
u(\hat{y}_{\text{B}})=\sqrt{\frac{a}{ab-c^2}}\,.
\eqlbl{33}
\end{equation}
Covariance of the KCRVs:
\begin{equation}
u(\hat{y}_{\text{A}},\hat{y}_{\text{B}})=\frac{c}{ab-c^2}\,.
\eqlbl{34}
\end{equation}
\section{Degrees of equivalence}
The degrees of equivalence (DOE) are defined to be
\begin{equation}
d_{\text{A},i}=x_{\text{A},i}-\hat{y}_{\text{A}}\,,
\qquad i\in\set{I}_\text{A}\,,
\eqlbl{35}
\end{equation}
and
\begin{equation}
d_{\text{B},i}=x_{\text{B},i}-\hat{y}_{\text{B}}\,,
\qquad i\in\set{I}_\text{B}\,.
\eqlbl{36}
\end{equation}
These values give the deviations of each laboratory from the respective KCRV.
Following the rules for the propagation of variances, we obtain
\begin{equation}
u(d_{\text{A},i})=\sqrt{u^2(x_{\text{A},i})-u^2(\hat{y}_{\text{A}})}\,,
\qquad i\in\set{I}_\text{A}\,,
\eqlbl{37}
\end{equation}
and
\begin{equation}
u(d_{\text{B},i})=\sqrt{u^2(x_{\text{B},i})-u^2(\hat{y}_{\text{B}})}\,,
\qquad i\in\set{I}_\text{B}\,,
\eqlbl{38}
\end{equation}
for the respective uncertainties. The negative sign under the square root is due to the correlations $u(\hat{y}_{\text{A}},x_{\text{B},k})=u^2(\hat{y}_{\text{A}})$ and $u(\hat{y}_{\text{B}},x_{\text{B},k})=u^2(\hat{y}_{\text{B}})$ as shown in the appendix.
\section{Conformity test}
After the evaluation procedure has been accomplished, we have to check, whether the determined results do conform to the data given. This can be achieved in the following way. We observe, that on the one hand $\chi^2$ is represented by equation \gl{9} and on the other hand by equation \gl{10}. Thus, we can calculate the expectation $\operatorname{\mathsf{E}}\left[\chi^2\right]$ in two different ways, yielding
\begin{equation}
\operatorname{\mathsf{E}}\left[\chi^2\right]=N=2+q^2\,,
\eqlbl{42}
\end{equation}
with
\begin{equation}
N=\operatorname{card}\set{I}_\text{A}+\operatorname{card}\set{I}_\text{B}
\eqlbl{44}
\end{equation}
and $q^2$ given by equation \gl{16}. Thus, we obtain
\begin{equation}
q^2=N-2\,.
\eqlbl{43}
\end{equation}
Note that the right hand side of this equation may be regarded as the number of the degrees of freedom, because it is just the total number of measured data diminished by the number of estimated quantities.
The condition \gl{43} can, however, not be expected to be strictly fulfilled, but $N-2$ is the most probable value of $q^2$, because if the data are actually statistically distributed according to the model and the measurement uncertainties given, equation \gl{43} is the consequence. A significant deviation from this result is a signal that either the model is wrong or the data are suspect. Assuming the model to be correct, an outcome $q^2>N-2$ leads to the conclusion, that there is a strong possibility that either some of the measurement uncertainties have been underestimated or the respective measurement results contain uncorrected (or unknown) systematic deviations. Thus
\begin{equation}
q^2\le N-2
\eqlbl{45}
\end{equation}
can be taken as an indication of the conformity of the estimated results with the data.
\section{Examples}
In order to demonstrate the application of the method proposed in this paper, we show two examples. Our first example is a synthetic one. Data have been produced by a simulated sampling from Gaussian distributions as given in equations \gl{4} to \gl{6}, using the parameters\footnote{$\sigma$ and $\rho$ denote the standard deviation and the correlation coefficient, respectively, of a Gaussian distribution.} $Y_\text{A}=110$, $\sigma_\text{A}=20$, $Y_\text{B}=120$, $\sigma_\text{B}=50$, $\rho=0,5$ and a sample size of $n=50$. Subsequently the sampled data have been evaluated by using the usual formulae for the sample mean, the sample variance and the sample covariance. The resulting data thus obtained as well as their evaluation are shown in \tab{2a} and \tab{2b}.
\begin{table}[ht]
\caption{Data of the synthetic example obtained from the sampled data by application of the usual formulae for the sample mean, the sample variance and the sample covariance.}
\tbl{2a}
\begin{center}
\begin{tabular}{|l|rr|rr|r|}
\hline
Institute & $x_{\text{A},i}$ & $u(x_{\text{A},i})$ &
$x_{\text{B},i}$ & $u(x_{\text{B},i})$ & $r_{\text{A}\text{B},i}$ \\
\hline
LAB-01 & 113,4 & 2,9 & & & \\
LAB-02 & 112,1 & 2,8 & & & \\
LAB-03 & 113,0 & 2,5 & & & \\
LAB-04 & 110,6 & 2,6 & & & \\
LAB-05 & 109,4 & 2,4 & & & \\
LAB-06 & 107,0 & 2,6 & & & \\
LAB-07 & 104,7 & 2,8 & & & \\
LAB-08 & 109,0 & 2,6 & & & \\
\hline
LAB-09 & 111,0 & 2,4 & 120,1 & 6,5 & 0,8 \\
LAB-10 & 109,4 & 2,8 & 117,3 & 7,3 & 0,8 \\
LAB-11 & 111,1 & 2,8 & 125,0 & 6,4 & 0,8 \\
LAB-12 & 115,3 & 2,4 & 135,7 & 6,7 & 0,7 \\
\hline
LAB-13 & & &129,7 & 6,1 & \\
LAB-14 & & &129,1 & 7,5 & \\
LAB-15 & & &125,0 & 7,1 & \\
LAB-16 & & &123,6 & 6,6 & \\
LAB-17 & & &123,0 & 6,9 & \\
\hline
\end{tabular}
\end{center}
\end{table}
\begin{table}[ht]
\caption{Degrees of equivalence for the deviation from the nominal values and their associated standard uncertainties, as well as the key comparison reference values and their associated standard uncertainties for the groups A and B, respectively, obtained by the application of the linking method proposed in this paper for the synthetic example.}
\tbl{2b}
\begin{center}
\begin{tabular}{|l|rr|rr|}
\hline
Institute & $d_{\text{A},i}$ & $u(d_{\text{A},i})$ & $d_{\text{B},i}$ & $u(d_{\text{B},i})$ \\
\hline
LAB-01 & 2,491 & 2,815 & & \\
LAB-02 & 1,191 & 2,712 & & \\
LAB-03 & 2,091 & 2,401 & & \\
LAB-04 & -0,309 & 2,505 & & \\
LAB-05 & -1,509 & 2,296 & & \\
LAB-06 & -3,909 & 2,505 & & \\
LAB-07 & -6,209 & 2,712 & & \\
LAB-08 & -1,909 & 2,505 & & \\
LAB-09 & 0,091 & 2,296 & -3,779 & 6,196 \\
LAB-10 & -1,509 & 2,712 & -6,579 & 7,030 \\
LAB-11 & 0,191 & 2,712 & 1,121 & 6,091 \\
LAB-12 & 4,391 & 2,296 & 11,821 & 6,405 \\
LAB-13 & & & 5,821 & 5,775 \\
LAB-14 & & & 5,221 & 7,238 \\
LAB-15 & & & 1,121 & 6,822 \\
LAB-16 & & & -0,279 & 6,300 \\
LAB-17 & & & -0,879 & 6,614 \\
\hline
& \multicolumn{4}{l|}{\rule{0pt}{4mm}$\hat{y}_{\text{A}}=110,909\,,\: u(\hat{y}_{\text{A}})=0,698$} \\
& \multicolumn{4}{l|}{$\hat{y}_{\text{B}}=123,879\,,\: u(\hat{y}_{\text{B}})=1,966$} \\
& \multicolumn{4}{l|}{\rule{0pt}{4mm}$q^2/(N-2)=0,89$} \\
\hline
\end{tabular}
\end{center}
\end{table}
As can be seen, the data do pass the conformity test. However, the KCRVs are slightly larger than the values used for the simulation. This is caused by the fact, that the correlation coefficients estimated from the sampled data tend to be too large compared to the correlation coefficient used for the simulation. This demonstrates the effect of small sample sizes.
As a second example data from the CCL-K1 \cite{THALMANN} and SIM.L-K1 \cite{DECKER} gauge block comparisons have been evaluated. The three institutes CENAM, NIST and NRC participated in both comparisons and hence serve as linking laboratories. For convenience, the results and their associated standard uncertainties, as reported in the respective publications, are repeated in \tab{2}. Values for the covariances have unfortunately not been reported, although systematic deviations present when measuring the gauge blocks in each of the linking laboratories with the same measuring system inevitably cause correlations, even if a correction for these systematic effects, as usually good practise, has been applied. Thus, following the recommendation given in the GUM in lack of this information\footnote{It should be noted, that this recommendation is consistent with the \emph{modus operandi} of the Bayesian theory, where only \emph{known} information is taken into account.}, all covariances have been assumed to be zero, knowing that this is certainly not true.
\begin{table}[ht]
\caption{Results for the deviations from the nominal length and their associated standard uncertainties as taken from the CCL-K1 \cite{THALMANN} and SIM.L-K1 \cite{DECKER} gauge block comparison reports (steel gauge block, nominal length 100 mm).}
\tbl{2}
\begin{center}
\begin{tabular}{|l|rr|rr|}
\hline
Institute & $x_{\text{A},i}/$nm & $u(x_{\text{A},i})/$nm & $x_{\text{B},i}/$nm & $u(x_{\text{B},i})/$nm \\
\hline
METAS & -96,0 & 13,0 & & \\
NPL & -140,0 & 33,0 & & \\
BNM-LNE & -110,0 & 16,0 & & \\
KRISS & -104,3 & 20,6 & & \\
NRLM & -89,4 & 16,3 & & \\
VNIIM & -104,0 & 15,0 & & \\
CSIRO & -114,0 & 16,0 & & \\
NIM & -90,0 & 10,3 & & \\
\hline
NIST & -117,0 & 17,9 & -100,0 & 18,0 \\
CENAM & -119,0 & 18,7 & -93,0 & 23,0 \\
NRC & -126,0 & 24,0 & -124,0 & 26,0 \\
\hline
INMETRO1 & & & -98,0 & \textbf{4,0} \\
INMETRO2 & & & -68,0 & 29,0 \\
INTI & & &-104,0 & 21,0 \\
CEM & & &-148,0 & 17,0 \\
\hline
\end{tabular}
\end{center}
\end{table}
\begin{table}[ht]
\caption{Degrees of equivalence for the deviation from the nominal values and their associated standard uncertainties, as well as the key comparison reference values and their associated standard uncertainties for the groups A and B, respectively, obtained by the application of the linking method proposed in this paper (results obtained for the original data in \tab{2}).}
\tbl{3}
\begin{center}
\begin{tabular}{|l|rr|rr|}
\hline
Institute & $d_{\text{A},i}/$nm & $u(d_{\text{A},i})/$nm & $d_{\text{B},i}/$nm & $u(d_{\text{B},i})/$nm \\
\hline
METAS & 7,6 & 12,1 & & \\
NPL & -36,4 & 32,6 & & \\
BNM-LNE & -6,4 & 15,2 & & \\
KRISS & -0,7 & 20,0 & & \\
NRLM & 14,2 & 15,6 & & \\
VNIIM & -0,4 & 14,2 & & \\
CSIRO & -10,4 & 15,2 & & \\
NIM & 13,6 & 9,1 & & \\
\hline
NIST & -13,4 & 17,2 & 0,5 & 17,6 \\
CENAM & -15,4 & 18,1 & 7,5 & 22,7 \\
NRC & -22,4 & 23,5 & -23,5 & 25,7 \\
\hline
INMETRO1 & & & 2,5 & 1,7 \\
INMETRO2 & & & 32,5 & 28,8 \\
INTI & & & -3,5 & 20,7 \\
CEM & & & -47,5 & 16,6 \\
\hline
& \multicolumn{4}{l|}{\rule{0pt}{4mm}$\hat{y}_{\text{A}}=-103,6\,,\: u(\hat{y}_{\text{A}})=4,9$ nm} \\
& \multicolumn{4}{l|}{$\hat{y}_{\text{B}}=-100,5\,,\: u(\hat{y}_{\text{B}})=3,6$ nm} \\
& \multicolumn{4}{l|}{\rule{0pt}{4mm}$q^2/(N-2)=1,07$ (conformity test failed)} \\
\hline
\end{tabular}
\end{center}
\end{table}
\begin{table}[ht]
\caption{Degrees of equivalence for the deviation from the nominal values and their associated standard uncertainties, as well as the key comparison reference values and their associated standard uncertainties for the groups A and B, respectively, obtained by the application of the linking method proposed in this paper (results obtained after the uncertainty of the institute INMETRO1 has been increased from 4,0 nm to 11,2 nm).}
\tbl{4}
\begin{center}
\begin{tabular}{|l|rr|rr|}
\hline
Institute & $d_{\text{A},i}/$nm & $u(d_{\text{A},i})/$nm & $d_{\text{B},i}/$nm & $u(d_{\text{B},i})/$nm \\
\hline
METAS & 7,6 & 12,1 & & \\
NPL & -36,4 & 32,6 & & \\
BNM-LNE & -6,4 & 15,2 & & \\
KRISS & -0,7 & 20,0 & & \\
NRLM & 14,2 & 15,6 & & \\
VNIIM & -0,4 & 14,2 & & \\
CSIRO & -10,4 & 15,2 & & \\
NIM & 13,6 & 9,1 & & \\
\hline
NIST & -13,4 & 17,2 & 6,7 & 16,6 \\
CENAM & -15,4 & 18,1 & 13,7 & 22,0 \\
NRC & -22,4 & 23,5 & -17,3 & 25,1 \\
\hline
INMETRO1 & & & 8,7 & 8,9 \\
INMETRO2 & & & 38,7 & 28,2 \\
INTI & & & 2,7 & 19,9 \\
CEM & & & -41,3 & 15,6 \\
\hline
& \multicolumn{4}{l|}{\rule{0pt}{4mm}$\hat{y}_{\text{A}}=-103,6\,,\: u(\hat{y}_{\text{A}})=4,9$ nm} \\
& \multicolumn{4}{l|}{$\hat{y}_{\text{B}}=-106,7\,,\: u(\hat{y}_{\text{B}})=6,8$ nm} \\
& \multicolumn{4}{l|}{\rule{0pt}{4mm}$q^2/(N-2)=1,00$} \\
\hline
\end{tabular}
\end{center}
\end{table}
The degrees of equivalence for the deviation from the nominal values and their associated standard uncertainties, as well as the key comparison reference values and their associated standard uncertainties for the groups A and B, respectively, obtained by the application of the linking method proposed in this paper are given in \tab{3} for the original data. It emerges, however, that the data do not pass the conformity test.
Assuming that the evaluation model is correct and that the respective measurement results have been corrected for systematic deviations, the test result leads to the conclusion, that there is a strong possibility that some of the measurement uncertainties have been underestimated. Inspecting the uncertainty values given in \tab{2}, it is obvious that ``\emph{\dots INMETRO1 submitted very optimistic uncertainty claims relative to conventional capabilities}'' \cite{DECKER} (the corresponding value is marked in bold face in the table). Since the gauge blocks have been measured at INMETRO by two completely different instruments, involving different staff members (data denoted by INMETRO1 represent results from a research grade instrument, while those denoted by INMETRO2 represent results from the routine gauge block calibration service offered by INMETRO) \cite{DECKER}, it is reasonable to question the very low uncertainty stated for INMETRO1. Therefore, this uncertainty value has been increased here from 4,0 nm to 11,2 nm, which is the smallest value in order that the data pass the conformity test. The respective results obtained with this change are given in \tab{4}.
A comparison of the results given in \tab{3} with those given in \tab{4} shows, that the reference value for group B and its associated standard uncertainty has noticeably changed. This strongly emphasises, that it is imperative to do a \emph{rigorous} conformity test. The $E_n$ criterion, for example, did not reveal any problem with the data from INMETRO1 during the SIM.l-K1 comparison. In fact the values $\abs{E_{95}}$ as well as $\abs{E}$ for the evaluation of the 100 mm steel gauge block data have even been comparable to those for the NRC \cite{DECKER}. The usual $\chi^2$ criterion at a 5 \% confidence level also fails to indicate a problem with the data.
\section{Conclusion}
In this contribution we have derived formulae for the linking parameters key comparison reference value (KCRV) and degree of equivalence (DOE) for analysis of two comparisons based on Bayesian statistics, {i.\,e.} taking all known information into account but without the necessity to choose one comparison as a primary one. The applicability of the linking approach was demonstrated on an example from gauge block metrology.
The proposed approach can be described as a type of `distributed linking', which has been looked for especially in the area of dimensional metrology within the working groups of the CCL.
This `distributed linking' method can, for example, be used to compare the results of two comparisons which were started in parallel using different transfer standards and where it does not seem meaningful to choose one comparison to be primary (defining a KCRV) and the other secondary (linking its comparison results to the KCRV of the first). Such comparisons with more than one loop have been organised if the number of interested participants became rather large, like e.~g. in the finished line scale comparison EUROMET.L-K7 with 31 participants \cite{EUROMET}.
In a strict Bayesian sense, using the `distributed linking' approach takes all known information from both comparisons into account to determine the linking parameters, i.~e. all the measurement results and their uncertainties. Instead, in application of the hierarchical linking approaches one could in principle argue that not all of the information is taken into account, because e.~g. the results gained from a RMO key comparison are not taken into account to reflect or re-analyse the results of a CIPM key comparison, although by the twofold involvement of the linking laboratories the level of knowledge is increased for both comparisons. On the other hand the results of a subsequent RMO comparison linked to a key comparison by the described `distributed linking' approach would also change the KCRV of the (already finished) key comparison. We see that this would impose new challenges for the operation of the KCDB.
The described approach can be extended to link more than two comparisons. Moreover, additional knowledge, as {e.\,g.} results from prior measurements, may be taken into account by choosing a suitable prior pdf.
\section{Acknowledgement}
The authors gratefully acknowledge the valuable comments of A. Balsamo from INRIM (Italy) and the suggestion to add the information given in appendix 2.
\section*{Appendix 1 (Calculation of covariances)}
In equations \gl{37} and \gl{38} the relations $u(\hat{y}_{\text{A}},x_{\text{B},k})=u^2(\hat{y}_{\text{A}})$ and $u(\hat{y}_{\text{B}},x_{\text{B},k})=u^2(\hat{y}_{\text{B}})$, respectively, have implicitly been used. In order to show that they are valid, we introduce the estimators
\begin{equation}
\hat{Y}_{\text{A}}=\frac{bS_1+cS_2}{ab-c^2}
\eqlbl{30b}
\end{equation}
and
\begin{equation}
\hat{Y}_{\text{B}}=\frac{cS_1+aS_2}{ab-c^2}\,,
\eqlbl{31b}
\end{equation}
where
\begin{equation}
S_1=\sum_{i\in(\set{I}_\text{A}\operatorname{\setminus}\set{I}_\text{B})}\frac{X_{\text{A},i}}{u^2(x_{\text{A},i})}
+\sum_{i\in(\set{I}_\text{A}\operatorname{\cap}\set{I}_\text{B})}
\frac{u^2(x_{\text{B},i})X_{\text{A},i}-u(x_{\text{A},i},x_{\text{B},i})X_{\text{B},i}}
{u^2(x_{\text{A},i})u^2(x_{\text{B},i})-u^2(x_{\text{A},i},x_{\text{B},i})}
\eqlbl{28b}
\end{equation}
and
\begin{equation}
S_2=\sum_{i\in(\set{I}_\text{B}\operatorname{\setminus}\set{I}_\text{A})}\frac{X_{\text{B},i}}{u^2(x_{\text{B},i})}
+\sum_{i\in(\set{I}_\text{A}\operatorname{\cap}\set{I}_\text{B})}
\frac{u^2(x_{\text{A},i})X_{\text{B},i}-u(x_{\text{A},i},x_{\text{B},i})X_{\text{A},i}}
{u^2(x_{\text{A},i})u^2(x_{\text{B},i})-u^2(x_{\text{A},i},X_{\text{B},i})}\,.
\eqlbl{29b}
\end{equation}
are functions of the random variables $X_{\text{A},i}$ ($i\in\set{I}_\text{A}$) and $X_{\text{B},i}$ ($i\in\set{I}_\text{B}$). We assume the measured values $x_{\text{A},i}$ ($i\in\set{I}_\text{A}$) and $x_{\text{B},i}$ ($i\in\set{I}_\text{B}$) to be a realisation of these random variables.
Since $\operatorname{\mathsf{E}}\left[X_{\text{A},i}\right]=Y_{\text{A}}$, $\operatorname{\mathsf{E}}\left[X_{\text{B},i}\right]=Y_{\text{B}}$, $\operatorname{\mathsf{Var}}\left[X_{\text{A},i}\right]=u^2(x_{\text{A},i})$, $\operatorname{\mathsf{Var}}\left[X_{\text{B},i}\right]=u^2(x_{\text{B},i})$ and $\operatorname{\mathsf{Cov}}\left[X_{\text{A},i},X_{\text{B},i}\right]=u(x_{\text{A},i},x_{\text{B},i})$ ($i\in\set{I}_\text{A}$, $i\in\set{I}_\text{B}$)
it can be verified, that
\begin{equation}
\operatorname{\mathsf{E}}\left[\hat{S}_1\right]=aY_{\text{A}}-cY_{\text{B}}\,,
\qquad
\operatorname{\mathsf{E}}\left[\hat{S}_2\right]=bY_{\text{B}}-cY_{\text{A}}\,,
\eqlbl{x1}
\end{equation}
as well as
\begin{equation}
\operatorname{\mathsf{Var}}\left[\hat{S}_1\right]=a\,,
\qquad
\operatorname{\mathsf{Var}}\left[\hat{S}_2\right]=b\,,
\qquad
\operatorname{\mathsf{Cov}}\left[\hat{S}_1,\hat{S}_2\right]=-c
\eqlbl{x2}
\end{equation}
is valid. Using these results, we obtain
\begin{equation}
\operatorname{\mathsf{E}}\left[\hat{Y}_{\text{A}}\right]=\hat{y}_{\text{A}}\,,
\qquad
\operatorname{\mathsf{E}}\left[\hat{Y}_{\text{B}}\right]=\hat{y}_{\text{B}}\,,
\eqlbl{x3}
\end{equation}
\begin{equation}
\operatorname{\mathsf{Var}}\left[\hat{Y}_{\text{A}}\right]=u(\hat{y}_{\text{A}})\,,
\qquad
\operatorname{\mathsf{Var}}\left[\hat{Y}_{\text{B}}\right]=u(\hat{y}_{\text{B}})\,,
\qquad
\operatorname{\mathsf{Cov}}\left[\hat{Y}_{\text{A}},\hat{Y}_{\text{B}}\right]=u(\hat{y}_{\text{A}},\hat{y}_{\text{B}})\,,
\eqlbl{x4}
\end{equation}
{i.\,e.} the estimators $\hat{Y}_{\text{A}}$ and $\hat{Y}_{\text{B}}$ yield the correct results as given by equations \gl{30} to \gl{34}.
We observe that
\begin{equation}
\operatorname{\mathsf{Cov}}\left[\hat{S}_1,X_{\text{A},k}\right]=\operatorname{\mathsf{Cov}}\left[\hat{S}_2,X_{\text{B},k}\right]=1
\eqlbl{x5}
\end{equation}
and
\begin{equation}
\operatorname{\mathsf{Cov}}\left[\hat{S}_1,X_{\text{B},k}\right]=\operatorname{\mathsf{Cov}}\left[\hat{S}_2,X_{\text{A},k}\right]=0\,.
\eqlbl{x6}
\end{equation}
Thus
\begin{equation}
u(\hat{y}_{\text{A}},x_{\text{A},k})=
\operatorname{\mathsf{Cov}}\left[\hat{Y}_{\text{A}},X_{\text{A},k}\right]=u^2(\hat{y}_{\text{A}})
\eqlbl{x7}
\end{equation}
and
\begin{equation}
u(\hat{y}_{\text{B}},x_{\text{B},k})=
\operatorname{\mathsf{Cov}}\left[\hat{Y}_{\text{B}},X_{\text{B},k}\right]=u^2(\hat{y}_{\text{B}})\,.
\eqlbl{x8}
\end{equation}
This is what we wanted to show.
\section*{Appendix 2 (Uncorrelated measurement results)}
Here we consider the possibility that the measurement results of the linking laboratories (group C) have to be treated as if they were uncorrelated, because the respective covariances have not been reported. In this case we have to set all covariances $u(x_{\text{A},i},x_{\text{B},i})$ ($i\in(\set{I}_\text{A}\operatorname{\cap}\set{I}_\text{B})$) equal to zero in the equations \gl{25} to \gl{29}. This yields
\begin{equation}
a=\sum_{i\in\set{I}_\text{A}}\frac{1}{u^2(x_{\text{A},i})}\,,
\eqlbl{x25}
\end{equation}
\begin{equation}
b=\sum_{i\in\set{I}_\text{B}}\frac{1}{u^2(x_{\text{B},i})}\,,
\eqlbl{x26}
\end{equation}
\begin{equation}
c=0\,,
\eqlbl{x27}
\end{equation}
\begin{equation}
s_1=\sum_{i\in\set{I}_\text{A}}\frac{x_{\text{A},i}}{u^2(x_{\text{A},i})}\,,
\eqlbl{x28}
\end{equation}
\begin{equation}
s_2=\sum_{i\in\set{I}_\text{B}}\frac{x_{\text{B},i}}{u^2(x_{\text{B},i})}\,.
\eqlbl{x29}
\end{equation}
Thus, we obtain from equations \gl{30} and \gl{31} for the key comparison reference values
\begin{equation}
\hat{y}_{\text{A}}=\frac{s_1}{a}
\qquad\text{and}\qquad
\hat{y}_{\text{B}}=\frac{s_2}{b}\,,
\eqlbl{x31}
\end{equation}
and from equations \gl{32} and \gl{33} for their associated standard uncertainties
\begin{equation}
u(\hat{y}_{\text{A}})=\frac{1}{\sqrt{a}}
\qquad\text{and}\qquad
u(\hat{y}_{\text{B}})=\frac{1}{\sqrt{b}}\,,
\eqlbl{x33}
\end{equation}
{i.\,e.} the usual weighted mean results of for the KCRVs separately for each group participating in the key comparisons. The results of the groups are independent of each other as indicated by the covariance of the KCRVs, $u(\hat{y}_{\text{A}},\hat{y}_{\text{B}})=0$, as obtained from equation \gl{34}. This demonstrates that the linking procedure essentially depends on the covariances reported by the linking laboratories.
\Bibliography{16}
\bibitem{CIPM-MRA}
Mutual recognition of national measurement standards and of calibration and measurement certificates issued by national metrology institutes, International Committee of Weights and Measures (CIPM), 14 October 1999,
(http://www1.bipm.org/utils/en/pdf/mra\_2003.pdf).
\bibitem{CIPM-MRA-GUIDE}
Guidelines for CIPM key comparisons, International Committee of Weights and Measures (CIPM), 1 March 1999,
(http://www1.bipm.org/utils/en/pdf/guidelines.pdf).
\bibitem{KCDB}
KCDB, http://www.bipm.org/en/cipm-mra/mra-kcdb/
\bibitem{STEELE}
Steele A G, Wood B M and Douglas R J 2005, \textit{Linking key comparison data to appendix B}, TEMPMEKO 2004 : 9th International Symposium on Temperature and Thermal Measurements in Industry and Science (Dubrovnik, Croatia, June 22-25, 2004), p. 1087-1092, vol. 2
\bibitem{NIELSEN}
Nielsen L 2000, \textit{Evaluation of measurement intercomparisons by the method of least squares}, Danish Institute of Fundamental Metrology, Technical Report DFM-99-R39
\bibitem{ELSTER}
Elster C, Link A and Wöger W 2003, \textit{Proposal for linking the results of CIPM and RMO key comparisons}, Metrologia \textbf{40}, 189–94
\bibitem{WHITE}
White D R 2004, \textit{On the analysis of measurement comparisons}, Metrologia \textbf{41}, 122–31
\bibitem{SUTTON}
Sutton C M 2004, \textit{Analysis and linking of international measurement comparisons}, Metrologia \textbf{41}, 272–7
\bibitem{THALMANN}
Thalmann R 2002, \textit{CCL key comparison: calibration of gauge blocks by interferometry}, Metrologia \textbf{39}, 165–77
\bibitem{CCL}
CCL/WGDM/09-22: http://www.bipm.org/wg/CCL/CCL-WG/Allowed/General\_CCL-WG\_docs/CCL-WG-MRA-GD-2.pdf, accessed 2014-07-22
\bibitem{DECKERJ}
Decker J et. al. 2008, \textit{Measurement science and the linking of CIPM and regional key comparisons}, Metrologia \textbf{45}, 223–232
\bibitem{DECKER}
Decker J E, Altschuler J, Beladie H, Malinovsky I, Prieto E, Stoup J, Titov A, Viliesid M and Pekelsky J R 2007, \textit{Report on SIM.L-K1 regional comparison: stage one calibration of gauge blocks by optical interferometry},
Metrologia \textbf{44} (Tech. Suppl.) 04001
\bibitem{GUM}
ISO 1993, \textit{Guide to the Expression of Uncertainty in Measurement} (Geneva: ISO)
\bibitem{WEISE-WOEGER}
Weise K and Wöger W 1992, \textit{A Bayesian theory of measurement uncertainty}, \textit{Meas. Sci. Technol.} \textbf{3}, 1--11
\bibitem{JEFFREYS}
Jeffreys H 1939, \textit{Theory of Probability} (Oxford: Oxford University Press)
\bibitem{JAYNES1}
Jaynes E T 1957, \textit{Information theory and statistical mechanics}, \textit{Phys. Rev.} \textbf{106}, 620--630
\bibitem{JAYNES2}
Jaynes E T 1957, \textit{Information theory and statistical mechanics II}, \textit{Phys. Rev.} \textbf{108}, 171--190
\bibitem{JAYNES3}
Jaynes E T 1968, \textit{Prior probabilities}, \textit{IEEE Trans. On Systems Science and Cybernetics} \textbf{SC-4}(3), 227--241
\bibitem{SHANNON}
Shannon C E 1948, \textit{A mathematical theory of communication}, Bell System Techn. Journ. \textbf{27}, 379--423 and 623--656
\bibitem{KULLBACK}
Kullback S and Leibler R A 1951, \textit{On information and sufficiency}, \textit{Ann. Math. Stat.} \textbf{22}, 79--86
\bibitem{EUROMET}
Acko B 2012, \textit{Final report on EUROMET Key Comparison EUROMET.L-K7: Calibration of line scales}, Metrologia \textbf{49}, 04006
\endbib
\end{document}
|
\section{Introduction}\label{s:intro}
The magnetic field and spin of a white dwarf (WD) are inherited from its progenitor star. Our interest here is
in how magnetism and rotation evolve together within the progenitor, in response to an active
hydromagnetic dynamo. We focus on the post-main sequence phase of stellar evolution, when the material comprising the outer
layers of the WD is deposited by a strong inflow from a hydrogen-rich convective envelope,
through burning shell(s), into the stellar core. The presence or absence of a magnetic field at the
surface of the WD remnant depends on the ability of the inner envelope to sustain a magnetic field.
Rotationally-driven dynamo activity on the red giant (RGB) and asymptotic giant (AGB) branch has received
only limited attention (e.g. \citealt{blackman01, nordhaus08}), due to a prevailing assumption that the radial angular velocity profile
$\Omega(r)$ in the convective envelope will mirror that observed in the Sun. Although localized magnetic field amplification in the outer,
slowly rotating envelope may result in weak surface magnetic fields
\citep{dorch04,auriere10}, a dynamo that manifests itself through strong chromospheric and coronal emission
appears to depend on an interaction with a binary stellar companion, as in RSCvn stars (e.g. \citealt{moss91}).
Magnetic activity appears to be prevalent during the first stages of post-MS expansion, and during core helium
burning \citep{auriere14}, but the sampling of more expanded giants is very limited.
On the other hand, the angular velocity profile in a giant is less sensitive to boundary conditions than it is in the
Sun, given the extreme depth of the envelope.
Here we are guided by i) recent measurements of rapid core rotation in subgiants and core He burning stars
using asteroseismic measurements by Kepler \citep{becketal2012,mossetal2012};
ii) measurements of a strong decrease of surface rotation speeds in subgiants as they evolve toward the RGB \citep{SchrP1993};
and iii) anelastic simulations which show a tendency to nearly uniform specific angular momentum in a deep convective envelope with slow rotation \citep{brunp2009}.
In a companion paper (\citealt{KissT2015}, hereafter paper I),
we show that the Kepler measurements
are consistent with a tight coupling between the radiative core and convective envelope, combined with
some inward pumping of angular momentum deep into the envelope. The favored rotation profile has
$\Omega(r)$ transitioning between $\propto r^{-1}$ in the inner envelope to $\propto r^{-2}$ in the outer and
more slowly rotating layers. The depth of the transition point depends on the size of the star and its angular momentum.
An important consequence of such an inward-peaked rotation profile is that the threshold for a rotationally
driven dynamo may be reached in the inner envelope, but {\it not} near the stellar photosphere.
A key figure of merit is the Coriolis parameter
\begin{equation}\label{eq:coriolis}
{\rm Co} \equiv \Omega \tau_{\rm con} \equiv {\Omega\ell_P\over v_{\rm con}}
\end{equation}
as measured at the base of the envelope; here $v_{\rm con}$ is the convective velocity and $\tau_{\rm con}$ is
the timescale for convection over one pressure scale height. Observations of MS stars indicate the
presence of dynamo activity when ${\rm Co} \gtrsim 1$ (corresponding to a Rossby number $\lesssim 2\pi$;
e.g. \citealt{reinbb2009}). The radial transition in the slope of the rotation profile in a deep convective
envelope is postulated to occur where ${\rm Co}$ approaches unity, with the most rapidly rotating zone lying
below this transition.
Inward pumping of angular momentum can also reduce
the need for magnetically driven spindown in subgiants (Paper I).
\subsection{Remnant Magnetism: Dependence on Stellar Mass, Binarity and Age}
Three hypotheses have dominated previous discussions of the origin of magnetic fields in WD stars:
the conservation of magnetic flux from the main sequence (MS) progenitor \citep{angebl1981,ferrw2005};
the amplification of such a seed field by convective episodes in the core of the progenitor star \citep{RudeS1973};
and the transient amplification of the field during a merger event \citep{ToutWLFP2008,NordWSMB2011}.
We begin our investigation by reviewing some constraints on each of these hypotheses; and
then summarize the arguments pointing to an amplification of the magnetic field
during the post-MS expansion of the WD progenitor. This process comprises
three steps: i) the advection of angular momentum inward through the convective envelope; ii)
the amplification of a magnetic field by a dynamo near the core-envelope boundary; followed by iii)
a flow of magnetic helicity into the growing core, which is sustained by a persistent angular velocity
gradient and magnetic twist below the boundary.
Whether the inner envelope can support a dynamo turns out to depend on a combination of stellar properties. The core rotation rate is
sensitive to the loss of angular momentum on the MS, and to any interaction with a planetary or stellar companion. Clearly solid rotation
would imply ${\rm Co} \ll 1$ throughout the envelope of a star approaching the tip of the RGB or AGB; and therefore suggest the absence of a
large-scale hydromagnetic dynamo.
But when the envelope rotation profile is inwardly peaked, we find differing answers for isolated stars of
initial mass $M_{\rm ZAMS}$ less or greater than $\sim 1.3\,M_\odot$, corresponding to the presence or absence of
a strong magnetic wind on the MS. Hence the greater importance of binary interaction for
sustaining ${\rm Co} \gtrsim 1$ in solar-mass stars.
Another consideration is the proportion of the material contained in the remnant WD that is deposited during single and double shell burning.
This comprises about a third of the mass of the WD when $M_{\rm ZAMS} \sim 1$ ($M_{\rm wd}
\sim 0.55\,M_\odot$), but is strongly reduced by dredge-up (after core helium
ignitition and before the onset of thermal pulses) in the intermediate-mass
stars that are the progenitors of massive WDs. In this second case, the magnetized mass shell turns out
to be thin enough that ohmic decay is initiated on a timescale $\sim 0.3$ Gyr (with some dependence
on the mass loss rate on the AGB).
The source of a strong magnetic field ($1-10^3$ MG) in a WD therefore depends on both its age and mass.
Indeed, expanded sampling of WD magnetism calls into
question the flux-freezing hypothesis.
Whereas between $\sim 2$ and $\sim 10$ percent of isolated
WDs are magnetic in this sense \citep{LiebBH2003,Liebetal2005} -- roughly similar to the proportion of chemically peculiar
and magnetic A stars -- the proportion is higher in isolated massive WDs \citep{kepletal2013} and extends to
$\sim 25$ percent in cataclysmic variables (CVs; \citealt{Gans2005}). A large proportion of magnetic
WDs in short-period CVs also appear to have very strong magnetic fields, exceeding $10$ MG, which are much rarer
in the isolated WD population.
It has also been noted that magnetic WDs appear to be rare in binary
systems with a non-interacting MS companion \citep{Liebetal2005}, suggesting that binary interaction may be involved
in their genesis. Previous theoretical ideas have centered around the impulsive growth of a magnetic field during a binary
merger \citep{NordBF2007,ToutWLFP2008,NordWSMB2011}. Here our focus is on a much longer-lived process that
is sustained by the inward pumping of angular momentum.
A solar-mass star is slowly rotating at the end of the MS, so that the absorption of a giant planet significantly
spins up the star and may trigger dynamo activity \citep{livio02}. The critical planetary mass that has this
effect near the base of the envelope depends on the angular velocity profile. For
a flat distribution of specific angular momentum at ${\rm Co} \lesssim 1$, we find that the absorption of even
a Neptune-mass planet can significantly spin up the core.
A star more massive than $M_{\rm ZAMS} \sim 1.3\,M_\odot$ retains enough angular
momentum at the end of the MS that only a planet more massive than Jupiter will significantly supplement its rotational angular momentum.
Nonetheless we find that such an interaction can lead to stronger magnetic fields in the WD remnant.
Two properties of giants are especially relevant here. The inward flow of matter across the core-envelope
boundary allows a net transport of magnetic helicity, which has been shown to be a necessary ingredient in
the long-term stabilization of white dwarf magnetic fields \citep{brais2004}; and has also been implicated in the establishment of
solid rotation in the radiative core of the Sun \citep{GougM1998}. We show that this helicity flux can be related to
the magnitude of the large-scale Maxwell torque that enforces (nearly) solid rotation in the outer core.
The outer core of an AGB star is also distinguished by an intense radiation flux.
This greatly facilitates the buoyancy of a toroidal magnetic field
that is anchored in a tachocline layer below the convective envelope.
A thickness approaching $\sim 10$ percent of a scale height can maintain
contact with the envelope over a dynamo period $\sim 10-10^2 \tau_{\rm con}$. This suggests that a given level
of magnetic activity (measured e.g. by the ratio of Maxwell to Reynolds stresses) may be achieved with a lower
Coriolis parameter in an AGB envelope than in a solar-type MS star. Magnetically induced mixing of core
material into the envelope \citep{nordhaus08} could also be significantly enhanced by this effect.
\subsection{Remnant Spin}
The measured spins of WDs are strongly inconsistent with solid rotation in the envelope of an AGB star, combined with
a tight core-envelope coupling. But even if we allow for inward convective pumping of angular momentum, the final
spin rate of a WD remnant is sensitive to the heavy loss of mass during the transition to the
post-AGB phase.\footnote{We thank Peter Goldreich for emphasizing this point to us.}
A decoupling of the rotation of core and envelope is only feasible
when the large-scale poloidal magnetic field drops below the equivalent of $10^2-10^3$ G in the WD during
the brief, thermally pulsating AGB phase. (Much weaker fields will couple the core and envelope of a subgiant
or core helium burning star; Paper I.) This
appears, at first sight, to be inconsistent with strong magnetism in the WD.
In fact, a magnetic coupling between core and envelope is naturally self-limiting: once ${\rm Co}$ drops
below a critical value, the dynamo operating in the inner envelope shuts off.
A reasonable first guess (${\rm Co}_{\rm crit} \sim 0.1-0.3$) yields WD rotation periods in the range observed
in most WDs.
As a result, the remnant rotation is {\it inversely} correlated with the net angular momentum
in the star: the slower the rotation of the AGB envelope, the sooner the decoupling between core and envelope,
when the core has a larger radius and moment of inertia. The weakly magnetized
surface layer of the WD is very thin, and we show that the magnetic field below quickly diffuses to the photosphere.
Our most surprising finding is that WDs with very long ($\sim$ yr) rotation
periods originate from stars which spin relatively rapidly
near the end of the AGB phase. Then the rotation of core and envelope
remain magnetically coupled during the contraction of the envelope.
The endpoint is sensitive to details of how the envelope is expelled: in particular, to
whether the envelope experiences a re-expansion following a late thermal
pulse. This taps most of the remaining angular momentum of the core.
Without such a re-expansion, the final rotation period can be shorter than a day.
It should be kept in mind that anisotropic contraction of the envelope can contribute to white dwarf rotation
\citep{Spru1998}. Indeed, it may be needed to generate centrifugally supported disks that provide seeds for
metal enrichment of white dwarf atmospheres.
\subsection{Stellar Models and Planetary Interaction}
In order to evaluate various pieces of the physics, we employ two stellar models of solar metallicity and initial mass
$M_{\rm ZAMS} = 1\,M_\odot$ and $5\,M_\odot$. Both models are constructed using the MESA code \citep[version 5527]{paxtetal2011}, with
rotational degrees of freedom turned off. The normalization
of the mass-loss rate on the AGB, and the masses of the WD remants, were fixed by matching the latest initial-to-final
mass relation obtained from open clusters \citep{Dobbetal2009}. The coefficient entering into the formula of
\cite{bloe1995} was taken to be $\eta_B = 0.05$. The corresponding masses of the WD remnants are $0.55\,M_\odot$ and
$0.87\,M_\odot$.
The rotational evolution was followed in post-processing using the rotation models
described in Section \ref{s:rotation}. The loss of angular momentum due to mass loss on the RGB and AGB was
taken into account, along with the interaction with a companion planet of a range of masses: Earth, Neptune, Jupiter
in the $1\,M_\odot$ model, and even heavier companions in the $5\,M_\odot$ model. In the first case, the companions were
started at a range of semi-major axes and orbital eccentricities. The orbit of the planet,
and the exchange of angular momentum with the star, were computed taking into account tidal drag
and the gravitational quadrupole that is generated by convection in the giant envelope.
Further details are given in Paper I.
\subsection{Plan of the Paper}
Section \ref{s:basic} and \ref{s:WD_magnetism} discuss several processes that are relevant to the growth and maintenance of a stable magnetic
field in the core of a post-MS star, and which determine the spin angular momentum trapped there at the end of the AGB.
Section \ref{s:finalB} presents detailed estimates of WD dipole magnetic fields.
Section \ref{s:WD_rotation} analyzes the magnetic decoupling of core and envelope near the tip of the AGB, and provides
estimates of the final WD spin. Ohmic transport through a thin, unmagnetized surface layer in a WD, and
the thicker magnetized layer below it, is considered in Section \ref{s:B_field_emergence_decay}.
Various outstanding issues -- such as the sensitivity of
our results to the assumed angular velocity profile in the progenitor envelope, the impulsive growth of a magnetic field during a merger,
and effects associated with the anisotropic contraction of the residual AGB envelope -- are addressed in the
concluding Section \ref{s:conclusions}. The Appendix gives further details of the accumulation of magnetic helicity in the stellar core.
\section{Rotation and Magnetism in a Rapidly Evolving Giant Star}\label{s:basic}
Here we summarize the rotation model for giant stars that was developed in Paper I. We then
review some physical constraints on generating sustainable white dwarf magnetic fields, with a focus on
the accumulation of magnetic helicity during the rapid growth of the hydrogen-depleted
core of a giant star. We also summarize
the relative effectiveness of different convective episodes at sourcing WD magnetic fields.
In this regard, there are some significant structural differences between the evolving cores of solar-mass and of intermediate-mass stars.
\subsection{Combined Effect of Convection and Poloidal Magnetic Field on Angular Momentum Redistribution}\label{s:rotation}
Our understanding of the origin of WD rotation has been impeded by an incomplete knowledge of mixing processes
in the radiative parts of the progenitor. Even a very weak poloidal magnetic field (corresponding to a magnetic
field $\ll$ MG in the WD under the assumption of magnetic flux freezing) will enforce solid rotation in the evolving
core.
A popular approach recently has been to assume that such a large-scale poloidal field is initially
absent, and to imagine that the radial magnetic field that is needed for angular momentum transport is generated by
a current-driven (`Tayler') instability of a predominantly toroidal field (\citealt{Spru2002,CantMBCP2014}, and references therein).
This leads to an incomplete coupling between core and envelope. We argue that even an inefficient dynamo
process operating near the core-envelope boundary will easily short-circuit this effect.
We develop, as a working alternative, a simple and deterministic model of angular momentum redistribution
in evolving stars. This model applies to stars which retain (or gain) enough angular momentum to support
some dynamo activity near the core-envelope boundary.
\vskip .1in \noindent
1. The outer core is assumed to rotate as a solid body,
as enforced by large-scale Maxwell stresses, and to co-rotate on the average with the inner envelope.
\vskip .1in \noindent
2. Inhomogeneous rotation is sustained mainly in convective regions of the star, especially the deep outer
envelope that forms following the completion of core H and He burning.
\vskip .1in
The rotation profile that is formulated and calibrated in paper I can be summarized as follows.
Where the Coriolis force can
be neglected, outside some transition radius $R_c$, the envelope maintains uniform specific angular momentum,
\begin{equation}\label{eq:omouter}
\bar\Omega(r) = \bar\Omega(R_c) \left({r\over R_c}\right)^{-2}; \quad r > R_c.
\end{equation}
Here $\bar\Omega$ is the angular velocity averaged over a spherical shell. We identify
$R_c$ with the base of the envelope,
\begin{equation}
R_c = R_{\rm benv} \quad (\bar{\Omega} \tau_{\rm con} < 1\;{\rm at} \; R_{\rm benv}),
\end{equation}
when the angular momentum of the star is small enough that $\bar\Omega(r) \tau_{\rm con} < 1$ throughout the envelope.
Throughout this paper, the subscript `benv' refers to the base of the envelope.
When the star has more angular momentum, we maintain the profile (\ref{eq:omouter}) in the
outer part of the convection zone, and consider
\begin{eqnarray}\label{eq:ominner}
\bar\Omega(r) &=& \bar\Omega(R_c) \left({r\over R_c}\right)^{-\alpha} \mbox{} \nonumber \\ \mbox{}
&=& {{\rm Co}_{\rm trans}\over \tau_{\rm con}(R_c)} \left({r\over R_c}\right)^{-\alpha}; \quad
R_{\rm benv} < r < R_c
\end{eqnarray}
in the inner envelope. The Coriolis parameter at the transition radius ${\rm Co}_{\rm trans}$ is taken to
be unity, and the radial index $1 < \alpha < 3/2$. As the angular momentum increases $R_c$ moves outward,
and may reach the surface of the star, in which case ${\rm Co}$ is larger than unity everywhere in the envelope.
As we discuss in Paper I, $\alpha=1$ best matches the observed core rotation periods of sub-giant and helium burning stars,
which we use to calibrate our rotational model. However, in stars of much greater size, such as those approaching the tips of
the RGB and AGB, the core is more centrally concentrated and the gravitational field in the inner
envelope, $g(r) \propto r^{-\beta}$, steepens from $\beta = 1$ to $\beta = 2$. Then the value of $\alpha$ that results from the inward transport of
angular momentum by deep convective plumes increases from $(1+\beta)/2 = 1$ to $3/2$. It is during this more advanced phase of
evolution that the helicity flux into the core is expected to peak, and the rotational fate of the WD is determined.
We therefore use $\alpha=3/2$ in our analysis.
\subsection{Magnetic Helicity Growth in Giant Cores}\label{s:htransport}
Net magnetic helicity ${\cal H} = \int{\bf A}\cdot{\bf B}\,dV$ appears to be an essential ingredient for maintaining long-term
magnetostatic stability in a static and entirely fluid star \citep{brais2004}. Here ${\bf B}$ is the magnetic field
and ${\bf A}$ the vector potential from which it is derived. ${\cal H}$ is a topological charge in a perfectly conducting
fluid without boundary. For example, when a closed poloidal magnetic loop carrying flux $\Phi_p$ is also twisted through an angle
$\psi$, the helicity is ${\cal H} = \Phi_p\Phi_\phi = \psi\Phi_p^2$, where $\Phi_\phi$ is the toroidal flux. ${\cal H}$ evolves
only on a very long ohmic timescale in an extended astrophysical fluid -- unless there is an exchange of magnetic twist with
the exterior of the fluid, or with a medium of low electrical conductivity.
The accumulation of helicity considered here is associated with rapid changes in the structure of the star.
We expect this to be possible in the outer core and inner envelope of a giant star,
since this zone is connected to the surface of the star by rapid convective motions. Core convection has a topology
less favorable to the accumulation of ${\cal H}$, as does the erasure of differential rotation by the
stretching of magnetic field lines in the interior of a radiative star. A simplified time-dependent model of
an internal shear-driven dynamo has been developed by \cite{WickTF2014}, but does not address the growth of ${\cal H}$.
Such a secular accumulation of magnetic helicity -- although occuring relatively rapidly near the tips of
the RGB and AGB -- is nonetheless a gradual process when measured over an individual dynamo cycle. There has also
been longstanding interest in the role that helicity conservation may play in limiting the exponential growth of
a seed magnetic field in a turbulent magnetofluid \citep{gruzinov94,bf00a,vishniac01}. In mean-field dynamo models, this involves
a suppression of the $\alpha$ effect, e.g. the toroidal electromotive force which sources a large-scale poloidal magnetic field.
This limitation on fast dynamo action could be removed by the transport of
magnetic helicity on a relatively short timescale, comparable to the rotation period. Such transport may involve
the removal of twist across the surface of the star \citep{bf00b}, or alternatively an internal flow of
helicity between rotational hemispheres \citep{vishniac14}. In either of these
cases, rapid dynamo growth is not directly dependent on changes in the net stellar helicity, obtained by summing the contribution from both
hemispheres. The longer, secular effect being considered here necessarily
depends on a breakdown of reflection symmetry between magnetic hemispheres; and, because it involves a change in an integral property of the
star, must depend on the transport of twist across the stellar surface.
The connection between helicity growth and mass inflow can be examined with a simple model of a toroidal
magnetic twist superposed on a radial magnetic field, all subjected to a radial MHD flow of speed $v_r$.
(The sign of this flow, relative to the convective boundary, fluctuates during thermal pulses on the AGB
but is negative on the average.)
Within a sphere coinciding with $R_{\rm benv}$, one finds
\begin{eqnarray}\label{eq:dhdt}
{d{\cal H}\over dt} &=& {d\over dt}\int {\bf A}\cdot{\bf B}\, dV \mbox{} \nonumber \\ \mbox{}
&=&
{\partial{\cal H}\over \partial t} + {dR_{\rm benv} \over dt}\int dS A_\phi B_\phi
\;=\; -\delta v_r\int dS\,A_\phi B_\phi\mbox{} \nonumber \\ \mbox{}
\end{eqnarray}
(Appendix \ref{s:twist}). The gauge
\begin{equation}
A_\phi(\theta,r) = {1\over r\sin\theta}\Phi_r(\theta)
\end{equation}
is convenient, where
\begin{equation} \label{eq:flux}
\Phi_r(\theta) = \int_0^\theta \sin\theta' r^2B_r(\theta') d\theta'
\end{equation}
is $(2\pi)^{-1}$ times the radial magnetic flux between the rotation axis and polar angle $\theta$.
Clearly $\Phi_r(\pi) = 0$.
One notices that the surface integral appearing in Equation (\ref{eq:dhdt}) has the dimensions of a torque:
it is comparable in magnitude to the Maxwell stress $B_r B_\phi/4\pi$, multiplied by the lever arm
$r\sin\theta$, and integrated over the core-envelope interface. Even if the net torque acting on
the core boundary vanishes, $d{\cal H}/dt$ may remain finite as a consequence of the different
angular dependences of $A_\phi$ and $B_r r\sin\theta$.
We are interested here in the case where $B_r B_\phi/4\pi$ is the dominant stress driving the core
material toward solid rotation. Then we can estimate the helicity flux into the core once we know
the source of the Maxwell torque. The largest effect turns out to be the latitude dependence of
the Reynolds stress that is applied to the core at the base of the convective envelope.
We may estimate this stress by considering a dynamo process at work in a tachocline layer
right below the convective boundary.
Convection above the tachocline supports strong latitudinal variations in $\Omega$ which impose a radial mis-match
$\Delta\Omega$ with the mean core rotation $\Omega_{\rm benv} = \bar\Omega(R_{\rm benv})$.
The tachocline can therefore be identified with the radial layer (of thickness $\Delta r$)
in which $\Delta\Omega$ is concentrated,
\begin{equation}
\mbox{\boldmath$\nabla$}\Omega \simeq {\partial\Omega\over \partial r}\hat r
\sim {\Delta\Omega\over \Delta r} \hat r.
\end{equation}
Our focus here is on the parts of the tachocline where $\partial\Omega/\partial r > 0$
(in the Sun, it is the equatorial band at latitudes $\lesssim 45^\circ$).
The magnetorotational instability \citep{BalbH1994} is excited where $\partial\Omega/\partial r < 0$,
and is sustained by rapid radiative diffusion even in the presence of stable radiative convective
equilibrium \citep{menou04}.
The poloidal magnetic field threading this part of the tachocline is wound up linearly on
timescales short compared with the dynamo period $P_{\rm dyn}$, which we normalize as
$N_{\rm dyn}\Delta\Omega^{-1}$. We limit the magnetic field strength in the tachocline
by noting that a radial magnetic field stronger than
\begin{equation}
B_r \sim (4\pi \rho)^{1/2} {\Delta r\over P_{\rm dyn}}
\end{equation}
will act to redistribute angular momentum in a layer of density $\rho$ and thickness $\Delta r$, thereby
shorting out the angular velocity gradient.
The ratio of toroidal to poloidal field components is obtained from the linear winding term in the
induction equation,
\begin{equation}\label{eq:bphir}
{2\pi\over P_{\rm dyn}} B_\phi \sim {r \Delta\Omega \over \Delta r} B_r.
\end{equation}
One finds that the toroidal magnetic field is in approximate equipartition with
the rotational motions at the base of the convective envelope,
\begin{equation}\label{eq:bphieq}
B_\phi^2 \sim 4\pi \rho \left(r\frac{\Delta\Omega}{2\pi}\right)^2 \sim
4\pi \left({{\rm Co}\over 2\pi}{r\over\ell_P}\frac{\Delta\Omega}{\bar{\Omega}}\right)^2\rho v_{\rm con}^2 .
\end{equation}
The Maxwell stress works out to
\begin{equation}\label{eq:reyn}
{B_r B_\phi\over 4\pi} \sim {\rho r \Delta r (\Delta\Omega)^2 \over 2\pi N_{\rm dyn}}.
\end{equation}
We emphasize that this estimate of the stress has been obtained from dynamical considerations,
and side-steps the challenging question of how the $\alpha$ effect is actually manifested in a
dynamo operating near a radiative-convective boundary. For recent approaches to this problem,
in the context of lower-luminosity MS stars, see \cite{bt15} and references therein.
We note that $\Delta\Omega \sim \bar\Omega$ is seen in anelastic calculations of deep and rapidly rotating
(${\rm Co} > 1$) convective envelopes \citep{brunp2009}, and is suggested by the analytic treatment of
angular momentum pumping in Paper I.
The latitudinal gradient is milder in the Sun, where the equator rotates more rapidly than the poles,
$(\Omega_{\rm eq}-\Omega_{\rm pole})/\bar{\Omega} \sim 0.1$; but the Solar convective envelope also has a very
different aspect ratio.
Now consider the helicity flux integrated over one rotational hemisphere. Even in a state of rotational
equilibrium -- where the angle-integrated Maxwell torque vanishes -- the integral (\ref{eq:dhdt}) generally
remains finite. Notice also that it is the product
of $B_r$ and $B_\phi$ that is regulated according to Equation (\ref{eq:reyn}). The Maxwell
stress does not depend on the sign of $B_r$ that is imposed by a dynamo operating near the core-envelope
boundary, nor does it depend directly on the coherence of the magnetic field.
The sum of the two hemispheric integrals for $d{\cal H}/dt$ would vanish if the magnetic field were
reflection-symmetric about the rotational equator. But in general it is not: the distribution of
radial magnetic flux across the northern hemisphere need not precisely mirror that threading the
southern hemisphere. (Only the net spherical flux vanishes.) Such an incomplete hemispheric
cancellation allows a net flow of magnetic twist into the core.
Other sources of helicity flow can be considered. The inflow through the core-envelope boundary
itself drives $\partial\Omega/\partial r$, but on a much longer timescale than the rotation period.
The WD magnetic fields that result from these two sources of
${\cal H}$ are compared in Section \ref{s:WD_magnetism}.
\subsection{Relative Importance of \\ Different Convective Episodes}
The magnetic field of a WD is influenced by a series of convective episodes in the progenitor. One measure
of the relative importance of different convective episodes is provided by the convective Mach number
${\cal M}_{\rm con} = v_{\rm con}/c_s$ (here $c_s$ is the adiabatic speed of sound), which in a stellar core can be
estimated as
\begin{equation}
{\cal M}_{\rm con} \sim \left({R_{\rm core} v_{\rm con}^2 \over GM_{\rm core}}\right)^{1/2}.
\end{equation}
As the core continues to contract during later stages of nuclear burning, its magnetic energy and gravitational
energy scale in the same way with central density. To the extent that the generated Maxwell stress is limited by the convective stress, e.g.
$B_\phi \lesssim (4\pi \rho)^{1/2} v_{\rm con}$, episodes of greater
${\cal M}_{\rm con}$ are capable of generating stronger magnetic fields.
One encounters much larger ${\cal M}_{\rm con}$ on the giant branches ($\sim 10^{-3}-10^{-2}$) than during steady core H or He
burning in intermediate-mass stars ($\sim 10^{-5}-10^{-4}$).
It is also worth noting that the intense period of convection during the AGB phase is also the shortest in duration. That implies
a weaker cancellation in the magnetic helicity that accumulates in the core, due to a shifting distribution of poloidal magnetic flux
(see Section \ref{s:htransport}).
Magnetic fields that are generated during a certain convective phase will remain trapped unless exposed to later
stages of convection, or the exterior mass shells are expelled. As we examine in Section \ref{s:B_field_emergence_decay},
ohmic diffusion in the WD remnants of intermediate-mass stars is restricted to a relatively thin magnetized layer.
The relevance of dynamo activity on the RGB, horizontal branch and AGB is determined, in good part,
by the maximum inward penetration of the convective envelope: one requires $M(R_{\rm benv}) \leq M_{\rm wd}$.
Only in relatively massive progenitors ($M_{\rm ZAMS} \gtrsim 4\,M_\odot$) does MS core convection
extend to the mass boundary of the WD remnant, where it can influence the visible magnetic field; we ignore this effect here.
The core helium flash occuring in stars of mass $< 2.3\,M_\odot$ is confined to a central subset of the eventual WD material.
Because a star does not become fully convective during the flash (e.g. \citealt{bildsten12}),
the magnetic helicity of the convecting core material is essentially conserved. A similar conclusion applies to the brief
convective layers that are triggered by thermal pulses in the helium-burning shell near the tip of the AGB.
\subsubsection{Stars with Minimal Dredge-up of Helium ($M_{\rm ZAMS} \lesssim 2.3\,M_\odot$)}
In stars of initial mass $M_{\rm ZAMS} \sim 1\,M_\odot$, magnetic fields generated on the RGB are buried inside the WD remnant
under a layer of CO-rich material that is produced during double shell burning (Figure \ref{fig:M_star_M_base_M_He_core_M_CO_core_TRGB_solar}).
They are ohmically coupled on a timescale $\sim 10^8-10^9$ yr to a thinner magnetized shell that appears on the
AGB (Section \ref{s:B_field_emergence_decay}).
The ohmic timescale across the entire magnetized shell is longer than $\sim 10^9$ yr, meaning that any magnetic field
deposited near the center of the star during the pre-MS phase would influence the surface of the star only on a
timescale of a few Gyr.
We conclude that the existence of young magnetized WDs with masses $\lesssim 0.6\,M_\odot$ points to dynamo activity in the convective envelope on the RGB and AGB.
\begin{figure}[!]
\figurenum{1}
\epsscale{1.0}
\plotone{f1a}
\vskip 0.2in
\plotone{f1b}
\caption{Interior profile of $M_{\rm ZAMS} = 1 M_\odot$ star around the tip of the RGB and first stages of core
He burning (top panel) and during AGB phase (bottom panel).
Dotted magenta line: maximum penetration of the convective envelope on the AGB ($0.525M_{\odot}$).
A dynamo operating on the RGB deposits magnetized material within mass shells below this line; above it, the magnetic
field is sourced on the AGB. Reference time $t_{10.9} \sim 12.37$ Gyr subtracts MS and early giant branch evolution
($R_{\star}(t_{10.9}) \sim 10.9R_{\odot}$).}
\vskip .2in
\label{fig:M_star_M_base_M_He_core_M_CO_core_TRGB_solar}
\end{figure}
\begin{figure}[!]
\figurenum{2}
\epsscale{1.0}
\plotone{f2a}
\vskip 0.2in
\plotone{f2b}
\caption{Interior profile of a $M_{\rm ZAMS} = 5M_{\odot}$ star during RGB expansion and core He burning phase (top panel) and AGB phase
(bottom panel). The deepest penetration of the convective envelope on the RGB remains outside the mass boundary of the WD remnant
($\sim 0.87M_{\odot}$). Dredge-up of helium occurs on the early AGB; the dotted magenta
line shows the corresponding maximum penetration of the convection.}
\vskip .2in
\label{fig:M_star_M_base_M_He_core_M_CO_core_TRGB_5M_o}
\end{figure}
\subsubsection{Intermediate-mass Stars ($M_{\rm ZAMS} \gtrsim 2.3\,M_\odot$)}\label{s:inter}
Stars of intermediate mass are distinguished from those of solar mass in that a dynamo operating
on the RGB cannot contribute directly to the WD magnetic field: the outer parts of the helium core
are dispersed by dredge up before the C/O core is assembled. In addition, only a limited mass is
converted to carbon and oxygen by double shell burning (Figure \ref{fig:M_star_M_base_M_He_core_M_CO_core_TRGB_5M_o}).
The fraction of the WD mass contained by this outer shell is sensitive to the rate of mass loss on the
thermally pulsating AGB.
The ohmic conduction time across this outer magnetized shell is several $10^8$ years.
The surface magnetic field of the remnant of an isolated, intermediate-mass star is therefore expected to grow and
then gradually decay on this timescale (Section \ref{s:B_field_emergence_decay}).
\section{Magnetic Field Amplification Near Core-Envelope Boundary} \label{s:WD_magnetism}
We now investigate the growth of a magnetic field above and below the core-envelope boundary of a post-MS star.
There are two profound differences here with a solar-type star. First, the radiative flux that is generated
by the burning shell(s) much exceeds the flux emerging from the core on the MS. Second, hydrogen-rich material
drifts downward rapidly through the convective boundary, as the core grows in mass. This leads to a secular accumulation of magnetic
helicity in the core, as this inward drift is combined with a persistent magnetic twist. Such a twist is generated when the magnetic field
couples the rotation of core and envelope. In these respects, our considerations extend beyond existing dynamo theory.
\subsection{Critical {\rm Co} for Dynamo Action: \\ Effect of an Intense Radiation Flux in the Tachocline}\label{s:dynthresh}
Late-type MS stars generally are compact enough to sustain ${\rm Co} \gtrsim 1$ in their convective envelopes, as
are subgiants massive enough to have retained most of their natal angular momentum at the completion of core H burning.
Dynamo activity, as measured by chromospheric lines and coronal X-ray emission,
begins to decline when the Rossby number ${\rm Ro} \equiv P_{\rm rot}/\tau_{\rm con} = 2\pi/{\rm Co}$ is larger than about 0.1,
corresponding to ${\rm Co} \lesssim 60$. Specifically in M-dwarfs, the closest MS analogs of giants, \cite{reinbb2009} observe
that surface magnetic flux scales as ${\rm Co}^2$ for ${\rm Co} \lesssim 60$.
There does not appear to be a sharp cutoff of magnetic activity with Coriolis parameter, as the simplest mean-field dynamo models would
suggest.
The evidence therefore points to the presence of weak, but finite, dynamo activity at ${\rm Co} \sim 1$ in MS stars.
It also begs the question as to whether the relative strength of the dynamo-generated magnetic field
(measured in terms of the convective stress at a given Coriolis parameter) differs in giants and MS stars.
First it should be emphasized that a large-scale Maxwell torque that couples the rotation of the outer core to the
envelope depends on a tiny poloidal field that is implanted from the dynamo layer. We quantify this minimal
field in Section \ref{s:bseed}.
Some additional insight is provided by mean-field dynamo theory, which suggests that the dynamo number
\begin{equation}\label{eq:dynum}
D \equiv {\alpha \Omega r^3\over \nu_t^2} \sim \left({\alpha\over v_{\rm con}}\right) \left({r\over\ell_P}\right)^3
{\rm Co}
\end{equation}
must exceed a threshold $D \gtrsim (2\pi)^3 \sim 10^2$ for magnetic field growth to be sustained \citep{Char2014}. Here $\alpha$ is the
parameter relating the toroidal magnetic field to the toroidal electromotive force, and $\nu_t$ is the turbulent
diffusivity. The prospect for a dynamo is enhanced if the magnetic field can be anchored in a part of the star with
relatively low (but finite) $\nu_t$.
This effect provides part of the motivation for considering a thin, stably-stratified layer below the convective
envelope as the locus of strong differential rotation, and a reservoir for a toroidal magnetic field
\citep{park1993}. However, the sharpness of the onset of stable stratification below the envelope boundary has presented
a difficulty for this hypothesis. The solar convective motions have a relatively low Mach number
(${\cal M}_{\rm con} \sim 10^{-4}$), and so the overshoot layer below the envelope is very thin. A toroidal magnetic
field reaching equipartition with the convective stresses overcomes the stable stratification in a layer that is
not much thicker.
In Section \ref{s:buoyrad}, we show that the intense flux of radiation emanating from the core of a RGB or AGB star
will induce an upward drift of magnetized material, one that more than counterbalances the downward
flow to the burning shell(s). A tiny seed field that is generated by the convective motions will sustain this feedback
(Section \ref{s:bseed}). The interchange between toroidal and poloidal fields will persist in a layer of considerable thickness
$\sim (0.01-0.1)\ell_P$ over a dynamo cycle $\sim (10-10^2)\tau_{\rm con}$. Convectively driven turbulence is strongly suppressed within
this layer.
\begin{figure}[!]
\figurenum{3}
\figurenum{3}
\epsscale{1.0}
\plotone{f3.eps}
\caption{A magnetic flux tube positioned in the tachocline layer experiences an upward drift that is driven by the intense radiation
flux through the tube. This drift can be strong enough to counterbalance the downward flow of hydrogen-rich material
into the burning shell(s) below.}
\vskip .2in
\label{fig:fluxtube}
\end{figure}
\subsection{Magnetic Buoyancy in the Tachocline}\label{s:buoyrad}
\begin{figure}[!]
\figurenum{4}
\epsscale{1.0}
\plotone{f4}
\caption{Horizontal (toroidal) magnetic field anchored in a tachocline experiences radiatively driven buoyancy.
Upward drift speed $v_r^B$, Equation (\ref{eq:buoy}), is compared with the downward flow of material from the
convective envelope into the growing radiative core. Curves correspond to
various timesteps near the tips of the RGB and AGB in our $1\,M_\odot$ model, and one taken from the thermally pulsating AGB. The
magnetic field is normalized to $B_\phi \sim (4\pi \rho v_{\rm con}^2)^{1/2}$, with $v_r^B$ proportional to $B_\phi^2$.
Buoyant drift accelerates toward the core-envelope boundary, due to the decrease in the Brunt-V\"ais\"al\"a frequency.
Here downflow is evaluated directly from the stellar model.}
\vskip .2in
\label{fig:vbuoy}
\end{figure}
We now examine the competition between the downward drift of hydrogen-rich material across the tachocline, and the upward
drift of magnetized material (Figure \ref{fig:fluxtube}). Consider a horizontal (toroidal) magnetic flux tube, of a radial
thickness comparable to the distance $\Delta r = R_{\rm benv}-r$ to the base of the convective envelope. The flux tube
starts in buoyancy equilibrium, with a density equal to that of its unmagnetized surroundings. The temperature deficit
in the flux tube (labelled `$B$') is
\begin{equation}
{T_B - T\over T} = -{B_\phi^2\over 8\pi P} \ ,
\end{equation}
which maintains a differential energy flux
\begin{equation}
\delta F_{\rm rad} \sim -{4ac\over 3\kappa\rho\Delta r} (T_B-T)T^3
\end{equation}
into the flux tube. This drives a buoyant rise of the flux tube at a speed $v_r^B$, which is determined by
\begin{equation}
\rho k_{\rm B}T v_r^B {dS\over dr} = -{dF_{\rm rad}\over dr} \sim {\delta F_{\rm rad}\over \Delta r}.
\end{equation}
The background entropy gradient is usefully expressed in terms of the Brunt-V\"ais\"al\"a frequency
\begin{equation}
N = \sqrt{g {\gamma-1\over \gamma} {dS\over dr}},
\end{equation}
so that
\begin{equation}\label{eq:buoy}
v_r^B \sim \left({B_\phi^2\over 8\pi P}\right) {F_{\rm rad}\over (\Delta r)^2 \rho N^2}.
\end{equation}
The background flow of mass into the radiative core is
\begin{equation}
{dM_{\rm core}\over dt} = 4\pi R_{\rm benv}^2 \rho(R_{\rm benv})\left({dR_{\rm core}\over dt} - v_r^{\rm nuc}\right) \sim
{L\over\varepsilon_i}.
\end{equation}
After the star passes onto the thermally pulsating AGB, the relation between core growth rate and luminosity $L$ only holds after averaging
over pulsations. The nuclear energy released per unit mass of new core material is $\varepsilon_1 = 6.3 X$ MeV/$m_u$
during the first step of converting material with hydrogen mass fraction $X$ to helium; and $\varepsilon_2 = \varepsilon_1 + 0.8$ MeV/$m_u$
during double shell burning. The corresponding velocity is
\begin{equation}
v_r^{\rm nuc} \sim -{F_{\rm rad}\over\rho \varepsilon_i}.
\end{equation}
The ratio $v_r^B/|v_r^{\rm nuc}|$ is plotted in Figure \ref{fig:vbuoy} at various intervals during the RGB and AGB
evolution of our $1\,M_\odot$ model.
One observes that the most strongly magnetized parts of the tachocline will
not be subducted if the magnetic pressure approaches the kinetic pressure at the base of the convective envelope.
It should be kept in mind that this effect depends on a difference in magnetization: subduction of the less weakly
magnetized parts of the tachocline must continue even where $v_r^B > |v_r^{\rm nuc}|$.
Magnetic buoyancy has the effect of expanding the active dynamo region below $R_{\rm benv}$, by allowing a layer
with strong $d\Omega/dr$ -- the tachocline -- to remain in contact with the convection zone.\footnote{We do not address here whether the
re-conversion of wound toroidal magnetic field back to the poloidal direction is concentrated in the tachocline, or instead
the convection zone, as in Parker's 1993 model.}
We can estimate the size of this region by comparing the buoyant rise time $\Delta r/v_r^B(r)$ with the dynamo timescale $P_{\rm dyn}$
(set equal to $100\tau_{\rm con}$)
\begin{equation} \label{eq:Delta_r_buoyancy}
\Delta r \equiv R_{\rm benv}-r = v_r^B(r) \cdot P_{\rm dyn}.
\end{equation}
To invert this expression, we fit $v_r^B(\Delta r)$ to a power law in $\Delta r$, measuring it in two different radial zones.
Figure \ref{fig:delta_r_TP_AGB} shows the resulting buoyant shell thickness, in comparison with the local scale height,
during the AGB phase of the $5\,M_\odot$ model.
The buoyant shell thickness reaches $\sim 0.1 l_P(R_{\rm benv})$ before and during the first thermal pulse, and then relaxes to a value
about 10 times smaller.
\begin{figure}[!]
\figurenum{5}
\plotone{f5}
\epsscale{1.0}
\caption{Thickness $\Delta r$ of magnetized shell below the core-envelope boundary which remains in contact with the inner convection zone
over the dynamo period $P_{\rm dyn}$. See Equation (\ref{eq:Delta_r_buoyancy}). Here $\Delta r$ is normalized to the local pressure
scaleheight. Plot covers the thermally pulsating AGB phase of the $5\,M_\odot$ model, with $B_\phi = (4\pi \rho v_{\rm con}^2)^{1/2}$.}
\vskip .2in
\label{fig:delta_r_TP_AGB}
\end{figure}
\subsection{Seeding a Toroidal Magnetic Field \\ in the Tachocline}\label{s:bseed}
We now consider the minimal seed field that will trigger dynamo feedback in the tachocline, thereby facilitating the magnetic coupling
of core and envelope. We suppose that this seed field is generated within the convective envelope, and is deposited in the
outer radiative core by the mass flow toward the burning shell(s).
This seed field must pass a certain threshold if linear winding in the tachocline is to generate a buoyantly unstable toroidal field.
The time available for linear winding depends on the depth $\Delta r$ below the radiative-convective boundary. Consider
an unmagnetized core in which matter flowing across the convective boundary sustains an angular velocity profile
$\Omega = \Omega_{\rm benv}(r/R_{\rm benv})^{-2}$. Near the boundary, the toroidal magnetic field that is sourced
by a seed radial field $B_{\rm seed}$ is $B_\phi = -2B_{\rm seed}\Omega_{\rm benv}\Delta r/v_r^{\rm nuc}$.
Substituting this into Equation (\ref{eq:buoy}) and setting $v_r^B > |v_r^{\rm nuc}|$ gives
\begin{equation}\label{eq:bseed}
{B_{\rm seed}^2\over 4\pi \rho(R_{\rm benv}) v_{\rm con}^2} > { N^2 \over 2 \Omega^2 } { P |v_r^{nuc}|^3 \over v_{con}^2 F_{rad} } \ .
\end{equation}
This minimal seed field for dynamo feedback turns out to be quite weak. Figure \ref{fig:B_seed_TAGB} shows
${B_{\rm seed} / \sqrt{4\pi \rho(R_{\rm benv})} v_{\rm con}}$ during the final AGB evolution of a $5\,M_\odot$ star, measured at a
distance $\Delta r = l_P(R_{\rm benv})/10$ below the core-envelope boundary. Much of the C/O mass deposition from double shell burning
is concentrated around $t \sim 1.0716\times10^8$ yr, when this quantity is as small as $\sim 10^{-9}$; it rises to $\sim 10^{-4}$
during the phase of rapid thermal pulses.
\begin{figure}[!]
\figurenum{6}
\plotone{f6}
\vskip 0in
\epsscale{1.0}
\caption{Minimal seed poloidal magnetic field, deposited in the outer core of an AGB star, which will grow by linear winding to a sufficient
strength that radiatively driven magnetic buoyancy overcomes the downward drift of envelope material to the burning shell(s).
Equation (\ref{eq:bseed}) is evaluated at depth $\Delta r = l_{P,\rm benv}/10$ below core-envelope boundary.
Plot covers same thermally pulsating AGB phase of $5\,M_\odot$ model as Figure \ref{fig:delta_r_TP_AGB}.}
\vskip .2in
\label{fig:B_seed_TAGB}
\end{figure}
Hydrodynamic turbulence is strongly suppressed in the tachocline. The mean-field approach to dynamo action then
suggests (see the review in Section \ref{s:dynthresh}) that dynamo activity should be easily sustained in the
tachocline as long as the toroidal magnetic field can be rotated back into the poloidal direction. This effect is present in numerical
simulations of a shear layer \citep{ClinBC2003}; and is conjectured to operate in a more global dynamo model where
toroidal magnetic field formed in the tachocline diffuses into the convective envelope \citep{park1993}.
Estimates of the equipartition toroidal magnetic field and the large-scale Maxwell stress are given by equations
(\ref{eq:bphieq}) and (\ref{eq:reyn}).
\begin{figure}[!]
\figurenum{7}
\plotone{f7}
\vskip 0in
\epsscale{1.0}
\caption{The magnetic helicity (expressed as a magnetic flux $\Delta {\cal H}^{1/2}$) that
accumulates in the time $t_{\rm drift}
= \ell_P/|v_r^{\rm nuc}|$. RGB (left) and AGB (right) evolution of $M_{\rm ZAMS} = 1\,M_\odot$ model
with a Jupiter companion in an initial orbit $a_i = 1$ AU.
Most of the helicity generated on the RGB is concentrated near the tip (orange short-dash curve shows cumulative helicity).
This gives $M_{\rm benv}(5.65\times10^7)$ $\sim 0.388M_\odot$ as the
effective base of the magnetized region. The net mass passing through the dynamo-active layer
is then $\sim0.12M_{\odot}$ on the RGB, and $\sim 0.014 M_{\odot}$ on the AGB.
The greater helicity accumulated through the `$\Omega(\theta)$' channel (Equation (\ref{eq:helicityf})),
as compared with the `inflow' channel (Equation (\ref{eq:helicityd})), is reflected in the final WD magnetic
field (Figure \ref{fig:B_WD_three_dynamos_vs_L_orb}).}
\vskip .2in
\label{fig:Delta_H_sqrt_TP-AGB_solar}
\end{figure}
\begin{figure}[!]
\figurenum{8}
\plotone{f8}
\epsscale{1.0}
\caption{Comparison of magnetic helicity generated over $t_{\rm drift} = \ell_P/|v_r^{\rm nuc}|$
through the same two channels as Figure \ref{fig:Delta_H_sqrt_TP-AGB_solar}, but now in the
$M_{\rm ZAMS} = 5\,M_\odot$ model without a planetary companion.
Only the AGB contributes directly to final WD magnetization in this case, within
a mass shell $\sim 0.044\,M_\odot$. The largest net contribution comes from the first thermal pulse.}
\vskip .2in
\label{fig:Delta_H_sqrt_TP-AGB_5M_solar}
\end{figure}
\section{White Dwarf Magnetic Field} \label{s:finalB}
After magnetic helicity accumulates in the core, we allow the magnetic field to relax to a more isotropic configuration.
A first estimate of the flux threading one hemisphere is
\begin{equation}\label{eq:phicore}
\Phi_{\rm core} \simeq {\cal H}^{1/2}.
\end{equation}
The final poloidal magnetic field of a WD of radius $R_{\rm wd}$ is
\begin{equation}\label{eq:bwd}
B_{\rm wd} \sim \frac{\Phi_{\rm core}}{\pi R_{\rm wd}^2}.
\end{equation}
We first evaluate the magnetic helicity that is deposited in a post-AGB star,
using the $1\,M_\odot$ and $5\,M_\odot$ stellar models constructed using MESA.
This stored helicity is then converted to a dipolar magnetic
field in the WD remnant using Equations (\ref{eq:phicore}) and (\ref{eq:bwd}).
This estimate presupposes
that toroidal and poloidal magnetic fluxes evolve to an energy minimum with $\Phi_p \sim \Phi_\phi$, all the while preserving
the product helicity ${\cal H} \sim \Phi_p\Phi_\phi$. Such an interchange may occur later in the life of a WD as the
result of ohmic diffusion (Section \ref{s:B_field_emergence_decay}); or earlier on as the result of relaxation within
episodes of confined convection (such as the core helium flash, or helium shell flashes).
\subsection{Accumulation of Magnetic Helicity in a \\ Radiative Core Relaxing to Solid Rotation} \label{s:mass_drift_dynamo}
The zone of interest here is the tenuous mantle of radiative material that is sandwiched between the hydrogen-depleted
core and the base of the convective envelope. Here only a small mass $\lesssim 10^{-3}\,M_\odot$ lies within
a scaleheight, and so we model the mass inflow through the mantle as quasi-steady,
\begin{equation}
\dot M_{\rm core} = 4\pi r^2 |v_r| \rho(r).
\end{equation}
The poloidal magnetic field is approximated as radial
with flux function (\ref{eq:flux}), and the toroidal magnetic field is expressed in terms of a radial magnetic twist
as $B_\phi(r,\theta) = (\partial\phi_B/\partial r) r\sin\theta B_r$. Then helicity accumulates in the core at the rate
\begin{equation}\label{eq:helicityf}
{d{\cal H}\over dt} = -2\pi\delta v_r\int d\theta \Phi_r {\partial \Phi_r\over\partial\theta} {\partial\phi_B\over \partial r},
\end{equation}
where $\delta v_r$ is the flow velocity relative to the core-envelope boundary.
Now consider the torque imparted to a mass shell by the Maxwell stress $B_rB_\phi/4\pi$,
\begin{equation}
{d\over dt}\left(\delta m {2\over 3}r^2\bar\Omega\right) = \delta\left[\int dS\; r\sin\theta {B_rB_\phi\over 4\pi}\right].
\end{equation}
Here the Lagrangian time derivative follows the radial flow at speed
$v_r$. Since $d(\delta m)/dt = 0$, one has, following Equation (\ref{eq:flux}),
\begin{equation} \label{eq:torque}
{d\over dt}\left({2\over 3}r^2\bar\Omega\right) = {1\over 8\pi \rho r^2}
\int d\theta \sin\theta \left({\partial \Phi_r\over\partial \theta}\right)^2 {\partial^2\phi_B\over \partial r^2}.
\end{equation}
Both hemispheres generally contribute in the same sense to the torque.
\subsubsection{Helicity Sourced by Radial Inflow}\label{s:massdrift}
We suppose that the magnetic field is strong enough that nearly solid rotation is maintained in the radiative layers,
$|\partial\Omega/\partial r| \ll \bar\Omega/r$. Each mass shell starts with a mean angular frequency $\Omega_{\rm benv}$ that is imposed by
the lower part of the convective envelope. The twist that is required to maintain a weak angular velocity gradient is found from Equation
(\ref{eq:torque}),
\begin{equation} \label{eq:twistb}
{\partial^2\phi_B\over\partial r^2} \sim -{2\dot M_{\rm core}\Omega_{\rm benv}r \over (B_r r^2)^2},
\end{equation}
where the coefficient on the right-hand side corresponds to $B_r$ independent of $\theta$. From Equation (\ref{eq:phiB}) one has
\begin{equation}
{\partial\phi_B\over\partial r} \sim {r\over |v_r|}{\partial\Omega\over\partial r}.
\end{equation}
Nearly solid rotation is maintained if
\begin{equation} \label{eq:solid}
v_{A,r}^2 \equiv {B_r^2\over 4\pi \rho} \gg v_r^2.
\end{equation}
Now consider the magnetic helicity that is advected into the growing radiative core. We evaluate the magnetic twist in Equation
(\ref{eq:helicityf}) from Equation (\ref{eq:twistb}), obtaining
\begin{equation} \label{eq:helicityd}
{d{\cal H}\over dt} = f(\Delta M_{\rm core})\cdot \pi \dot M_{\rm core} \Omega_{\rm benv} R_{\rm benv} ^2 \delta v_r(R_{\rm benv} )
\end{equation}
at $r = R_{\rm benv}$.
The factor $f$, with indeterminate sign but magnitude $|f| < 1$, represents the cancellation between hemispheres.
We discuss its origin in Section \ref{sec:cancellation_factor}.
\subsubsection{Magnetic Twist Compensating a Latitude-dependent Convective Torque}\label{s:contorque}
A greater Maxwell stress results from the application of convective stresses to the outer radiative core.
A convection zone generally sustains latitudinal gradients in rotation. Part of the differential
Reynolds stress is transmitted through the tachocline to the core by a dynamo-generated magnetic
field (Equation (\ref{eq:reyn})). As a working estimate, we use
\begin{equation}\label{eq:maxcon}
{B_rB_\phi\over 4\pi} = \varepsilon_B \rho(R_{\rm benv}) (\Omega_{\rm benv} R_{\rm benv})^2,
\end{equation}
with $\varepsilon_B \sim 10^{-3}$, corresponding to a dynamo period $P_{\rm dyn} \sim 10
[\Delta\Omega(R_{\rm benv})]^{-1}$ and a pole-equator offset $\Delta\Omega(R_{\rm benv}) \sim \Omega_{\rm benv}$.
Given that the deeper parts of the radiative core are coupled magnetically to the tachocline,
uniform rotation can be maintained in the outer core only if this stress is divergence-free below the tachocline.
The corresponding magnetic helicity flow into the core through the advection of this twisted field is
\begin{equation} \label{eq:helicitye}
{d{\cal H}\over dt} = f(\Delta M_{\rm core})\cdot \pi \varepsilon_B \dot M_{\rm core} R_{\rm benv} (\Omega_{\rm benv} R_{\rm benv})^2.
\end{equation}
The contributions to ${\cal H}$ from equations (\ref{eq:helicityd}) and (\ref{eq:helicitye}) are labelled
in the Figures as `inflow' and `$\Omega(\theta)$', respectively.
\subsubsection{Cancellation Between Hemisphere} \label{sec:cancellation_factor}
One feature of Equations (\ref{eq:helicityd}) and (\ref{eq:helicitye}) stands out: the growth of ${\cal H}$ is
independent of the seed radial magnetic field, as long as it is strong enough to satisfy the inequality (\ref{eq:solid}).
This means that the contribution to ${\bf A}\cdot{\bf B}$ {\it in one hemisphere} is not strongly affected by fluctuations
in the magnitude and sign of the poloidal magnetic field that is supplied to the radiative layer by a dynamo operating near
the convective boundary.
A net mass $\Delta M_{\rm core}$ flows through the convective-radiative boundary over the duration of dynamo activity. In the case of
an AGB star, this is approximately the increase in the mass of the C/O core due to double-shell burning. Note that $\Delta M_{\rm core}$
is generally much larger than the mass of the rarefied core mantle ($\sim 10^{-3}\,M_\odot$).
The low mass of the mantle imposes a short-timescale cutoff to fluctuations in the summed ${\cal H}$ from the two
hemispheres. We express this cutoff in terms of the mass within a scale height below the core-envelope boundary,
\begin{equation}
\Delta M_{\rm min} \sim {4\pi Pr^4\over GM_{\rm core}}\biggr|_{R_{\rm benv} }.
\end{equation}
Then we take
\begin{equation}\label{eq:fM}
f(\Delta M_{\rm core}) \sim \left({\Delta M_{\rm min}\over \Delta M_{\rm core}}\right)^{1/2}.
\end{equation}
\subsection{Results}
First consider the $M_{\rm ZAMS} = 1\,M_\odot$ model. The angular momentum profile of the star is evolved
in post-processing using the method of Section \ref{s:rotation} and Paper I. Companion planets with a range of
masses (Earth, Neptune, and Jupiter) are placed at an initial semi-major axis $a_i = [0.5,0.75,1,1.5,2]$ AU and with a range of
orbital eccentricities.
Magnetic fields generated during the RGB and AGB phases both contribute to the
remnant WD field, but at different depths in the star (see Figure \ref{fig:M_star_M_base_M_He_core_M_CO_core_TRGB_solar}). A net mass
$\sim 0.135M_{\odot}$ passes through the core-envelope boundary while conditions are favorable for magnetic field growth.
\footnote{For example, we do not include the first dredge-up phase, during which the convective envelope grows in mass.}
\begin{figure*}[!]
\figurenum{9}
\epsscale{1.0}
\plottwo{f9a}{f9b}
\vskip 0in
\caption{Dipole magnetic field, Equation (\ref{eq:bwd}), left behind in the WD remnant of a $1\,M_\odot$ star
interacting with a planet (black = Earth mass, red = Neptune mass, blue = Jupiter mass).
Result is plotted separately for the two helicity channels described in the text. Horizontal axis shows
initial orbital angular momentum of planet companion. Each tick represents an average over realizations of
a given set of orbital initial conditions ($a_i$ and $e_i$). Magenta squares show result for
$5\,M_\odot$ progenitor with initial equatorial rotation speed 50 km s$^{-1}$ and no companion; orange squares same
model with angular momentum of $3M_J$, $a_i = 2$ AU planet added during early AGB.
Dynamo shuts off when ${\rm Co}_{\rm benv} < {\rm Co}_{\rm benv,crit} = 0.1$ (left panel) or ${\rm Co}_{\rm benv,crit} = 0.3$ (right panel).}
\vskip .2in
\label{fig:B_WD_three_dynamos_vs_L_orb}
\end{figure*}
Figure \ref{fig:Delta_H_sqrt_TP-AGB_solar} shows the helicity that accumulates in the brief period $t_{\rm drift} = \ell_P/|v_r^{\rm nuc}|$
when stellar material settles through the outer scale height of the radiative core. The dominant contribution comes from the
`$\Omega(\theta)$' channel.
In the $5M_{\odot}$ star, only helicity accumulated on the AGB contributes directly to the final WD magnetic field
(Figure \ref{fig:M_star_M_base_M_He_core_M_CO_core_TRGB_5M_o}). During this phase $\sim 0.044 M_{\odot}$ passes
through the active dynamo region (given our normalization $\eta_B = 0.05$ of the Bl\"{o}cker mass loss formula).
We see in Figure \ref{fig:Delta_H_sqrt_TP-AGB_5M_solar} that the
initial phase of the dynamo, which occurs during the first thermal pulse, contributes the majority of ${\cal H}$.
In the 5 $M_{\odot}$ model, we simplified the treatment of the thermal pulses by linearly interpolating
all of the relevant quantities between the local minima in $R_{\rm benv}$ after each He shell flash, so as to focus on the
mean growth of core mass and ${\cal H}$. The stellar model considered in Figure \ref{fig:Delta_H_sqrt_TP-AGB_5M_solar}
has no planetary companion.
The resulting WD dipolar field is shown in Figure \ref{fig:B_WD_three_dynamos_vs_L_orb}
for both the $1\,M_\odot$ and $5\,M_\odot$ models, using
the mass-radius relation of \cite{zapos1969}.
A strong magnetic field in the $0.55\,M_\odot$ WD remnant of the $1\,M_\odot$ progenitor depends on the
injection of angular momentum from a planetary companion (as considered here), or a tidal interaction with a stellar
companion. We find that $B_{\rm wd}$ is an increasing function of the angular momentum absorbed;
the various points correspond to different realizations of the orbital evolution.
Dipole fields approaching $10^8$ G are achievable through the `$\Omega(\theta)$' channel, following the absorption of a Jupiter-mass planet.
Note that an Earth-mass planet does not signicantly augment the angular momentum left over at the end of the MS spindown, and so
this case corresponds closely to an isolated star that does not interact with planets.
Figure \ref{fig:B_WD_three_dynamos_vs_L_orb} focuses on those runs which end in the engulfment of the planet. However there are cases in which
the planet is pushed out into a large orbit due to the convective quadrupole. The planet therefore gains angular momentum from the star, which
in turn causes the star to spin in the opposite direction. In some cases this leads to the development of a dynamo as well. We ignore these
particular cases.
An external source of angular momentum injection is not required to sustain a dynamo in the $5\,M_\odot$ progenitor.
Nonetheless, $B_{\rm wd}$ depends on the C/O mass that accumulates during double shell burning, and therefore on the
mass-loss rate during the expulsion of the envelope.
The effect of adding a relatively massive companion to the $5\,M_\odot$ model is considered in Figure
\ref{fig:B_WD_high_mass}. The absorption of a $3\,M_J$ planet significantly augments the remnant dipole
field in the $0.87\,M_\odot$ WD. We see a steady increase in the remnant field as the companion mass increases to the brown
dwarf range, or even to $\sim 0.1\,M_\odot$, for the angular velocity profile considered here ($\alpha = 3/2$).
For comparison we show the magnetic fields obtained by a shallower inner rotation profile ($\alpha = 1$), which may be applicable
depending on the strength of the Coriolis back reaction. In this case an increase in angular momentum results in a decrease of remnant
magnetic field, caused by a decrease in ${\rm Co_{benv}}$ due to the shallow $\tau_{\rm conv}(r)$ profile.
\begin{figure}[!]
\figurenum{10}
\epsscale{1.0}
\plotone{f10}
\caption{Dipole magnetic field obtained via the $\Omega(\theta)$ channel in the $0.87\,M_\odot$ remnant of a $5\,M_\odot$ star that
absorbs an orbiting companion ($a_i = 2$AU) of various masses during the early AGB phase ($R_\star = 200R_\odot$).
Top points: rotation profile $\Omega(r) \propto r^{-1.5}$ ($r^{-2}$) in the inner (outer) convective envelope
($\alpha = 1.5$ in Equation (\ref{eq:ominner})).
Bottom points: $\alpha = 1$.}
\vskip .2in
\label{fig:B_WD_high_mass}
\end{figure}
\section{White Dwarf Rotation} \label{s:WD_rotation}
We now analyze the spin evolution of the core and envelope of our stellar models.
Our focus is on the magnetic {\it de}-coupling between the core and envelope. Three factors determine the
effectiveness with which a poloidal magnetic field transfers angular momentum across the core-envelope boundary,
and establishes solid rotation in the outer core: i) the timescale over which mass is exchanged between
core and envelope; ii) the direction of this exchange; and iii) the Coriolis parameter in the inner envelope.
The internal structure of the star evolves slowly enough during the early ascent of the giant branches, and during
core helium burning, that a magnetic coupling between core and envelope is very difficult to avoid
(see Figure 2 of Paper I). There is a much more rapid exchange of mass between envelope and core
near the tips of the RGB and AGB. The first giant phase is (except in cases of close binary interaction)
followed by more extended evolution. So our focus here is on the final AGB superwind phase during which most
of the envelope mass is removed.
The ability of the inner envelope to generate a magnetic field is most relevant
while the radiative core is growing in mass. Then magnetized material is advected downward into the outer
core, where it can communicate changes in rotation rate in the envelope to the inner core. The core
grows in mass during the thermally pulsating AGB phase, after averaging over the pulsations; but following
the contraction of the envelope the superwind causes a small but rotationally significant decrease in mass.
An important effect involves the re-expansion of the envelope following a late thermal pulse:
following this about $10^{-3}\,M_\odot$ of hydrogen-rich material is transfered
to the convective envelope. If core and envelope are still magnetically coupled, this
causes a dramatic spindown of the core through the internal exchange of angular momentum.
Faster WD spin is mediated by an early loss of angular momentum from
the surface of an AGB star. This loss, if strong enough, will push the Coriolis parameter ${\rm Co}_{\rm benv}$ of
the inner envelope below unity well before the envelope contracts. The remnant WD spin period depends on the critical value of
${\rm Co_{ benv}}$ below which the dynamo effectively shuts off.
If the transition to ${\rm Co}_{\rm benv} < {\rm Co}_{\rm benv,crit}$ takes place before the final shell flash, then the angular momentum of
the core is effectively frozen in and inherited by the WD. On the other hand,
if this transition occurs afterward, then we expect that the rotation of the outer core remains coupled to the inner
envelope as the photosphere of the star contracts.
Figures \ref{fig:Co_base_and_J_star_J_core_TAGB_closer_solar} and \ref{fig:Co_base_and_J_TAGB_closer_5M_solar_time_shift} show
the evolution of the net stellar angular momentum $J_\star$ and the core angular momentum $J_{\rm core}$ during the AGB phase, under the assumption
of a tight core-envelope coupling. The internal rotation profile, the prescription for mass and angular momentum loss,
and the interaction with a planetary companion, are the same as those described in Sections \ref{s:intro} \& \ref{s:basic}
and detailed in Paper I. As the star experiences strong mass loss, there is a rapid drop in $J_\star$.
\begin{figure}[!]
\figurenum{11}
\epsscale{1.0}
\plotone{f11}
\vskip 0in
\caption{Rotational angular momentum of $M_{\rm ZAMS} = 1\,M_\odot$ AGB star (dotted red line) and its core (short-dash orange line) just
before and during the super-wind phase. The star has absorbed a Jupiter-mass companion near the tip of the RGB.
Black line: Coriolis parameter at base of convective envelope. Rotation profile
$\Omega(r) \propto r^{-1.5}$ ($r^{-2}$) in the inner (outer) convective envelope.
When calculating remnant WD spins, we freeze $\Omega_{\rm core}$ when and if ${\rm Co}_{\rm benv}$ drops below a critical value,
beyond which time the curves plotted here do not apply.}
\vskip .2in
\label{fig:Co_base_and_J_star_J_core_TAGB_closer_solar}
\end{figure}
\begin{figure}[!]
\figurenum{12}
\epsscale{1.0}
\plotone{f12}
\vskip 0in
\caption{Similar to Figure \ref{fig:Co_base_and_J_star_J_core_TAGB_closer_solar}, but for the $M_{\rm ZAMS} = 5M_{\odot}$ star.}
\vskip .2in
\label{fig:Co_base_and_J_TAGB_closer_5M_solar_time_shift}
\end{figure}
During this phase, $J_\star$ is still dominated by the envelope.
Also plotted is ${\rm Co}_{\rm benv}$ for comparison. Core-envelope decoupling becomes plausible at
${\rm Co}_{\rm benv} = {\rm Co}_{\rm benv,crit} \sim 0.1$-0.3, given the strong dependence of magnetic field
on rotation that is observed in deeply convective MS stars \citep{reinbb2009}.
Whether the transition to ${\rm Co}_{\rm benv} < {\rm Co}_{\rm benv,crit}$ is encountered before the final
He shell flash depends on both $J_\star$ and the value of ${\rm Co}_{\rm benv,crit}$. The transition occurs earlier for a larger critical
Coriolis parameter, as in our $1\,M_\odot$ model with ${\rm Co}_{\rm benv} = 0.3$.
But we find a tighter restriction on $J_\star$ when ${\rm Co}_{\rm benv}$ is reduced to 0.1:
then the absorption of a Jupiter will cause core and envelope to remain coupled past the final flash
in the $1\,M_\odot$ model, but the absorption of a Neptune (or of no planet) allows a transition to a weakly magnetized envelope.
We find that ${\rm Co}_{\rm benv}$ generally remains above ${\rm Co}_{\rm benv,crit}$ at the final shell flash in the $5\,M_\odot$ model,
implying a tight core-envelope coupling during the transition to a post-AGB star.
In the following two sections, we separately consider the rotational evolution of our stellar models with, and without,
a transition to weak core-envelope coupling.
\subsection{Dependence of White Dwarf Spin on \\ Coriolis Parameter for Core-Envelope Decoupling} \label{s:Co_base_critical}
The WD spin period that results from core-envelope decoupling at ${\rm Co}_{\rm benv,crit} = 0.1$ and 0.3 in the $1\,M_\odot$
model is shown in Figure \ref{fig:P_rot_WD_Co_crit_w_5M_o}.
For the more conservative estimate ${\rm Co}_{\rm benv,crit} = 0.1$, we find that $P_{\rm wd}$ ranges over
$0.5$-$1.5$ d.
In this situation, where the timing of the transition to weak magnetization in the envelope depends on $J_\star$,
$P_{\rm wd}$ depends on the initial orbital angular momentum of the planetary companion. A lower total angular momentum results in faster
decoupling, and therefore a {\it larger} trapped core angular momentum: the core rotation frequency at decoupling is tied to the convective
timescale in the inner envelope, and to the structure of the star.
This behavior is explained in Figure \ref{fig:J_Co_base_eq_1_TAGB}, which shows the minimum angular momentum required to
maintain ${\rm Co}_{\rm benv} = 0.1$ in the $1\,M_\odot$ model near the tip of the AGB. This is compared with the actual angular momentum
evolution that results from the absorption of a Jupiter, Neptune or Earth, initially orbiting at $a_i=1$ AU with eccentricity $e_i=0$.
\begin{figure}[!]
\figurenum{13}
\epsscale{1.0}
\plotone{f13a}
\vskip 0.2in
\plotone{f13b}
\vskip 0in
\caption{Rotation period of $0.55\,M_\odot$ WD remnant of $1\,M_\odot$ star, as a function of the initial orbital
angular momentum of planetary companion. Squares: Core and envelope remain magnetically coupled until the post-AGB phase.
x's: Core-envelope decoupling before the final helium shell flash, with $J_{\rm wd}$ set to $J_{\rm core}$ at decoupling.
Decoupling takes place when ${\rm Co}_{\rm benv}$ drops below ${\rm Co}_{\rm benv,crit} = 0.1$ (upper panel) or
$0.3$ (lower panel). Horizontal magenta line: result for $0.87\,M_\odot$ remnant of $M_{\rm ZAMS} = 5\,M_\odot$ star with
50 km s$^{-1}$ equatorial rotation period.}
\vskip .2in
\label{fig:P_rot_WD_Co_crit_w_5M_o}
\end{figure}
\begin{figure}[!]
\figurenum{14}
\epsscale{1.0}
\plotone{f14}
\vskip 0in
\caption{Black solid line: minimum $J_\star$ that maintains ${\rm Co}_{\rm benv} = 0.1$ near the tip of the AGB of a
$M_{\rm ZAMS} = 1\,M_\odot$ star. For comparison, colored lines show calculated angular momentum of star
absorbing a Jupiter (red dotted), Neptune (orange short-dashed) or Earth (blue long-dashed) starting with $a_i=1$ AU, $e_i=0$.
$J_{\rm core}$ is shown separately for Neptune and Earth interaction, where core and envelope decouple relatively early.
Decoupling is delayed to the final super-wind phase when the companion has a Jupiter mass,
resulting in lower $J_{\rm core}$ and the longer WD spin seen in Figure \ref{fig:P_rot_WD_Co_crit_w_5M_o}.}
\vskip .2in
\label{fig:J_Co_base_eq_1_TAGB}
\end{figure}
\subsection{White Dwarf Spin in Case of Continued Core-Envelope Coupling}
We observe a significant difference in the remnant angular momentum in two models which maintain a tight
core-envelope coupling through to the post-AGB phase: the $5\,M_\odot$ model, and the $1\,M_\odot$ model
which absorbs a Jupiter-mass planet. In the first case,
downward angular momentum pumping in the convective envelope allows $J_{\rm core}$ to remain relatively
constant during the ejection of the envelope, and only modest spindown is caused by a wind from
the surface of the post-AGB star (Figure \ref{fig:Co_base_and_J_TAGB_closer_5M_solar_time_shift}).
The spin of the WD remnant is directly
proportional to the initial spin period of the star, except in the case of the absorption of
a relatively massive ($\gtrsim 3M_J$) companion.
An equatorial rotation speed $50$ km s$^{-1}$ on the zero-age MS results in $P_{\rm wd} \simeq 0.6$ days.
The initial contraction of the envelope also does not cause much of a reduction in $J_{\rm core}$ in the
$1\,M_\odot$ model; but we observe a much stronger decrease following the re-expansion and re-contraction
of the envelope that is triggered by a late helium shell instability (Figure \ref{fig:Co_base_and_J_star_J_core_TAGB_closer_solar}).
This means that the WD remnant of the $5\,M_\odot$ star has a spin period $P_{\rm wd} \sim 15\,$ hr, as compared with $\sim 2$ yr for the
remnant of the $1\,M_\odot$ star.
To understand how such a dramatic difference in final WD spin could arise, we consider a simplified analytic model
of mass and angular momentum loss from an AGB star. The envelope of the model star has uniform specific angular momentum,
the core has a fixed radius $R_{\rm benv}$, and the core rotates as a solid body with angular velocity
$\Omega_{\rm core} = \Omega_{\rm benv}$. Mass is lost from the outer boundary of the star, and is also
exchanged between core and envelope,
\begin{equation}
{dM_{\rm env}\over dt} = {dM_\star\over dt} - {dM_{\rm benv}\over dt}.
\end{equation}
Here $M_{\rm env}$ is the mass of the convective H-rich envelope, and $M_{\rm benv}$
is the mass of (mainly radiative) material inside the base of the envelope.
The moment of inertia of the core is initially a small fraction of the total effective moment of inertia
\begin{equation}
I_{\rm eff} = {J_\star\over \Omega_{\rm benv}} = {2\over 3}M_{\rm env}R_{\rm benv}^2 + I_{\rm core},
\end{equation}
where we parameterize $I_{\rm core} = (2\epsilon_{\rm core}/3) M_{\rm benv} R_{\rm benv}^2$.
The coefficient $\epsilon_{\rm core}$ is very small, but a key consideration is its relative
size compared with the fraction $M_{\rm env}/M_\star$ of the stellar mass that is contained in
the envelope. We find that $\epsilon_{\rm core}$ ranges over $\sim 3\times 10^{-5}-10^{-3}$ during
the superwind and post-AGB phases. The envelope of the $1\,M_\odot$ model completes its first contraction
while it retains $\sim 10^{-3}\,M_\odot$, but following its re-expansion and re-contraction this mass has dropped
to $M_{\rm env} \sim 10^{-4}\,M_\odot$. In the $5\,M_\odot$ model, the envelope experiences a single contraction when its mass
drops below $\sim 10^{-4}\,M_\odot$. See Figure \ref{fig:jvsm}.
Consider first the stage(s) when the envelope is still inflated, and the core radius $R_{\rm benv}$ changes only
slowly. Then
\begin{eqnarray}
{dJ_\star\over dt} &=& {2\over 3}R_\star^2\Omega(R_\star){dM_\star\over dt} \mbox{} \nonumber \\ \mbox{}
&=& {2\over 3}R_{\rm benv}^2\Omega_{\rm benv} {dM_{\rm env}\over dt} + I_{\rm eff} {d\Omega_{\rm benv}\over dt}.
\end{eqnarray}
and
\begin{equation}
{1\over\Omega_{\rm benv}}{d\Omega_{\rm benv}\over dt} = {dM_{\rm benv}/dt\over M_{\rm env} + \epsilon_{\rm core} M_{\rm benv}}.
\end{equation}
Changes in the radius of the star do not enter into these expressions as a result of the uniform
specific angular momentum profile in the envelope.
The core mass increases relatively slowly compared with the rapid drop in $M_{\rm env}$ during the peak of
the superwind. So $\Omega_{\rm benv}$ increases only slightly up to the contraction of the envelope, where
it takes the value $\Omega_{\rm benv,col}$. Beyond that point, the entire star rotates nearly as a solid body.
A wind from its surface, which carries away a mass $\Delta M_\star$ causes a net spindown
\begin{equation}
{\Omega_{\rm benv,col}'\over\Omega_{\rm benv,col}} \simeq \exp\left[-{1\over\epsilon_{\rm core}}{\Delta M_\star\over
M_\star}\right].
\end{equation}
This works out to a factor $\sim 0.2$ decrease in $\Omega_{\rm benv}$ in the $1\,M_\odot$ model at the onset of the
late helium shell flash.
The key step in the dramatic spindown of the post-AGB star involves the re-expansion of the envelope. This occurs
by the transfer of $\sim 10^{-3}\,M_\odot$ of hydrogen-rich material from the radiative part of the star into a rejuvenated convective layer,
all occuring at nearly constant total angular momentum. Now the core rotation rate (still equal to the rotation rate at the base
of the envelope due to a tight magnetic coupling) decreases to
\begin{equation}\label{eq:omegaf}
{\Omega_{\rm benv}\over\Omega_{\rm benv,col}'} = {\epsilon_{\rm core}M_{\rm benv}\over M_{\rm env} + \epsilon_{\rm core}M_{\rm benv}} \sim
{\epsilon_{\rm core}M_{\rm benv}\over M_{\rm env}} \sim {1\over 30}.
\end{equation}
(Although the star contracts by a factor $\sim 10$ in between the initial envelope contraction and the late
helium shell flash, $\epsilon_{\rm core}$ returns to its pre-contraction value after the re-expansion of the envelope.)
The rotation rate of the core remains nearly constant at the value (\ref{eq:omegaf}), as most of the rejuvenated envelope is expelled.
The net spindown from the initial contraction of the envelope works out to $\Omega_{\rm benv}/\Omega_{\rm benv,col} \sim 1/150$.
This behavior is neatly encapsulated by the trend of $J_\star$ and $J_{\rm core}$ with the mass $M_\star-M_{\rm wd}$ remaining to be expelled
(Figure \ref{fig:jvsm}).
\begin{figure}[!]
\figurenum{15}
\epsscale{1.0}
\plotone{f15a}
\vskip .2in
\plotone{f15b}
\vskip 0in
\caption{Total spin angular momentum of AGB star during the contraction of its hydrogen envelope (black solid line)
and the angular momentum of the core (red dotted line). Top panel: $1\,M_\odot$ model, including both the initial
envelope contraction and the re-expansion following a late helium shell flash. Bottome panel: $5\,M_\odot$ model.
We show for comparison the dimensionless core moment of inertia $\epsilon_{\rm core} = I_{\rm core}/(2/3)M_{\rm benv}R_{\rm benv}^2$
(orange short-dash line).}
\vskip .2in
\label{fig:jvsm}
\end{figure}
\hfil
\subsection{Minimal Magnetic Flux Enforcing \\ Strong Core-Envelope Coupling} \label{s:coupling_flux}
Here we evaluate the minimal magnetic flux that must thread the core-envelope boundary in order for
i) the outer radiative core to remain in solid rotation; and ii) the rotation of the inner core to remain coupled
to the outer core and inner envelope. A related (although less quantitative) estimate has been made by \cite{sp98} for
the central cores of massive stars during the later stages of nuclear burning.
First consider the outer core. Although it contains a small fraction of the core mass, it
can contribute significantly to the moment of inertia. We define a characteristic magnetic flux
threading a single hemisphere by setting $B_r/\sqrt{4\pi \rho(R_{\rm benv})} = |v_r - dR_{\rm benv}/dt|$:
\begin{equation}\label{eq:phi_solid}
\Phi_{\rm solid} = \pi R_{\rm benv}^2 |v_r - dR_{\rm benv}/dt| \sqrt{4\pi\rho(R_{\rm benv})}.
\end{equation}
Next consider changes in the angular momentum of the entire core, which are sensitive to the evolving
rotation profile of the envelope. The core angular momentum
responds to the applied Maxwell stress $B_rB_\phi/4\pi$ according to
\begin{equation}
I_{\rm core}{d\Omega_{\rm core}\over dt} \sim
\int\frac{B_r B_{\phi}}{4\pi} R_{\rm benv} \sin\theta \cdot 2\pi R_{\rm benv} ^2 \sin\theta d\theta.
\end{equation}
Taking a second time derivative, making use of the induction equation, we get
\begin{equation}
I_{\rm core} \frac{d^2\Omega_{\rm core}}{dt^2}
\sim \frac{1}{2} \int B_r^2 \frac{\partial\Omega}{\partial r} R_{\rm benv} ^4\sin^3\theta d\theta.
\end{equation}
Here the poloidal magnetic field is assumed to be purely radial, so that $\partial_r(r^2B_r) = 0$, and the
integral is evaluated at radius $R_{\rm benv}$. Integrating over $\theta$ then gives
\begin{equation}
I_{\rm core} \frac{d^2\Omega_{\rm core}}{dt^2} \sim \frac{2}{3} B_r^2 R_{\rm benv} ^4 \frac{\partial\Omega}{\partial r}.
\end{equation}
We take the rotation rate to vary according to
\begin{equation}
\Omega(r,t) \sim \Omega_{\rm benv} \left( \frac{R_{\rm benv} }{r} \right)^{2} \exp\left[-\frac{t}{\tau_{\Omega}} \right].
\end{equation}
Then to avoid spindown on a timescale $\tau_\Omega$, the magnetic flux must exceed
\begin{eqnarray}\label{eq:phicoup}
\Phi_{\rm couple} &=& \pi R_{\rm benv}^2 B_r \sim \sqrt{\frac{3}{4}} \pi \frac{I_{\rm core}^{1/2} R_{\rm benv} ^{1/2}}
{\tau_{\Omega}} \\ \nonumber
&\simeq& 7.1 \times 10^{20} \ {\rm Mx} \left( \frac{I_{\rm core}}{10^{-4}M_{\odot}R_{\odot}^2} \right)^{1/2} \\\nonumber
&& \quad\quad \left( \frac{R_{\rm benv} }{R_{\odot}} \right)^{1/2} \left( \frac{\tau_{\Omega}}{10^3 {\rm yr}} \right)^{-1}.
\end{eqnarray}
We note that $\tau_\Omega$ is controlled by the rate at which material is ejected from the envelope during
the final superwind phase, and as a result $\Phi_{\rm couple}$ can in principle be larger or smaller than
$\Phi_{\rm solid}$ (which is determined by the requirement that the core itself remain in solid rotation).
These estimates of the magnetic flux that will enforce a tight rotational coupling between core and envelope
are compared in Figures \ref{fig:Co_base_and_Phi_couple_Phi_dynamo} and
\ref{fig:Co_base_and_Phi_couple_Phi_dynamo_time_shift} with the radial flux $(\Delta{\cal H})^{1/2}$
that accumulates at the core-envelope boundary. This
increment $\Delta{\cal H}$ is calculated using Equation (\ref{eq:helicitye}), taking the timescale
for helicity accumulation to be the minimum of the radial drift time $\ell_P/|v_r^{\rm nuc}|$ and $\tau_\Omega$.
The larger value of the dynamo-generated flux supports our choice of ${\rm Co}_{\rm benv} < 1$ for
core-envelope decoupling. Recall also from Section \ref{s:bseed} that
a tiny poloidal magnetic field -- corresponding to $10^{-4} \sqrt{4\pi \rho(R_{\rm benv})} v_{\rm con}$
during the later stages of the AGB -- will seed a toroidal field in the tachocline
that is strong enough to interact buoyantly with the layers above (see Figure \ref{fig:B_seed_TAGB}).
\begin{figure}[!]
\figurenum{16}
\epsscale{1.0}
\plotone{f16}
\caption{Comparison between magnetic flux $\Phi_{\rm couple}$ (Equation (\ref{eq:phicoup}), red dotted line)
that maintains core-envelope coupling on the AGB, flux $\Phi_{\rm solid}$ (Equation (\ref{eq:phi_solid}), purple dot-dashed line)
that maintains solid rotation in the outer core against the inward mass flow,
and flux $\Delta{\cal H}^{1/2}$ generated by the $\Omega(\theta)$ dynamo process on radial drift time (orange short-dashed line).
Blue long-dashed line: magnetic flux corresponding to a poloidal field in equipartition with the convective motions,
$\Phi_{\rm eq} = \pi R_{\rm benv}^2 \sqrt{4\pi \rho(R_{\rm benv}) v_{\rm con}^2}$.
Solid black line: Coriolis parameter at base of envelope. We argue that the strong dependence of the dynamo-generated field
on ${\rm Co}_{\rm benv}$ allows for core-envelope decoupling below ${\rm Co}_{\rm benv} \sim 0.1-0.3$. Fluxes are only
plotted up to the final thermal pulse: afterward $dM_{\rm benv}/dt < 0$ and the coupling of the outer core
to the envelope is less sensitive to the instantaneously generated magnetic field.}
\vskip 0in
\label{fig:Co_base_and_Phi_couple_Phi_dynamo}
\end{figure}
\begin{figure}[!]
\figurenum{17}
\epsscale{1.0}
\plotone{f17}
\caption{Similar to Figure \ref{fig:Co_base_and_Phi_couple_Phi_dynamo}, but for the $M_{\rm ZAMS} = 5\,M_\odot$ model.}
\vskip .2in
\label{fig:Co_base_and_Phi_couple_Phi_dynamo_time_shift}
\end{figure}
\hfil
\section{Magnetic Field Emergence and Decay}\label{s:B_field_emergence_decay}
We now consider the emergence of magnetic fields from post-AGB stars, and their subsequent decay. The lifetime of the
visible magnetic field at the surface of the star depends on the thickness of the magnetized layer, which in turn
depends on the mass of the progenitor and the mass loss rate during the expulsion of the hydrogen envelope.
The strong dredge-up of helium that is experienced by stars of mass $> 2.3\,M_\odot$ leaves only
a thin mass shell to be processed on the AGB. The remainder of the C/O material is generated during core He burning,
when the central convective core is decoupled from the exterior and therefore must conserve ${\cal H}$.
On the other hand, the helium generated on the RGB is not dredged up in solar-mass stars, and so
magnetic fields deposited during both the red giant and asymptotic giant phases contribute directly to the remnant
WD field.
We found that mass loss on the AGB can push the convective envelope below the threshold
for dynamo activity (Figures \ref{fig:Co_base_and_J_star_J_core_TAGB_closer_solar} and
\ref{fig:Co_base_and_J_TAGB_closer_5M_solar_time_shift}). In this case, a thin outer layer of the remnant is initially
unmagnetized; a strong magnetic field only emerges by ohmic diffusion. We do not consider any contribution to the magnetic
field from the contraction of the final $10^{-4}\,M_\odot$ of envelope material, which might be rapidly rotating as
the result of anisotropic mass loss \citep{Spru1998}.
The timescale for flux emergence, and the asymptotic decrease in the trapped magnetic flux,
are easily obtained from the induction equation,\footnote{Hall drift does not modify an
axisymmetric magnetic field in a fluid star, and so we consider only ohmic effects here.}
\begin{equation} \label{eq:induction}
\frac{\partial {\bf B}}{\partial t} = - \nabla \times (\eta\nabla \times {\bf B}).
\end{equation}
Here $\eta = c^2/4\pi \sigma$ is the magnetic diffusivity expressed in terms of electrical conductivity $\sigma$ and speed of light $c.$
For degenerate matter in the liquid state the conductivity can be estimated using \citep{yacku1980,itohmii1983},
\begin{equation}\label{eq:sigma}
\sigma = 8.5 \times 10^{21} {\rm s}^{-1} \ \frac{1}{\Lambda_{ei} \left<Z\right>} \frac{x^3}{1 + x^2}.
\end{equation}
Here $\Lambda_{ei} \simeq 1$ is the Coulomb logarithm; $\langle Z\rangle = \mu_e \sum X_i Z_i^2 / A_i$, with $X_i$ the mass fraction,
$Z_i$ and $A_i$ the nuclear charge and mass of species $i$, and $\mu_e$ is the mean molecular weight per electron.
Finally $x = p_F / m_e c$ measures the Fermi momentum of the electrons (of mass $m_e$).
A good approximation to the diffusion timescale through a pressure scale height $l_P$ is
\begin{equation} \label{eq:t_diff_l_P}
t_{{\rm diff},l_P} \simeq \frac{4\pi \sigma l_P^2}{c^2}.
\end{equation}
An analytic approximation to $t_{{\rm diff},l_P}$ is easily obtained in a thin outer
shell of mass $\Delta M$ that is supported by the pressure of non-relativistic electrons,
\begin{equation}
P = K\rho^{5/3} = \frac{(3\pi^2)^{2/3}}{5} \cdot \frac{\hbar^2 }{m_e ( \mu_e m_u )^{5/3}} \rho^{5/3},
\end{equation}
where
\begin{equation}
P \simeq g{\Delta M\over 4\pi R_{\rm wd}^2} = {GM_{\rm wd}\Delta M\over 4\pi R_{\rm wd}^4}
\end{equation}
in a nearly uniform surface gravitational field $g$. The integration of the equation of hydrostatic equilibrium gives
a pressure scale height $l_P = (2/5)(R_{\rm wd}-r)$ and density $\rho(l_P) = (gl_P/K)^{3/2}$.
Substituting $\sigma$ from equation (\ref{eq:sigma}) leads to
\begin{eqnarray} \label{eq:t_diff_analytic}
t_{\rm diff,analytic} &\sim& 1.8 \times 10^{10} \ {\rm years} \left(\frac{\Delta M}{M_\odot}\right)^{7/5}
\left(\frac{ M_{\rm wd}}{M_\odot}\right)^{-3/5} \mbox{} \nonumber \\ \mbox{}
&& \left(\frac{R_{\rm wd}}{10^{-2}R_\odot}\right)^{-8/5}
\left(\frac{Y_e}{0.5}\right)^{2} \left(\frac{\left<Z\right>}{6}\right)^{-1} \Lambda_{ei}^{-1}. \mbox{} \nonumber \\ \mbox{}
\end{eqnarray}
From this expression it is easy to see that the diffusion time through a shell of fixed $\Delta M$
depends weakly on the total WD mass.
The diffusion time through the magnetized surface layer of our 0.55 and $0.87\,M_\odot$ model WDs is shown in
Figures \ref{fig:T_Ohm_diffusion_mass_even_earlier_model} and \ref{fig:T_Ohmic_diffusion_B_layer_mass_non_degenerate}.
The longer diffusion time in the $0.55\,M_\odot$ remnant is due to the larger magnetized mass.
A direct integration of the induction equation is straightforward if we ignore the continuing hydromagnetic adjustment
that must accompany ohmic diffusion. Making the same approximation of a geometrically thin magnetized layer, one has
\begin{equation}
{\partial \Phi_{r,\phi}\over\partial t} = {c^2\over 4\pi\sigma(r)}{\partial^2 \Phi_{r,\phi}\over\partial r^2}
\end{equation}
for both the hemispheric poloidal flux $\Phi_r(r) \simeq 2\pi R_{\rm wd}^2\int d(\cos\theta) B_r(r,\theta)$,
and the toroidal flux $\Phi_\phi(r,\theta) \simeq 2\pi R_{\rm wd}\int_r^{R_{\rm wd}} dr' B_\phi(r',\theta)$ (defined
locally in $\theta$).
The decay of the flux is shown in Figure \ref{fig:fdecay}, along with the toroidal magnetic energy. We choose a
simple initial magnetic configuration with toroidal flux concentrated between depth $0.5\Delta r$ and $\Delta r$,
where $\Delta r = R_{\rm wd}-r(\Delta M)$ is the depth of the magnetized layer. Both the toroidal
and poloidal fluxes decrease by a factor $\sim 1/7$ over a time interval 10 times longer than the local decay time
$\ell_P^2/\eta(\Delta r)$, as measured at depth $\Delta r$.
\begin{figure}[!]
\figurenum{18}
\epsscale{1.0}
\plotone{f18}
\caption{Ohmic diffusion time in magnetized layers of $0.55\,M_\odot$ WD remnant of $1\,M_\odot$ star,
as given by Equations (\ref{eq:t_diff_l_P}) (black line)
and the analytic fit (\ref{eq:t_diff_analytic}) (magenta dashed line).
Mass shells passing through the dynamo-active layer on the RGB (AGB) are bounded by green (blue) dashed lines (when the star has absorbed a
Jupiter-mass planet).
The hump in the diffusion times close to the surface is caused by a shift in composition and electrical conductivity.}
\vskip .2in
\label{fig:T_Ohm_diffusion_mass_even_earlier_model}
\end{figure}
\begin{figure}[!]
\figurenum{19}
\epsscale{1.0}
\plotone{f19}
\caption{Similar to Figure \ref{fig:T_Ohm_diffusion_mass_even_earlier_model}, but for $0.87\,M_\odot$ WD remnant of$5\,M_\odot$ star.
Mass processed during the tail of the first thermal pulse, which produces the majority of the magnetic flux (Figure
\ref{fig:Delta_H_sqrt_TP-AGB_5M_solar}) is bounded by the dashed green line.}
\vskip .2in
\label{fig:T_Ohmic_diffusion_B_layer_mass_non_degenerate}
\end{figure}
\begin{figure}[!]
\figurenum{20}
\epsscale{1.0}
\plotone{f20a}
\plotone{f20b}
\vskip 0in
\caption{Top panel: ohmic diffusion of toroidal magnetic field intially concentrated at depths $0.5\Delta r$ to $\Delta r\ll R_{\rm wd}$
below the surface of a cold white dwarf. Here $\Delta r$ is the maximum depth of the magnetized layer of mass $\Delta M$.
Bottom panel: decrease of the toroidal (or radial) magnetic flux threading this layer, along with the
toroidal magnetic energy, both normalized to the initial values.
This calculation does not allow for an interchange between toroidal and poloidal fluxes
resulting from a hydromagnetic instability that might accompany ohmic drift.}
\vskip .2in
\label{fig:fdecay}
\end{figure}
At the top of the magnetized layer of both model WDs, $t_{{\rm diff},l_P}\sim 10^7$ yr.
We conclude that the magnetic field, although initially buried, will emerge at a moderate age in both intermediate-mass and
high-mass white dwarfs.
In the $0.55\,M_\odot$ remnant of the solar-mass star
$t_{{\rm diff},l_P}$ extends up to $\sim 1-2 \times 10^{9}$ yr at the base of the magnetized layer; whereas it reaches a maximum
$\sim 2\times 10^8$ yr in the $0.87\,M_\odot$ remnant of the $5\,M_\odot$ star.
Although essentially all WD progeny of intermediate-mass stars will have strong buried toroidal magnetic fields -- due to the
high angular momentum remaining at the end of the MS -- some decay in this buried field is expected at ages exceeding $\sim 1$ Gyr.
These ohmic timescales depend on the thickness of the C/O layer accumulated on the thermally pulsating AGB. The WD masses
we have obtained correspond to a normalization $\eta_B = 0.05$ to the mass-loss rate on the AGB using the formula
of \cite{bloe1995}. There are some suggestions of a lower normalization
(e.g. $\eta_B \sim 0.01$, based on the abundance of Li-rich giants in the LMC: \citealt{ventura00}),
which implies a larger final C/O core. For example, increasing the remnant mass of the $5\,M_\odot$ progenitor
to $\sim 0.90\,M_\odot$ results in an ohmic diffusion time $\sim 4\times 10^8$ yr at the base of the magnetized layer).
One sees from Figure \ref{fig:fdecay} that any toroidal flux buried in such a WD will have decreased by a factor
$\sim 1/5$ from its post-AGB value by an age $3\times 10^9$ yr.
Whether this decay of a buried field corresponds to significant decay of the surface dipole magnetic field -- or even
to growth -- remains uncertain because the two components do not evolve independently. Ohmic diffusion of the magnetic field
is so slow that the magnetic field easily makes a continuing hydromagnetic adjustment to something close to
magnetostatic equilibrium. An interchange between the toroidal and poloidal components is possible
without any change in ${\cal H}$.
\section{Summary and Comparison with Alternative Theoretical Approaches} \label{s:conclusions}
We have investigated how the rotation and magnetism of a giant star respond
to the inward advection of a small amount of angular momentum by deep
convective plumes in the extended envelope. Some features of the
dynamo process operating near the core-envelope boundary are unique to giants:
namely those driven by the intense radiation flux and the
rapid inward drift of material to the burning shell(s). The buoyancy that
is induced by radiative heating of magnetized material strongly enhances
the rate of mixing between core and envelope.
Even a weak helical magnetic field enforces a rotational coupling between
core and envelope, and maintains nearly solid rotation in the core.
The inflow of mass into the core is accompanied by an
inflow of magnetic helicity, which is needed to sustain a stable, large-scale
magnetic field in the white dwarf remnant. The compensating helicity is
lost through the surface of the star.
This strong core-envelope coupling is easily maintained in subgiants
and core He burning stars, as we found in Paper I. It is also sustained
near the tips of the RGB and AGB in isolated intermediate-mass stars
if angular momentum is distributed broadly within the slowly
rotating parts of the envelope. We found that the threshold for dynamo
action is easily reached when $\Omega \propto r^{-2}$ at ${\rm Co} \lesssim 1$.
Injection of angular momentum from a planetary or stellar companion
is needed to maintain a magnetic core-envelope coupling during the giant
expansion of late-type stars that lose most of their initial spin to magnetized
winds (corresponding to $M_{\rm ZAMS} \lesssim 1.3\,M_\odot$).
Although a strong coupling does not depend on angular momentum injection
in intermediate-mass stars, the remnant magnetic field may be significantly
enhanced by the absorption of a massive planet or brown dwarf, due
to the relatively low mass of the outer shell that becomes magnetized.
These results are broadly consistent with an increased incidence of
strong magnetism in i) massive WDs, and ii) accreting WDs
in CVs, which experienced a tidal interaction before and during a
common-envelope phase.
The strength of the remnant dipole magnetic field is directly tied to the magnetic helicity
that has accumulated in the hydrogen-depleted core.
Helicity growth depends on a combination of i) efficient angular momentum
transport by poloidal magnetic fields through the outer core; and ii) a persistent radial
angular velocity gradient in the tachocline that is driven by angular inhomogeneities
in the rotation rate in the surrounding convective envelope. The helicity
flux is roughly proportional to the rotational kinetic energy of the inner
envelope.
The WD dipole fields so obtained can exceed $10^7$ G in both
the $0.55\,M_\odot$ WD remnant of a solar-mass star that absorbs a Jupiter,
and the $0.87\,M_\odot$ remnant of a $5\,M_\odot$ star that has no binary interaction
(Figure \ref{fig:B_WD_three_dynamos_vs_L_orb}).
In the former case, the engulfment of companion much more massive than Jupiter requires
exceptional circumstances, and this field can be considered a approximate upper limit.
In the latter case, the
engulfment of much more massive companion, even $\sim0.1M_\odot$, may be significantly more
common. This results in an increased remnant field, reaching as high as $\sim10^8$ G
(Figure \ref{fig:B_WD_high_mass}).
These field strengths do not approach the strongest measured in WDs
($\sim 10^9$ G), which however are very rare. The strongest fields
may therefore originate in more extreme merger events such
as the collision of two WDs (e.g. \citealt{gb12}).
The engulfment of an Earth-mass planet by an evolved solar-mass star, which barely augments the angular momentum of
the star, results in a WD dipole field of $\sim10^6$ G. In this case, we find that most of the magnetic helicity is deposited
during a brief interval at the start of core He burning. The star falls below the threshold for dynamo activity near the tips of the
RGB and AGB. Even a mild enhancement of wind-driven angular momentum loss, or a softening of the rotation profile in the envelope, would
eliminate most of this MG magnetic field.
The magnetic field generated on the giant branches is initially buried in the WD (in cases where
AGB core and envelope decouple before the final thermal pulse) but diffuses
ohmically outward over the first $\sim 10^7$ yr. A strong internal
toroidal field begins to decay at an age $\sim 4\times 10^8$ yr
in a $\sim 0.9\,M_\odot$ WD, as compared with $\sim 2\times 10^9$ yr
in a $0.55\,M_\odot$ WD. We find that the net decay is by a factor $\sim 0.2$
at an age $\sim 3\times 10^9$ yr in a $0.9\,M_\odot$ WD. The initial
decay of the toroidal field may be accompanied by transient growth of
the external dipole, which is possible at constant internal magnetic helicity.
On the observational side, the compilation of magnetic WDs in \cite{kepletal2013}
only extends to an age $\sim 1$ Gyr, and the incidence
of magnetism in older WDs remains uncertain.
We find that the spin angular momentum of the remnant WD of a
late-type star can, in some circumstances,
{\it anti-correlate} with the angular momentum of the star during the AGB phase.
In stars that retain only a moderate angular momentum at the end
of the AGB (less than the orbital angular momentum of Jupiter),
this decoupling sets in before the final helium shell flash.
The final WD spin then depends most directly on the critical Coriolis parameter
below which the envelope dynamo fades away in the inner envelope.
In the case of absorption of a Neptune-mass planet during the post-MS
evolution of a $1\,M_\odot$ star, $P_{\rm wd}$ lies between 0.1 and 1 d.
This residual spin rate is a lower bound, as it does not include a
contribution from a final anisotropic contraction of the envelope
\citep{Spru1998}.
The situation changes if the star is massive enough to retain most of
its natal angular momentum, or if it absorbs a Jupiter-mass planet
or brown dwarf. Then core and envelope should remain coupled beyond
the final helium shell flash, and spindown of the core can be directly
related to a transient reduction in core mass. The angular momentum that is retained
by the core during the contraction of the envelope depends on whether
the core experiences a late re-expansion due to a delayed helium shell
instability. A transfer of mass $M_{\rm env} \gg \epsilon_{\rm core}M_\star$
to the envelope, combined with ejection of most of this envelope material,
implies strong spindown of the core.
The net result is that the final WD spin period can exceed a year in our
$1\,M_\odot$ model. This final spindown was found to be significantly
smaller in the $5\,M_\odot$ model, resulting in $P_{\rm wd}$ shorter than a day,
because of the smoother transition to the post-AGB phase.
\subsection{Some Outstanding Issues}
{\it Sensitivity to strength of angular-momentum pumping in giant envelope.} The Kepler
asteroseismological data give a calibration of the envelope rotation profile in subgiants -- as
investigated in detail in Paper I -- but not in stars larger than $\sim 10\,R_\odot$. Here we
have considered the consequences of uniform $dJ/dM$ in the slowly rotating parts of the envelope.
This zone with ${\rm Co} < 1$ extends deep into the star near the tips of the RGB and AGB, unless
the star interacts with a relatively massive binary companion. A rotation profile $\Omega \propto r^{-2}$
in the outer envelope is supported by the numerical simulations of \cite{brunp2009}.
The orbital angular momentum of a Jupiter can be compared with the
minimum rotational angular momentum that will sustain a dynamo near the tip of the AGB.
Starting the planet at $a_i = 1$ AU, the excess is about a factor
$\sim 30 ({\rm Co}_{\rm benv,crit}/0.3)^{-1}$ in our $1\,M_\odot$ model. The
shallowest rotation profile which could maintain a minimal Coriolis parameter ${\rm Co}_{\rm benv} \sim 0.3$
in an envelope of depth $R_{\rm benv}/R_\star \sim 0.01$ is $\Omega(r) \propto r^{-4/3}$.
In this situation, the amplitude of the helicity flux into the core, which is proportional to
$\Omega_{\rm benv}^2$ (Equation (\ref{eq:helicitye})), would be reduced by a factor $\sim 10^{-3}$.
The corresponding reduction in $B_{\rm dipole}$, by a factor $\sim 0.03$,
would still allow for fields of order MG (see Figure \ref{fig:B_WD_three_dynamos_vs_L_orb}),
but not magnetic fields $\gtrsim 10$ MG.
A related consideration is the steepness of the angular velocity profile in the inner envelope,
where ${\rm Co} \gtrsim 1$, and its effect on the upper range that is obtained for the WD magnetic field.
In this paper we did not simply take the profile that was determined in Paper I from asteroseismic
models of sub-giant and helium burning stars. During these relatively compact evolutionary phases,
an inner rotation profile $\Omega(r) \propto r^{-1}$ corresponds to ${\rm Co_{benv}} \sim 10$-$30$. This profile is consistent with
the inward pumping of angular momentum by deep convective plumes in an adiabatic envelope with gravity $g(r) \propto r^{-1}$.
A lower inner Coriolis parameter would be maintained near the tips of the RGB and AGB if the same rotation
profile were sustained there.
In fact, the rotation profile is expected to steepen to $\Omega(r) \sim r^{-3/2}$ in the
presence of a similar convective structure, due to the contraction of the core and the steepening of the gravity profile.
For this reason, we allow the inner index in Equation (\ref{eq:ominner}) to increase to $\alpha = 3/2$ during the
expanded phase where magnetic helicity growth is concentrated, and the remnant WD spin angular momentum is determined.
The net effect is to raise the upper envelope of the magnetic field distribution in massive WDs from
$\sim 3$ MG ($\alpha=1$) to 10-30 MG ($\alpha=3/2$). The
limiting magnetization and spins of lower-mass WDs hardly change over
this range of $\alpha$.
{\it What happens to any core magnetic field left over from the MS phase, when exposed to
the turbulent motions in a slowly rotating convective envelope} (${\rm Co}_{\rm benv}~\ll~1$)?
Consider taking the magnetic flux threading a WD and spreading it across the
core-envelope boundary of the progenitor during its AGB expansion. Even a MG magnetic field in the
WD becomes dynamically insignificant, $B_r(R_{\rm benv}) \sim 10^{-4}(4\pi \rho v_{\rm con}^2)^{1/2}$.
Such a weak poloidal magnetic field would rapidly diffuse over the boundary
sphere, causing fluid elements that are threaded by radial field of
opposing signs to be mixed together. In this situation
it is difficult to see how the material flow from the envelope into the core could maintain a large-scale poloidal
magnetic field unless it were self-consistently maintained by a dynamo process. A numerical experiment
studying the diffusion and reconnection of such a seed field across the spherical boundary between radiative
and convective layers is feasible and would provide interesting results.
{\it Relaxation of helical magnetic field to the `maximum-dipole' configuration.} The surface dipole magnetic of the WD
remnant has been estimated assuming that $\Phi_r \sim \Phi_\phi \sim {\cal H}^{1/2}$. Other magnetic configurations are
possible, especially those with relatively stronger toroidal fields, corresponding to $\Phi_\phi \gg \Phi_r
= ({\cal H}/\Phi_\phi)^{1/2}$ (\citealt{Brai2009}). Ohmic diffusion in WD stars with strong toroidal fields
may cause some transient growth of the dipole field, as the magnetic field relaxes to a more isotropic configuration.
{\it Dispersion in magnetic field strength in massive WDs.} Although a greater
fraction of massive WDs exhibit strong magnetic fields than do their $\sim 0.55-0.6\,M_\odot$ cousins, a majority do not.
Intermediate-mass stars retain more angular momentum at the end of the MS, thereby facilitating a dynamo.
This angular momentum can, nonetheless, be enhanced by the absorption of a massive planet, brown dwarf, or
low-mass star, or a tidal interaction with a stellar companion. Other effects causing dispersion in the
visible surface magnetic field include exchange between toroidal and poloidal magnetic fluxes, and the
time spent by the progenitor star during the thermally pulsating AGB phase.
{\it Emergence of `frozen' magnetic field generated during the pre-MS phase (or during MS core convection)?} The outer
$0.3\,M_\odot$ of the $0.55\,M_\odot$ WD is processed through a deep convective envelope during the
RGB and AGB phases. The proportion of this material which experiences rapidly rotating convection,
and develops a strong magnetic field, depends on the initial placement of planets around the star. For
example, if the closest Neptune or Jupiter orbits at a distance $\sim 1$ AU, then only the outer half
of the processed layer is strongly magnetized.
The ohmic diffusion time from the base of this processed layer approaches a Hubble time, independent
of the architecture of the planetary system. From this we conclude that the magnetization of all but
the very oldest $\sim 0.55$-$0.6\,M_\odot$ WDs is not significantly influenced by a `frozen' core magnetic field.
The assembly history is very different for the massive WD remnants of intermediate-mass stars. Now the bulk of the
WD mass is processed during core H and He burning, which leave unaltered the magnetic helicity
stored in the stellar core. In the case of an isolated $\sim 0.9\,M_\odot$ WD, a core magnetic field could emerge
ohmically through the outer $\sim 0.05-0.1\,M_\odot$ that is processed on the thermally pulsating AGB, but
only with a delay of $\sim 2-5\times 10^8$ yr (depending on the normalization of mass loss during this phase).
A growth in the incidence of magnetism in older WDs could therefore point to a contribution of such a `frozen' field
to the visible surface field. (But not exclusively so, given the competing possiblity of dipole field growth
mediated by the decay of a toroidal field that is buried in a thin outer mass shell.)
{\it Implications of Binary Magnetic WDs.} The incidence of strong magnetism (exceeding $10^7$ G
and implying classification as a `polar') exceeds $\sim 30\%$ in
short-period CVs with ages exceeding $\sim 1$ Gyr \citep{Gans2005}. Our calculations
suggest that these fields might decay significantly in more massive isolated WDs. Long-lived
magnetism in old binary WDs with periods $< 2$ hr (below the `period gap') could be maintained
by the accretion of $\gtrsim 0.1\,M_\odot$ from the companion, which would have the effect of pushing
down the magnetized layers to a greater depth and a longer ohmic timescale.
\begin{figure}[!]
\figurenum{21}
\epsscale{1.0}
\plotone{f21}
\vskip 0in
\caption{Dependence of AGB stellar radius $R_\star$ (black line) and base of convective envelope $R_{\rm benv}$
(dotted red line) on residual H mass. Endpoint of $M_{\rm ZAMS} = 1\,M_\odot$ model.}
\vskip .2in
\label{fig:AGBendpoint}
\end{figure}
\subsection{Competing effects of a collapsing low-mass ($\sim 10^{-4}\,M_\odot$) AGB envelope}
The final contraction of an AGB envelope, beginning at a photospheric radius $\sim 10^2\,R_\odot$,
leaves behind a hydrogen layer of mass $M_{\rm env,f} \sim 10^{-4}\,M_\odot$.
Mass loss slows down significantly as the star contracts inside a radius $R_{\rm pAGB} \sim R_\odot$ (Figure \ref{fig:AGBendpoint}).
It is possible that the process driving mass loss
is so anisotropic that the collapsing envelope carries with it a significant angular momentum,
equal and opposite to that lost during the final stages of the super-wind \citep{Spru1998}.
An upper bound to the angular momentum thus added to the post-AGB star can be obtained as follows.
The collapsed envelope will experience a bar instability and
spread outward into a disk if its angular velocity exceeds
$\sim 0.3(GM_{\rm wd}/R_{\rm pAGB}^3)^{1/2}$ at radius $R_{\rm pAGB}$. Any angular momentum in excess of
$J_{\rm max} \sim 0.3 M_{\rm env,f} (GM_{\rm wd}R_{\rm pAGB})^{1/2}$ is transported outward by viscous torques in the disk, and is
presumably lost in a wind.
Setting the envelope angular momentum to this limiting value, and assuming that it is shared with the rest of the star
at radius $R_{\rm pAGB}$, we obtain a minimum spin period for the cold WD remnant,
\begin{equation}
P_{\rm wd} \geq {2\pi I_{\rm wd}\over J_{\rm max}} = 0.4\,\left({M_{\rm env,f}\over 10^{-4}\,M_\odot}\right)^{-1}
\left({R_{\rm pAGB}\over R_\odot}\right)^{-1/2} \quad {\rm hr}.
\end{equation}
A comparison with Figure \ref{fig:P_rot_WD_Co_crit_w_5M_o} shows that somewhat faster WD
spins can be imparted by rotating envelope contraction than would result from a cutoff of the envelope dynamo
before the final intense phase of mass loss. However, the fastest WD spins may depend on a merger between two WDs.
\subsection{Impulsive Dynamo Amplification in a Merger?}
The formation of magnetic WDs in stellar binaries is considered by \cite{ToutWLFP2008}. It is noted that the injection
of angular momentum from a binary companion (either a low-mass MS star or a substellar companion) into a giant star
will trigger strong differential rotation in its core and envelope.
This differential rotation may spark the impulsive growth of the magnetic
field in the core, further details of which have been examined by \cite{WickTF2014}.
\cite{NordWSMB2011} consider the tidal shredding of a planetary companion close to the giant core.
They argue that this material initially forms a disk in nearly Keplerian rotation about the core, in which a magnetic field is generated by
the magnetorotational instability. The most promising application of this process is to isolated magnetic WDs, since it may be difficult
for a planet and a stellar companion (later the mass donor in a CV system) to co-exist on similar orbits.
The contribution made by such an impulsive dynamo to the net magnetic helicity stored in the stellar core can be compared with
the longer term effect examined in this paper. When the companion is tidally disrupted (either within the convective
envelope, or within the rarefied mantle surrounding the burning shells), the mass $M_2$ deposited would
be re-distributed back through the convective envelope on the thermal timescale
\begin{equation}\label{eq:tth}
t_{\rm th} \sim {GM_{\rm core} M_2 \over R_{\rm tide} L_{\rm AGB}} \sim 600 \left({M_2\over 0.1\,M_\odot}\right)
\left({R_{\rm tide}\over R_\odot}\right) \quad {\rm yr}.
\end{equation}
As envisaged by \cite{NordWSMB2011}, a centrifugally supported disk may spread on a shorter timescale
than (\ref{eq:tth}), in which case the internal heat generated in the disk is advected around without being
radiated back into the surrounding stellar envelope.
Studies of advective accretion flows in other contexts (e.g. \citealt{bb99}) suggest that a fraction of the disk
material would be accreted deeper into the gravitational potential of the degenerate core,
which may be enough to transiently increase its rotation.
Since the energy released by H burning easily suffices to unbind the ashes from the degenerate
core, any such rapidly accreted material must remain H rich. On the longer thermal timescale,
this nuclear energy can be radiated away, accompanied by a relaxation of the outer core to its equilibrium structure.
The mass $\delta M_{\rm nuc}$ processed by H and He burning during that relaxation is small compared with
the total deposited mass $M_2$. Given that a mass $M_{\rm acc}$ is accreted close to the burning radius
$R_{\rm burn} \sim 0.1\,R_\odot$, one has
\begin{eqnarray}
\delta M_{\rm nuc} &\sim& {GM_{\rm core}\over \varepsilon_{\rm nuc} R_{\rm burn}} M_{\rm acc} \mbox{} \nonumber \\ \mbox{}
&\sim& 2\times 10^{-3}\left({R_{\rm burn}\over 0.1~R_\odot}\right)^{-1} \left({M_{\rm acc}\over M_2}\right) M_2.
\end{eqnarray}
The normalization is about $\sim 10$ times larger if H burning is extinguished,
so that only the He-burning shell remains active.
We see that a relatively small mass is incorporated into the
C/O core before the star adjusts to a new equilibrium structure. Given that the magnetic coupling between core and envelope
is sustained, most of the added angular momentum in this new equilibrium state will be stored in the outer envelope.
The large-scale Maxwell stress $B_r B_\phi/4\pi$ in the outer core may easily increase by a factor $\sim 10^4$ in the
immediate aftermath of the tidal disruption, when the rotation frequency increases by $\gtrsim 10^2$. However, as the
envelope dredges up the injected angular momentum, we expect the Maxwell stress to adiabatically adjust downward, since
little mass flows into the core. Such a downward adjustment could be avoided if the magnetic field lines
became decoupled from the envelope, but that does not seen consistent with the rapid buoyant motion of such a
strong magnetic field (see Equation (\ref{eq:buoy})).
From this perspective, the {\it impulsive} growth of the magnetic field following the tidal disruption of a low-mass star or
substellar companion may leave little permanent mark on the magnetization of the C/O core. A merger during the early
post-MS expansion could induce some direct hydrodynamic mixing with the hydrogen-depleted core, but only when the accumulated
helium mass was still well below the final WD mass.
We argue that the dominant effect of such a merger is, instead, to i) push the envelope above the threshold for a gentler and more persistent
dynamo process in low-mass giant stars, $M_{\rm ZAMS} \lesssim 1.3\,M_\odot$; and ii) to augment the inward flux of magnetic helicity in stars
which remain rapidly rotating at the end of the MS.
The merger of two WDs (or of a WD with the core of an evolved companion) provides the main counter-example
to these conclusions. Now hydrogen-depleted material in the more massive WD experiences direct
hydrodynamic mixing with the tidally shredded remnants of the companion. The rate of such events appears to
be relatively small, and may therefore only accomodate the most strongly magnetized WDs.
\acknowledgements We would like to thank David Arnett and Peter Goldreich for conversations. This work was supported by NSERC.
\begin{appendix}
\section{Twisting of a Radial Magnetic Field} \label{s:twist}
Here we construct expressions describing the large-scale flow of magnetic
helicity in the evolving core of a giant star. The rotation profile
and magnetic field are assumed to be axisymmetric, but departures from
reflection symmetry about the rotation axis are considered.
Our treatment is simplified
by considering a weak radial magnetic field with angular flux profile
\begin{equation}
\Phi_r(\theta) = \int_0^\theta \sin\theta' r^2B_r(\theta') d\theta'
\end{equation}
and corresponding potential
\begin{equation}
A_\phi(\theta,r) = {1\over r\sin\theta}\Phi_r(\theta).
\end{equation}
In what follows, we therefore neglect any radial change in $\Phi_r(\theta)$.
A dynamo operating near the convective boundary will generally supply a
poloidal field of a fluctuating sign, but the contribution to
${\bf A}\cdot{\bf B}$ in one hemisphere is insensitive to this sign. The
contributions from the two hemispheres generally have opposing signs.
Consider the case where the rotation profile is reflection-symmetric,
so that convective motions enforce a certain $\partial\Omega/\partial r$
at polar angles $\theta$ and $\pi-\theta$. The evolution of the
twist in the core depends on the connectivity of the poloidal magnetic field.
If both angles $\theta$ and $\pi-\theta$ are connected by the same
(axially-symmetric) magnetic flux surface, then there is no evolution
of the net twist along this flux surface. On the other hand, if the distribution
of magnetic flux is different across the two hemispheres, then a
given flux surface will experience a net differential winding.
The rate of change of ${\cal H}$ inside a spherical boundary of a fixed radius $r$ is
\begin{equation} \label{eq:helicity}
{\partial{\cal H}\over \partial t} = -\int dS\,\hat r\cdot\left[ c\phi {\bf B} + ({\bf A}\cdot{\bf B})
{\bf v} - ({\bf A}\cdot {\bf v}){\bf B}\right],
\end{equation}
where the bounding sphere is placed just inside the convective boundary.
The electrostatic potential $\phi$ appearing on the right-hand side of (\ref{eq:helicity}) is driven by the mean rotation of the star.
The corresponding electric field ${\bf E} = -(\mbox{\boldmath$\Omega$}\times {\bf r})\times{\bf B}/c$ is sourced by
\begin{equation}
\phi(\theta,r) = {1\over c}\Omega(r) \Phi_r(\theta).
\end{equation}
Differential rotation generates $A_r$ (and thence helicity) via
\begin{equation}
{\partial A_r\over \partial t} = -{\partial\Omega\over\partial r}\Phi_r(\theta).
\end{equation}
The electrostatic potential cancels off the term $A_\phi \hat\phi\cdot(\mbox{\boldmath$\Omega$}\times{\bf r})$ in Equation (\ref{eq:helicity}), and we are
left with
\begin{equation} \label{eq:helicityb}
{d{\cal H}\over dt} \;=\; {\partial{\cal H}\over \partial t} + {dR_{\rm benv} \over dt}\int dS A_\phi B_\phi
\;=\; -\delta v_r\int dS\,A_\phi B_\phi.
\end{equation}
Here we have subtracted off the velocity of the convective boundary from the radial flow speed $v_r$ of spherical mass shells:
\begin{equation}
\delta {\bf v} = {\bf v} - {dR_{\rm benv} \over dt}\hat r.
\end{equation}
On the AGB, $v_r$ is negative and larger in magnitude than $dR_{\rm benv} /dt$, but we include this correction for completeness.
Because the poloidal flux surfaces are, in this treatment, fixed in position, it is useful to re-express the toroidal field in terms of a
twist angle $\phi_B$:
\begin{equation} \label{eq:bphi}
B_\phi(r,\theta) = {\partial\phi_B\over\partial r} r\sin\theta B_r.
\end{equation}
The toroidal component of the induction equation is
\begin{equation}
{\partial B_\phi\over \partial t} = B_r r\sin\theta {\partial\Omega\over\partial r} -
{1\over r}{\partial\over\partial r} \left(r B_\phi v_r\right),
\end{equation}
which simplifies to
\begin{equation} \label{eq:phiB}
{d\over dt}\left({\partial\phi_B\over \partial r}\right) = {\partial^2\phi_B\over\partial t\partial r} + v_r {\partial^2\phi_B\over
\partial r^2} = {\partial\Omega\over \partial r} - {\partial v_r\over\partial r}{\partial\phi_B\over\partial r}.
\end{equation}
in the case (considered here) that $\Phi_r$ only depends on latitude.
Substituting expression (\ref{eq:bphi}) into (\ref{eq:helicityb}) gives
\begin{equation} \label{eq:helicityc}
{d{\cal H}\over dt} = -2\pi\delta v_r\int d\theta \Phi_r {\partial \Phi_r\over\partial\theta} {\partial\phi_B\over \partial r},
\end{equation}
which mirrors the standard expression for a static plasma.
In general the radial magnetic flux profile differs in shape, as well as sign,
between the two hemispheres. A complete
cancellation in the integral over $\theta$ in Equation (\ref{eq:helicityc}) remains possible if $\phi_B$ is independent of $\theta$.
That would be the case if fluid stresses were to smooth out latitudinal differential rotation within each shell, so that
$\partial\Omega/\partial r$, the source for $\phi_B$, were only a function of radius. We see that a net flux of magnetic helicity into the
core depends on {\it latitudinal} differential rotation -- differential rotation across flux surfaces -- which is a generic consequence of
convection.
\end{appendix}
|
\section*{Introduction}
Rigidity results of elastic materials have been of great interest in mathematical continuum mechanics in recent years, in particular since the seminal work by Friesecke, James and M\"uller \cite{fjm}.
Such results yield a deeper insight into the properties of materials through an estimate of the distance of the deformation gradient from the set of rotations; this distance is in turn estimated from above by the free energy of the system.
The rigidity estimates turn out to be crucial steps in various proofs as for instance of $\Gamma$-convergence results in the context of dimension reduction. This was also the case in our earlier paper \cite{LaPaSc2014}, in which we derived a discrete to continuum limit and a dimension reduction of an energy of a heterogeneous nanowire (see \cite{LPSpro} for an abridged version). There we presented a detailed analysis of the passage from the two-dimensional setting to the one-dimensional limit, and we gave a summary of the corresponding dimension reduction from three dimensions to one dimension. The purpose of this article is to show the rigidity estimates (Section~\ref{sec:rigidity}) and the main features of the latter case in detail.
Further, we elaborate on various three-dimensional lattices that are of importance in applications: the face-centred cubic lattice, the hexagonal close-packed, the body-centred cubic lattice and the diamond cubic lattice, see Section~\ref{3d}. These lattices occur for instance in aluminum and gold, magnesium and zinc, iron and tungsten, and germanium and silicon, respectively. Note that Si/Ge nanowires have applications in the semiconductor optoelectronics \cite{Kavanagh,Schmidtetal2010}.
In Section~\ref{sec:rigidity} we show that our discrete rigidity result applies to all these lattice structures.
The main property of such lattices is their geometric rigidity: they define a tessellation of the space into rigid polyhedra whose edges correspond to bonds in the lattice.
Our approach does not work in non-rigid lattices, like a simple cubic crystal with nearest-neighbour interactions only.
\par
We are interested in the mathematical modeling of dislocations in heterogeneous nano\-wires.
We assume that the material consists of two parts with the same lattice structure but different lattice constants. The interface between the two parts is assumed to be flat. The material overcomes the lattice mismatch either defect-free or by creating dislocations. As was pointed out by Ertekin et al. \cite{egcs}, it is the radius of the nanowire which determines whether the material creates dislocations or is defect free. In our model the radius roughly corresponds to the number of layers of atoms parallel to the direction of the wire, see Section \ref{sec:not}.
We prove that it is energetically more favourite to create dislocations than to relieve the mismatch in a defect-free way if the thickness of the nanowire is sufficiently large (see Remark~\ref{finalrem}).
The underlying idea of our mathematical model, which we introduce in Section~\ref{sec:not} in detail, goes back to the variational model proposed in \cite{mp} in the context of nonlinear elasticity
and which was later generalized
to a discrete to continuum setting in \cite{LaPaSc2014}. As before we assume that the total energy only consists of nearest-neighbour interactions which are harmonic, though it is possible to generalize this as discussed in \cite[Section~4]{LaPaSc2014}. In order to be able to apply a rigidity estimate, we always impose a non-interpenetration condition, which ensures that the deformations of the discrete setting preserve the orientation of each cell; similar assumptions were made e.g.\ in \cite{BSV,FT}. The non-interpenetration assumption can be dropped if one takes into account interactions beyond nearest neighbours, see the recent work \cite{ALP}. It is worth mentioning that a related variational model for misfit dislocations has been recently proposed in \cite{FPP}.
As in \cite{LaPaSc2014} we distinguish the systems with and without defects already in the given reference configuration. For both such systems we study the corresponding free energy of nearest- neighbour interactions in a discrete to continuum limit with dimension reduction. For the definition of the nearest neighbours in the discrete settings close to the interface it is useful to work with the notion of Delaunay triangulations and Voronoi cells, see \cite{LaPaSc2014}, where this was introduced for the first time to describe configurations with dislocations, see also Section~\ref{3d} for an introduction.
In Section~\ref{sec:4} we compare the minimizers of the limiting functionals, which characterize the minimum cost needed to compensate the lattice mismatch with and without defects, respectively. It turns out that this cost depends on the thickness of the wire described by a mesoscale parameter $k$. More precisely, it depends quadratically on $k$ if there are dislocations, and scales like $k^3$ if there are no defects. Hence for sufficiently large $k$, i.e. large radius of the wire, dislocations are energetically preferred. The result is based on a scaling argument. In particular for applications in semiconductor optoelectronics it would be interesting to know the threshold $k_c$ below which the nanowire deforms defect-free. This is however out of reach with our current methods so that we leave this as an open problem for future research.
\par
\section{Three-dimensional lattices}\label{3d}
We consider various three dimensional lattices whose unit cells are rigid convex polyhedra.
In this context, rigidity is understood in the following sense:
once the lengths of the edges of a polyhedron are given,
then the polyhedron is determined up to rotations and translations,
under the assumption that the polyhedron itself is convex.
We recall that a convex polyhedron is rigid if and only if its facets are triangles,
according to the classical Cauchy Rigidity Theorem (see, e.g., \cite{handbook-geometry}).
We consider four types of discrete lattices in dimension three:
the face-centred cubic, the hexagonal close-packed, the body-centred cubic,
and the diamond cubic. They should be interpreted as prototypes to which our approach can be applied,
under slight modifications in each case. For a general overview on lattice structures see, e.g., \cite{Grosso}.
\par
All the lattices we will introduce, fulfil a property of rigidity.
Indeed, the corresponding nearest-neighbour bonds provide a tessellation of the space
into rigid convex polyhedra, as we will make precise case by case.
(In the diamond cubic, also next-to-nearest neighbours will be used.)
We always assume a non-interpenetration condition, see \eqref{ad-3d} below.
\par
A major role in modeling is then played by the choice of the
nearest neighbours of each lattice.
Here they are defined according to the notion of
\emph{Delaunay pretriangulation}, as given in the following definitions.
Such a general definition can be applied also when the lattice is irregular,
so in particular across the interface between the phases, see Section \ref{subsec:setting}.
\par
For later convenience we give the definition for all dimensions $N\ge2$.
Let $\L\subset{\mathbb R}^N$ be a countable set of points such that there exist $R,r>0$ with
$\inf_{x\in{\mathbb R}^N} \# \big(\L\cap B(x,R)\big)\ge1$ and
$|x-y|\ge r$ for every $x,y\in\L$, $x\neq y$,
where $B(x,R):=\{y\in{\mathbb R}^N\colon |x-y|<R\}$.
\begin{definition}[Voronoi cells] \label{def:Voronoi}
The Voronoi cell of a point $x\in\L$ is the set
$$
C(x):= \{ z\in{\mathbb R}^N\colon |z-x|\le|z-y| \ \forall\, y\in\L \} \,.
$$
The Voronoi diagram associated with $\L$ is the partition $\{C(x)\}_{x\in\L}$.
\end{definition}
\begin{definition}[Delaunay pretriangulation]\label{def:Del-pre}
The Delaunay pretriangulation associated with $\L$ is a partition of ${\mathbb R}^N$ in open nonempty hyperpolyhedra with vertices in $\L$,
such that two points $x,y\in\L$ are vertices of the same hyperpolyhedra if and only if $C(x)\cap C(y)\neq\emptyset$.
\end{definition}
\begin{definition}[Nearest neighbours]\label{def:NN}
Two points $x,y\in\L$, $x\neq y$, are said to be nearest neighbours (and we write: $\NN{x}{y}$)
if they are vertices of an edge of one of the hyperpolyhedra of the Delaunay pretriangulation.
\end{definition}
\begin{definition}[Next-to-nearest neighbours]\label{def:NNN}
Two points $x,y\in\L$, $x\neq y$, are said to be next-to-nearest neighbours
(and we write: $\NNN{x}{y}$)
if, setting
$$
\L_*(x):=\L\setminus\{y\colon \NN{x}{y}\}\,,
$$
we find that $\H^{N-1}(C^x_*(x)\cap C^x_*(y))>0$, where $\{C^x_*(y)\}_{y\in\L_*(x)}$ is
the Voronoi diagram associated with $\L_*(x)$.
\end{definition}
The Voronoi diagram and the Delaunay pretriangulation
associated with a lattice are unique.
For these and other properties we refer to \cite[Section 1]{LaPaSc2014} and references therein.
\subsection{FCC lattice}
The face-centred cubic lattice is the typical structure of metals such as aluminium, gold, nickel, and platinum.
It is the Bravais lattice generated by the vectors
$$
\mathrm{v}\fcc{}_1:=\sqrt2(1,0,0) \,, \quad
\mathrm{v}\fcc{}_2:=\sqrt2\big(\tfrac12,\tfrac12,0\big) \,, \quad
\mathrm{v}\fcc{}_3:=\sqrt2\big(0,\tfrac12,\tfrac12\big) \,,
$$
namely
$$
\L\fcc{}:= \{ \xi_1\mathrm{v}\fcc{}_1+\xi_2\mathrm{v}\fcc{}_2+\xi_3\mathrm{v}\fcc{}_3 \colon \ \xi_1,\xi_2,\xi_3\in{\mathbb Z} \}.
$$
The resulting lattice is obtained by repeating periodically in the space
a cubic cell of side $\sqrt2$,
where the atoms lie at the vertices and at the centre of each facet.
It is readily seen that two points $x,y\in\L\fcc{}$ are nearest neighbours in the sense of Definition \ref{def:NN}
if and only if $\mod{x-y}=1$, i.e.,
they are joined by half a diagonal of a facet of the cubic cell.
Each atom has twelve nearest neighbours.
The Delaunay pretriangulation provides a subdivision of the space
into regular tetrahedra and octahedra of side one, thus in rigid convex polyhedra,
see Figure \ref{fig:fcc}.
Remark that the diagonals of the octahedra, whose length is $\sqrt2$,
correspond to next-to-nearest neighbours. The latter will not enter the definition of the energy
\eqref{en:fcc}.
\begin{figure}[p]
\centering
\subfloat[]{
\includegraphics[width=.24\textwidth]{fcc11.eps}
\label{fig:subfig1}
}
\hfill
\subfloat[]{
\includegraphics[width=.24\textwidth]{fcc31p.eps}
\label{fig:subfig3}
}
\hfill
\subfloat[]{
\includegraphics[width=.24\textwidth]{fcc21.eps}
\label{fig:subfig2}
}
\caption{In the face-centred cubic lattice
the nearest-neighbour structure of the atoms provides a subdivision of the space into tetrahedra (a) and octahedra (b). Figure~(c) shows a quarter of an octahedron in the same unit cell.
Grey dots denote points lying on the hidden facets.
}\label{fig:fcc}
\end{figure}
\subsection{HCP lattice}
Our approach works also for non-Bravais lattices such as
the hexagonal close-packed structure found in some metals as, e.g., magnesium and zinc.
It is defined by
$$
\L\hcp{}:=\{\u\hcp{}_i + \xi_1\mathrm{v}\hcp{}_1+\xi_2\mathrm{v}\hcp{}_2+\xi_3\mathrm{v}\hcp{}_3 \colon \ \xi_1,\xi_2,\xi_3\in{\mathbb Z} \,, \ i=1,2\} \,,
$$
where
$$
\mathrm{v}\hcp{}_1:=\big(0,0,\tfrac{2\sqrt{6}}{3}\big)\,, \quad \mathrm{v}\hcp{}_2:=\big(\tfrac{1}{2},\tfrac{\sqrt{3}}{2},0\big)\,,\quad
\mathrm{v}\hcp{}_3:=\big(-\tfrac{1}{2},\tfrac{\sqrt{3}}{2},0\big)
$$
are generators of two sublattices and
$$
\u\hcp{}_1:=(0,0,0)\,,\quad
\u\hcp{}_2:=\big(0,\tfrac{\sqrt{3}}{3}, \tfrac{\sqrt{6}}{3}\big)
$$
are called vectors of the basis.
The lattice is thus obtained by merging two Bravais sublattices (defined for $i=1$ and $i=2$, respectively).
As in the previous case, the nearest neighbours are those couples with distance one,
each atom has twelve nearest neighbours,
and the Delaunay pretriangulation consists of regular tetrahedra and octahedra of side one,
see Figure \ref{fig:hcp}. As before, the diagonals of the octahedra, which correspond to
next-to-nearest neighbour interactions, will not enter the definition of the energy \eqref{en:hcp}.
\begin{figure}[p]
\centering
\subfloat[]{
\includegraphics[width=.28\textwidth]{hcp11.eps}
}
\hspace{.15\textwidth}
\subfloat[]{
\includegraphics[width=.28\textwidth]{hcp21.eps}
}
\caption{The hexagonal close-packed lattice is associated with
a tessellation of tetrahedra and octahedra as the ones in the figure.
Only some of the bonds and some of the polyhedra of the pretriangulation are displayed.
}\label{fig:hcp}
\end{figure}
\subsection{BCC lattice}
The body-centred cubic lattice
\begin{figure}[p]
\centering
\includegraphics[width=.45\textwidth]{bcc11.eps}
\caption{The body-centred cubic lattice is associated with
a tessellation of irregular tetrahedra as the one in the figure.
}\label{fig:bcc}
\end{figure}
is typical of some metals as, e.g., iron and tungsten.
It is the Bravais lattice generated by the vectors
$$
\mathrm{v}\bcc{}_1:=\tfrac{\sqrt2}{2}(-1,1,1) \,, \quad
\mathrm{v}\bcc{}_2:=\tfrac{\sqrt2}{2}(1,-1,1) \,, \quad
\mathrm{v}\bcc{}_3:=\tfrac{\sqrt2}{2}(1,1,-1) \,,
$$
namely
$$
\L\bcc{}:= \{ \xi_1\mathrm{v}\bcc{}_1+\xi_2\mathrm{v}\bcc{}_2+\xi_3\mathrm{v}\bcc{}_3 \colon \ \xi_1,\xi_2,\xi_3\in{\mathbb Z} \} \,.
$$
The resulting lattice can be viewed by repeating periodically in the space a cubic cell of side $\sqrt2$,
where the atoms lie at the vertices and at the centre of the cube.
According to Definition \ref{def:NN},
the nearest neighbours are those couples with distance $\frac{\sqrt6}{2}$,
(i.e., those joined by half a diagonal of the cubic cell,)
as well as those couples with distance $\sqrt2$
(i.e., those joined by an edge of the cubic cell).
Thus, in contrast with the face-centred cubic,
in this case the notion of nearest neighbours differs
from other notions based on the Euclidean distance.
According to this definition each atom has 14 nearest neighbours.
Correspondingly, the Delaunay pretriangulation consists
in a subdivision of the space into irregular tetrahedra,
with four edges of length $\frac{\sqrt6}{2}$ and two of length $\sqrt2$,
see Figure \ref{fig:bcc}.
Such an asymmetry in the definition of nearest neighbours leads to consider an anisotropic energy, see \eqref{bcc-energy}.
\subsection{DC lattice}
Finally, we present the diamond cubic lattice, which is composed
of two interpenetrating face-centred cubic lattices (thus, it is non-Bravais).
It is relevant in applications to nanowires, since it is the structure
of materials of use, such as silicon and germanium \cite{Kavanagh}.
When the sites of the two interpenetrating lattices are filled with
two different species of atoms, the structure is called zincblende
and is typical of Gallium arsenide (GaAs) and Indium arsenide (InAs),
also used in technical applications to semiconductor optoelectronics \cite{Kavanagh}.
\par
The diamond cubic structure is defined by
$$
\L\dc{}:=\{\u_i\dc{} + \xi_1\mathrm{v}\dc{}_1+\xi_2\mathrm{v}\dc{}_2+\xi_3 \mathrm{v}\dc{}_3 \colon \ \xi_1,\xi_2,\xi_3\in{\mathbb Z} \,, \ i=1,2\} \,,
$$
where $\mathrm{v}_j\dc{}:=\mathrm{v}_j\fcc{}$, $j=1,2,3$, are as in the face-centred cubic and
$$
\u_1\dc{}:=(0,0,0)\,,\
\u_2\dc{}:=\sqrt2\big(\tfrac14,\tfrac14,\tfrac14\big)
$$
compose the basis. It is convenient to split the lattice as follows,
\begin{equation}\label{def:subdc}
\L\dc{}= \L\dc{_1} \cup \L\dc{_2} \,,
\end{equation}
$$
\L\dc{_i}:=\{\u_i\dc{} + \xi_1\mathrm{v}\dc{}_1+\xi_2\mathrm{v}\dc{}_2+\xi_3 \mathrm{v}\dc{}_3 \colon \ \xi_1,\xi_2,\xi_3\in{\mathbb Z}\} \,, \ i=1,2 \,,
$$
where the sublattices $\L\dc{_i}$, $i=1,2$, are face-centred cubic,
see Figure \ref{fig:zincblende}.
Each atom of the sublattice $x\in\L\dc{_i}, i=1,2$, has four nearest neighbours
at distance $\frac{\sqrt6}4$, all belonging to the sublattice $\L\dc{_j} , j\neq i$.
Such bonds are not enough to provide a rigid tessellation of the space. Therefore we need
to take into account also the next-to-nearest neighbours.
By Definition \ref{def:NNN}, the next-to-nearest neighbours of $x$ in $\L\dc{}$
turn out to be its nearest neighbours as an element of $\L\dc{_i}$.
More precisely, each atom $x$ lies at the barycentre of a tetrahedron whose vertices are the nearest neighbours of $x$; the edges of such a tetrahedron correspond to next-to-nearest bonds.
Thus, when next-to-nearest neighbours are considered,
$\L\dc{}$ inherits some rigid structure from the (face-centred cubic) sublattices $\L\dc{_i}$, $i=1,2$.
\par
For a better understanding of the diamond cubic lattice,
we also refer to the simpler example of the planar honeycomb lattice,
which can be treated by the same methods as presented here. This two-dimensional example contains the main ideas for treating non-Bravais lattices
with next and next-to-nearest neighbours, see Figure \ref{fig:honey}.
\begin{figure}[p]
\centering
\includegraphics[width=.35\textwidth]{zincblende1.eps}
\caption{Cubic cell in the diamond lattice $\L\dc{}$.
Atoms from the sublattice $\L\dc{}_1$ are represented in black/grey, while white atoms are from the sublattice $\L\dc{}_2$.
Nearest-neighbour bonds are displayed by solid thick lines.
Moreover, the picture shows a tetrahedron from the Delaunay pretriangulation of $\L\dc{}_1$:
its edges (solid and dashed thin lines) correspond to next-to-nearest neighbours in $\L\dc{}$.
A white atom lies at the barycentre of the tetrahedron, which is further divided
into four irregular tetrahedra by the bonds between the barycentre and each vertex.
}\label{fig:zincblende}
\end{figure}
\begin{figure}[p]
\centering
\includegraphics[width=.85\textwidth]{honey.eps}
\caption{Bonds and triangulation in a honeycomb lattice.
The lattice is given by
$\L^{\text{\davidsstar}}} % ^{\text{\varhexagon}:=\L^{\text{\davidsstar}}} % ^{\text{\varhexagon}_1 \cup \L^{\text{\davidsstar}}} % ^{\text{\varhexagon}_2$, where
$\L^{\text{\davidsstar}}} % ^{\text{\varhexagon}_i:=\{\u_i^{\text{\davidsstar}}} % ^{\text{\varhexagon} + \xi_1\mathrm{v}^{\text{\davidsstar}}} % ^{\text{\varhexagon}_1+\xi_2\mathrm{v}^{\text{\davidsstar}}} % ^{\text{\varhexagon}_2 \colon \ \xi_1,\xi_2 \in{\mathbb Z}\}$,
$\mathrm{v}^{\text{\davidsstar}}} % ^{\text{\varhexagon}_1:=(1,0)$, $\mathrm{v}^{\text{\davidsstar}}} % ^{\text{\varhexagon}_2:=(\frac12,\frac{\sqrt3}2)$,
$\u^{\text{\davidsstar}}} % ^{\text{\varhexagon}_1:=(0,0)$, $\u^{\text{\davidsstar}}} % ^{\text{\varhexagon}_2:=(0,\frac{\sqrt3}3)$.
This results into two interpenetrating sublattices $\L^{\text{\davidsstar}}} % ^{\text{\varhexagon}_1$ and $\L^{\text{\davidsstar}}} % ^{\text{\varhexagon}_2$,
both being hexagonal (i.e., equilateral triangular).
Atoms from $\L^{\text{\davidsstar}}} % ^{\text{\varhexagon}_1$ and $\L^{\text{\davidsstar}}} % ^{\text{\varhexagon}_2$ are displayed in different colors
in the picture, respectively in black and in white.
In the left part of the figure we indicate nearest neighbour (solid)
and next-to-nearest neighbour bonds (dashed lines).
The right part of the figure shows a possible triangulation,
that is the natural triangulation of $\L^{\text{\davidsstar}}} % ^{\text{\varhexagon}_1$
enriched by considering the nearest-neighbour bonds between atoms $x\in\L^{\text{\davidsstar}}} % ^{\text{\varhexagon}_1$ and $y\in\L^{\text{\davidsstar}}} % ^{\text{\varhexagon}_2$.
This corresponds to ignoring the bonds between atoms of $\L^{\text{\davidsstar}}} % ^{\text{\varhexagon}_2$, cf.\ Section~\ref{subsec:admiss}.
}
\label{fig:honey}
\end{figure}
\section{Setting of the model}\label{sec:not}
In order to mathematically describe the three-dimen\-sional heterostructured nanowires we introduce four parameters $\varepsilon$, $k$, $\lambda$ and $\rho$, next to the lattice structures discussed above.
\par
The parameter $\varepsilon>0$ scales the equilibrium lattice distances and allows considering a passage from the discrete to the continuous setting by letting $\varepsilon\to 0^{+}$. The parameter $k\in{\mathbb N}$, $k\ge1$, mimics the thickness of the nanowire. The shape of the nanowire in the discrete setting is a parallelepiped of length $2L$, $L>0$, and width and the height $k\varepsilon$, see Section~\ref{sec:refconfintene} for details. In the continuum limit $\varepsilon\to 0^{+}$, the length is conserved whereas the width and height tend to zero thus giving a dimension reduction of the system from three to one dimension. Still, the microscopic parameter $k$ has an impact on the continuum energy, which then allows investigating the limiting behaviour in dependence of the microscopic thickness $k$ of the wire.
\par
The parameters $\lambda$ and $\rho$ allow modeling the microscopic biphase structure of the nanowire. Here, $\lambda\in(0,1)$ denotes the ratio of the equilibrium distances in the deformed configuration of the material on the right hand side of the interface and of the material on the left hand side of the interface, see Section~\ref{subsec:setting} for details.
\par
The parameter $\rho\in(0,1]$ gives the ratio of the lattice distances of the two parts of the material in the reference configuration, where $\rho\in[\lambda,1]$ is the most interesting case. This allows treating different geometries of the nearest neighbours and in particular for dislocations.
The case of a defect-free body is modeled by $\rho=1$; the coordination number, i.e., the number of nearest neighbours of any internal atom, is constant in the lattice. If the crystal contains dislocations in the reference configuration, the coordination number is not constant. As we will show, this is the case for $\rho\neq 1$ and $k$ sufficiently large.
\subsection{Biphase lattices and rigid tessellations}
\label{subsec:setting}
Given $\rho\in(0,1]$ and vectors $\mathrm{v}_1,\mathrm{v}_2,\mathrm{v}_3,$ $\u_1,\u_2\in{\mathbb R}^3$,
we define the biphase atomistic lattice
\begin{equation}
\label{def:biphase}
\L_\rho:= \L_1^-\cup\L_\rho^+
\end{equation}
by juxtaposing the two lattices
$\L_1^-$ and $\L_\rho^+$ given by
\begin{align*}
\L_1^- &:= \{\u_i + \xi_1\mathrm{v}_1+\xi_2\mathrm{v}_2+\xi_3\mathrm{v}_3 \colon \ \xi_1,\xi_2,\xi_3\in{\mathbb Z} \,, \ i=1,2 \,, \ \xi_1<0 \} \,, \\
\L_\rho^+ &:= \{\rho\u_i + \xi_1\mathrm{v}_1+\xi_2\mathrm{v}_2+\xi_3\mathrm{v}_3 \colon \ \xi_1,\xi_2,\xi_3\in\rho{\mathbb Z} \,, \ i=1,2 \,, \ \xi_1\ge0 \} \,.
\end{align*}
We will apply the above definitions to the crystals introduced in Section \ref{3d} and denote by
$\L_\rho\fcc{}$, $\L_\rho\hcp{}$, $\L_\rho\bcc{}$, and $\L_\rho\dc{}$
the lattices obtained by taking the vectors
$\{\mathrm{v}_j\fcc{},\u_i\fcc{}\}_{\substack{j=1,2,3\\i=1,2\,\,\,}}$,
$\{\mathrm{v}_j\hcp{},\u_i\hcp{}\}_{\substack{j=1,2,3\\i=1,2\,\,\,}}$,
$\{\mathrm{v}_j\bcc{},\u_i\bcc{}\}_{\substack{j=1,2,3\\i=1,2\,\,\,}}$, and
$\{\mathrm{v}_j\dc{},\u_i\dc{}\}_{\substack{j=1,2,3\\i=1,2\,\,\,}}$,
respectively, where $\u_i\fcc{}=u_i\bcc{}:=0$.
\par
In each of the four cases we find similar structures
for the planes at the interface between the lattices $\L_1^-$ and $\L_\rho^+$.
More precisely, for the face-centred cubic and the hexagonal close-packed,
the interfacial planes are two-dimensional equilateral triangular Bravais lattices, see Figure \ref{fig:fcc-int}.
In the body-centred cubic, the interfacial planes are triangular Bravais lattices, but not equilateral,
since the distance between nearest neighbours is not constant.
Finally, in the diamond cubic, whose properties are similar to the face-centred cubic,
we also find equilateral triangular planes composed by atoms of one of the sublattices.
(For a lower dimensional idea, see Figure \ref{fig:honey2}.)
\par
\begin{figure}[p]
\vspace{1cm}
\centering
\subfloat[]{
\includegraphics[height=.28\textwidth]{interface21.eps}
\label{fig:interface21}
}
\hfill
\subfloat[]{
\includegraphics[height=.28\textwidth]{interface31.eps}
\label{fig:interface31}
}
\caption{By cutting a cubic lattice along certain transverse planes,
one finds two-dimensional hexagonal Bravais lattices.
(a) face-centred; (b) body-centred.
}
\label{fig:fcc-int}
\end{figure}
\begin{figure}[p]
\vspace{1.5cm}
\centering
\includegraphics[width=.85\textwidth]{honey2.eps}
\caption{Dislocations in a honeycomb-type lattice.
The bonds at the interface are chosen in the following way:
First one considers only black atoms and finds a Delaunay pretriangulation,
which is then refined to a triangulation (dashed lines);
the same is done for white atoms (dotted lines).
The dashed and dotted lines thus obtained give the bonds between next-to-nearest neighbours.
Finally, each white (resp.\ black) atom lying inside a triangle formed by three black (resp.\ white) atoms
is connected to the vertices of that triangle by nearest-neighbour bonds (solid lines).
}
\label{fig:honey2}
\end{figure}
\clearpage
Next we define the interfacial bonds in the case when $\rho\neq1$.
Following the idea already used in the regular parts of the lattices,
we consider the (unique) Delaunay pretriangulation $\mathcal T'_\rho$ of $\L_\rho$
(Definition \ref{def:Del-pre}).
This defines,
in the case of $\L_\rho\fcc{}$, $\L_\rho\hcp{}$, and $\L_\rho\bcc{}$, a tessellation of the space into
rigid polyhedra away from the interface.
At the interface, the partition $\mathcal T'_\rho$ may contain polyhedra with quadrilateral facets
(to see this, one should recall that the interfacial atoms lie on two parallel planes
consisting of two-dimensional triangular Bravais lattices, with parallel primitive vectors):
in such a case we refine $\mathcal T'_\rho$ further, in order to obtain rigid polyhedra.
More precisely, given a quadrilateral facet we introduce a further bond along a diagonal of the facet;
correspondingly, the region around the interface is subdivided into (irregular) tetrahedra and octahedra.
\par
Following this construction we define a partition of the space into rigid polyhedra
and call it the \emph{rigid Delaunay tessellation} associated to $\L_\rho$, denoted by $\mathcal T_\rho$.
The nearest neighbours are the extrema of the edges of the polyhedra of the subdivision.
Such procedure can be followed for $\L_\rho\fcc{}$, $\L_\rho\hcp{}$, and $\L_\rho\bcc{}$.
Instead, in the diamond cubic lattice, applying Definition \ref{def:NN} may result in nearest-neighbour
bonds between interfacial atoms of the same sublattice
(which should instead be next-to-nearest neighbours).
This would not be consistent with the structure defined away from the interface,
therefore we follow a different construction.
\par
Recall that $\L_\rho\dc{}$ consists of two interpenetrating face-centred cubic lattices
$\L_{\rho}\dc{_1}$ and $\L_{\rho}\dc{_2}$, see \eqref{def:subdc} and \eqref{def:biphase}.
We introduce Delaunay pretriangulations for the sublattices $\L_{\rho}\dc{_1}$ and
$\L_{\rho}\dc{_2}$ individually,
which we further refine in order to obtain triangular facets
as before; we say that the vertices of the resulting edges are next-to-nearest neighbours
in $\L_{\rho}\dc{}$.
The tessellation of $\L_{\rho}\dc{_i}$ consists of (possibly irregular) tetrahedra and octahedra;
some of them may contain one atom $x$ of the other sublattice $\L_{\rho}\dc{_j}$,
$j\neq i$. In this case,
we connect $x$ to the vertices of the surrounding polyhedron and say that each of those vertices
is a nearest neighbour for $x$.
When applied to the regular parts of the lattice, this construction is consistent with the notion
of nearest and next-to-nearest neighbours presented in the previous section.
(For a simpler idea about the resulting structure, we refer to Figure \ref{fig:honey2} in the case of a honeycomb-type lattice.)
\par
\subsection{Reference configurations and interaction energies} \label{sec:refconfintene}
We now pass to rescaled, bounded lattices.
Given $L>0$, $\varepsilon\in(0,1]$, and $k\in{\mathbb N}$, we define
$$
\L_{\rho,\varepsilon}(k):=(\varepsilon\L_\rho)\cap\overline\Omega_{k\varepsilon}\,,
$$
where
$$
\Omega_{k\varepsilon} := \{ \xi_1\mathrm{v}_1+\xi_2\mathrm{v}_2+\xi_3\mathrm{v}_3\colon \xi_1\in(-L,L)\,,\ \xi_2,\xi_3\in(0, k\varepsilon) \}
$$
and $\overline\Omega_{k\varepsilon}$ is the union of all (closed) polyhedra of $\mathcal T_{\rho,\varepsilon}$ that intersect $\Omega_{k\varepsilon}$ on a nonempty set
(see \eqref{uffa} for the definition of $\mathcal T_{\rho,\varepsilon}$). Notice that the lattices $\L_{\rho,\varepsilon}\fcc{}$, $\L_{\rho,\varepsilon}\hcp{}$, $\L_{\rho,\varepsilon}\bcc{}$, and $\L_{\rho,\varepsilon}\dc{}$ denote the corresponding rescaled, bounded lattices for the other crystal structures, with the lattice vectors chosen as above.
We set
\begin{align*}
\displaybreak[0]
\Omega_{k\varepsilon}^- &:= \{ \xi_1\mathrm{v}_1+\xi_2\mathrm{v}_2+\xi_3\mathrm{v}_3\colon \xi_1\in(-L,0)\,,\ \xi_2,\xi_3\in(0, k\varepsilon) \} \,,\\
\displaybreak[0]
\Omega_{k\varepsilon}^+ &:= \{ \xi_1\mathrm{v}_1+\xi_2\mathrm{v}_2+\xi_3\mathrm{v}_3\colon \xi_1\in(0,L)\,,\ \xi_2,\xi_3\in(0, k\varepsilon) \} \,,\\
\displaybreak[0]
\L_{1,\varepsilon}^-(k) &:= \{\varepsilon \u_i + \xi_1\mathrm{v}_1+\xi_2\mathrm{v}_2+\xi_3\mathrm{v}_3\in\L_{\rho,\varepsilon}(k)\colon \xi_1<0\} \,,\\
\displaybreak[0]
\L_{\rho,\varepsilon}^+(k) &:= \{\rho\varepsilon\u_i + \xi_1\mathrm{v}_1+\xi_2\mathrm{v}_2+\xi_3\mathrm{v}_3\in\L_{\rho,\varepsilon}(k)\colon \xi_1\ge0\} \,.
\end{align*}
Two points $x,y\in \L_{\rho,\varepsilon}$
are said to be nearest (resp., next-to-nearest) neighbours if $x/\varepsilon$, $y/\varepsilon$ fulfil the corresponding property in the lattice $\L_\rho$.
This definition applies to each of the four cases presented above. Notice that we denote these lattices by $\L_{1,\varepsilon}\fcc{^\pm}(k)$, $\L_{1,\varepsilon}\hcp{^\pm}(k)$, $\L_{1,\varepsilon}\bcc{^\pm}(k)$ and $\L_{1,\varepsilon}\dc{^\pm}(k)$ as well as by $\L_{\rho,\varepsilon}\fcc{^\pm}(k)$, $\L_{\rho,\varepsilon}\hcp{^\pm}(k)$, $\L_{\rho,\varepsilon}\bcc{^\pm}(k)$ and $\L_{\rho,\varepsilon}\dc{^\pm}(k)$, respectively.
\par
We have introduced so far the bonds that enter the definition of the energy, which we generally
denote by ${\mathcal E}_{\varepsilon}^{\lambda}$.
Next we specialise ${\mathcal E}_{\varepsilon}^{\lambda}$ for each of the four lattices introduced above.
In the cases of the face-centred cubic and of the hexagonal close-packed,
the total interaction energy is defined respectively by
\begin{equation}\label{en:fcc}
\begin{split}
{\mathcal E}_{\varepsilon}\fcc{,\,\lambda}(u_\varepsilon,\rho,k) := &\
\tfrac{1}{2}\!\!\! \sum_{\substack{\NN{x}{y}\\x\in \L_{1,\varepsilon}\fcc{^-}(k) \\y\in\L_{\rho,\varepsilon}\fcc{}(k)}}
\left(\Big|\frac{u_\varepsilon(x)-u_\varepsilon(y)}{\varepsilon}\Big|-1\right)^2
\\ &
+
\tfrac{1}{2}\!\!\! \sum_{\substack{\NN{x}{y}\\x\in \L_{\rho,\varepsilon}\fcc{^+}(k)\\y\in\L_{\rho,\varepsilon}\fcc{}(k)}}
\left(\Big|\frac{u_\varepsilon(x)-u_\varepsilon(y)}{\varepsilon}\Big|-\lambda\right)^2
\end{split}
\end{equation}
for every deformation $u_\varepsilon\colon\L\fcc{}_{\rho,\varepsilon}(k)\to{\mathbb R}^3$
and by
\begin{equation}\label{en:hcp}
\begin{split}
{\mathcal E}_{\varepsilon}\hcp{,\,\lambda}(u_\varepsilon,\rho,k) := &\
\tfrac{1}{2}\!\!\! \sum_{\substack{\NN{x}{y}\\x\in \L_{1,\varepsilon}\hcp{^-}(k) \\y\in\L_{\rho,\varepsilon}\hcp{}(k)}}
\left(\Big|\frac{u_\varepsilon(x)-u_\varepsilon(y)}{\varepsilon}\Big|-1\right)^2
\\ &
+
\tfrac{1}{2}\!\!\! \sum_{\substack{\NN{x}{y}\\x\in \L_{\rho,\varepsilon}\hcp{^+}(k)\\y\in\L_{\rho,\varepsilon}\hcp{}(k)}}
\left(\Big|\frac{u_\varepsilon(x)-u_\varepsilon(y)}{\varepsilon}\Big|-\lambda\right)^2
\end{split}
\end{equation}
for every deformation $u_\varepsilon\colon\L\hcp{}_{\rho,\varepsilon}(k)\to{\mathbb R}^3$.
\par
For the body-centred cubic, we need to use an anisotropic energy,
because of the different length of the bonds between nearest neighbours
in the reference configuration.
For every deformation $u_\varepsilon\colon\L\bcc{}_{\rho,\varepsilon}(k)\to{\mathbb R}^3$ we define
\begin{equation}\label{bcc-energy}
\begin{split}
{\mathcal E}_{\varepsilon}\bcc{,\,\lambda}(u_\varepsilon,\rho,k) := &\
\tfrac{1}{2}\!\!\! \sum_{\substack{\NN{x}{y}\\x\in \L_{1,\varepsilon}\bcc{^-}(k) \\y\in\L_{\rho,\varepsilon}\bcc{}(k)}}
\left(\Big|\frac{u_\varepsilon(x)-u_\varepsilon(y)}{\varepsilon}\Big|-\varphi^x_y\right)^2
\\ &
+
\tfrac{1}{2}\!\!\! \sum_{\substack{\NN{x}{y}\\x\in \L_{\rho,\varepsilon}\bcc{^+}(k)\\y\in\L_{\rho,\varepsilon}\bcc{}(k)}}
\left(\Big|\frac{u_\varepsilon(x)-u_\varepsilon(y)}{\varepsilon}\Big|-\lambda\,\varphi^x_y\right)^2 \,,
\end{split}
\end{equation}
where $\varphi^x_y:=\varphi(\frac{x-y}{\mod{x-y}})$ and
$\varphi\colon \mathbb{S}^2\to(0,+\infty)$ is a smooth function such that
\begin{alignat*}{2}
\varphi(\mathrm{v})&=\sqrt2 \quad&&\text{if } \mathrm{v}\in\{(\pm1,0,0),(0,\pm1,0),(0,0,\pm1)\} \,, \\
\varphi(\mathrm{v})&=\tfrac{\sqrt6}{2} \quad&&\text{if } \sqrt3\,\mathrm{v}\in\{(\pm1,1,1), (\pm1,-1,1), (\pm1,1,-1), (\pm1,-1,-1)\}.
\end{alignat*}
\par
Finally, recall that the diamond cubic lattice consists of two interpenetrating face-centred cubic lattices. Therefore we set for a deformation $u_\varepsilon\colon\L\dc{}_{\rho,\varepsilon}(k)\to{\mathbb R}^3$
\begin{equation*}
{\mathcal E}_{\varepsilon}\dc{,\,\lambda}(u_\varepsilon,\rho,k) :=
c_1 {\mathcal E}\dc{_1}_{{\rm NNN}} + c_2 {\mathcal E}\dc{_2}_{{\rm NNN}} + {\mathcal E}\dc{}_{{\rm NN}} \,,
\end{equation*}
where the first two summands account for next-to-nearest neighbour interactions and are defined as in \eqref{en:fcc}, namely
\begin{equation*}
{\mathcal E}\dc{_i}_{{\rm NNN}} :=
\tfrac{1}{2}\!\!\! \sum_{\substack{\NNN{x}{y}\\x\in \L_{1,\varepsilon}\dc{_i^-}(k) \\y\in\L_{\rho,\varepsilon}\dc{}(k)}}
\left(\Big|\frac{u_\varepsilon(x)-u_\varepsilon(y)}{\varepsilon}\Big|-1\right)^2 +
\tfrac{1}{2}\!\!\! \sum_{\substack{\NNN{x}{y}\\x\in \L_{\rho,\varepsilon}\dc{_i^+}(k)\\y\in\L_{\rho,\varepsilon}\dc{}(k)}}
\left(\Big|\frac{u_\varepsilon(x)-u_\varepsilon(y)}{\varepsilon}\Big|-\lambda\right)^2 \,,
\end{equation*}
while the last term is
\begin{equation*}
{\mathcal E}\dc{}_{{\rm NN}} :=
\tfrac{1}{2}\!\!\! \sum_{\substack{\NN{x}{y}\\x\in \L_{1,\varepsilon}\dc{^-}(k) \\y\in\L_{\rho,\varepsilon}\dc{}(k)}}
\left(\Big|\frac{u_\varepsilon(x)-u_\varepsilon(y)}{\varepsilon}\Big|-\tfrac{\sqrt6}{4}\right)^2 +
\tfrac{1}{2}\!\!\! \sum_{\substack{\NN{x}{y}\\x\in \L_{\rho,\varepsilon}\dc{^+}(k)\\y\in\L_{\rho,\varepsilon}\dc{}(k)}}
\left(\Big|\frac{u_\varepsilon(x)-u_\varepsilon(y)}{\varepsilon}\Big|-\lambda\tfrac{\sqrt6}{4}\right)^2 \,.
\end{equation*}
The choice of the constants $c_1,c_2>0$ determines how strong the interactions between
atoms of the same sublattice $\L\dc{_i}$ are.
\par
\subsection{Admissible configurations}
\label{subsec:admiss}
In order to define the admissible deformations, we introduce piecewise affine functions.
To this end, we need to refine $\mathcal T_\rho$ to a proper triangulation.
However, we do not change the definition of the nearest neighbours, i.e., we do not introduce new interactions in the energy.
\begin{remark}
For the reader's convenience, we summarise here the different tessellations of the space associated to a biphase discrete lattice $\L_\rho$,
adopted in our setting.
\begin{itemize}
\item We have started from the (unique) \emph{Delaunay pretriangulation} $\mathcal T'_\rho$ (Definition \ref{def:Del-pre}),
which may contain non-rigid polyhedra at the interface.
\item We have refined $\mathcal T'_\rho$, obtaining a \emph{rigid Delaunay tessellation} $\mathcal T_\rho$,
a partition of the space into (possibly irregular) tetrahedra and octahedra.
Such a tessellation is not unique, indeed we have chosen a diagonal
for each quadrilateral facet of polyhedra of $\mathcal T'_\rho$.
The corresponding bonds enter the definition of the interaction energy.
\item In order to work with piecewise affine functions, in this section
we further refine $\mathcal T_\rho$ to get three possible \emph{triangulations}
(i.e., subdivisions of the space into tetrahedra only), denoted by
$\mathcal T_\rho^{(1)}$, $\mathcal T_\rho^{(2)}$, and $\mathcal T_\rho^{(3)}$, respectively.
\end{itemize}
The above construction is used to work in the case of
$\L_\rho\fcc{}$, $\L_\rho\hcp{}$, and $\L_\rho\bcc{}$.
For $\L_\rho\dc{}$, the definition of $\mathcal T_\rho$ is different,
as made precise in Sections \ref{subsec:setting} and \ref{subsec:admiss}.
\end{remark}
In the case of $\L_\rho\fcc{}$, $\L_\rho\hcp{}$, and $\L_\rho\bcc{}$,
given a (possibly irregular) octahedron of $\mathcal T_\rho$, we divide it into four irregular tetrahedra
by cutting it along one of the three diagonals.
We choose the diagonal starting from the vertex with the largest $x_1$-coordinate;
if two or three vertices have the same largest $x_1$-coordinate, we take among them the point with largest $x_2$-coordinate;
if two of such vertices have also the same largest $x_2$-coordinate, we take the one with the largest $x_3$-coordinate.
By repeating the process on every octahedron of $\mathcal T_\rho$, we obtain a triangulation that we denote by $\mathcal T_\rho^{(1)}$.
Other two triangulations $\mathcal T_\rho^{(2)}$ and $\mathcal T_\rho^{(3)}$ are obtained by repeating the same procedure,
but with different ordering of the indices, namely $x_2,x_3,x_1$ and $x_3,x_1,x_2$ respectively.
\par
In the case of the diamond-cubic lattice, we define a triangulation as follows:
we consider the Delaunay pretriangulation of $\L\dc{_1}$, which is rigid.
As already observed, some of the tetrahedra of the latter pretriangulation contain
an atom of $\L\dc{_2}$
at the barycentre (more precisely, every other tetrahedron has this property,
see Figure \ref{fig:zincblende}).
Such tetrahedra are further subdivided by connecting
the barycentre to the vertices.
In other words, we define a tessellation into tetrahedra and octahedra by considering
the (nearest neighbour) interactions between atoms
$x\in\L\dc{_1}$ and $y\in\L\dc{_2}$,
as well as the interactions between atoms of $\L\dc{_1}$
(nearest neighbour if restricted to $\L\dc{_1}$,
next-to-nearest neighbour if viewed in the whole $\L\dc{}$),
and ignoring the interactions between atoms of $\L\dc{_2}$.
We apply the same rule to the biphase lattice $\L_\rho\dc{}$ and
further subdivide the resulting octahedra as done for $\L_\rho\fcc{}$, $\L_\rho\hcp{}$,
and $\L_\rho\bcc{}$,
obtaining three possible triangulations.
For a better
understanding we illustrate the tessellation thus defined in the simpler case
of the honeycomb lattice in Figure \ref{fig:honey}.
\par
Given a function $u\colon \L_\rho\to{\mathbb R}^3$, we denote by $u^{(1)}$, $u^{(2)}$, and $u^{(3)}$ its piecewise affine interpolations
with respect to the triangulations $\mathcal T_\rho^{(1)}$, $\mathcal T_\rho^{(2)}$, and $\mathcal T_\rho^{(3)}$, respectively.
Analogous definitions and notations hold for the rescaled bounded lattices of the type $\L_{\rho,\varepsilon}(k)$.
More precisely, define
\begin{equation}\label{uffa}
\mathcal T_{\rho,\varepsilon}:=\{\varepsilon T\colon T\in\mathcal T_\rho\} \quad \text{and} \quad \mathcal T_{\rho,\varepsilon}^{(i)}:=\{\varepsilon T\colon T\in\mathcal T_\rho^{(i)}\}
\end{equation}
for $i=1,2,3$.
The set of admissible deformations is
\begin{equation}\label{ad-3d}
\begin{split}
{\mathcal A}_{\rho,\varepsilon}(\Omega_{k\varepsilon}):= \big\{ u_\varepsilon\in C^0(\overline\Omega_{k\varepsilon};{\mathbb R}^3) \colon & u_\varepsilon \ \text{piecewise affine,}\\
& \nabla} %{\mathrm{D} u_\varepsilon \ \text{constant on}\ \Omega_{k\varepsilon}\cap T\quad \forall\, T\in\mathcal T_{\rho,\varepsilon}^{(1)}\,, \\
& \det \nabla} %{\mathrm{D} u_\varepsilon>0 \ \text{a.e.\ in}\ \Omega_{k\varepsilon}\,, \\
& u_\varepsilon(\mathcal P) \ \text{is convex}\ \forall\, \mathcal P\in\mathcal T_{\rho,\varepsilon} \big\} \,.
\end{split}
\end{equation}
The restriction of $u_\varepsilon\in{\mathcal A}_{\rho,\varepsilon}(\Omega_{k\varepsilon})$ to $\L_{\rho,\varepsilon}(k)$ is still denoted by $u_\varepsilon$.
We will see that the limiting functional is independent of the choice of the triangulation $\mathcal T_{\rho,\varepsilon}^{(1)}$ in \eqref{ad-3d}, cf.\ Remark~\ref{rmk5}.
\par
\begin{remark}\label{rmk:3d}
The assumption of convexity on the images of the octahedra of $\mathcal T_{\rho,\varepsilon}$ is needed to enforce rigidity:
without such an assumption an octahedron could be compressed without paying any energy.
On the other hand, the notion of non-interpenetration used in \eqref{ad-3d} is independent of the choice of the triangulation $\mathcal T_{\rho,\varepsilon}^{(1)}$
provided the image of each octahedron is assumed to be convex, as clarified by Lemma \ref{lemma:convex}.
\end{remark}
It will be convenient to introduce
\begin{equation*}
\Omega_{k,\infty}:=\,
\{ \xi_1\mathrm{v}_1+\xi_2\mathrm{v}_2+\xi_3\mathrm{v}_3\colon \xi_1\in(-\infty,+\infty)\,,\ \xi_2,\xi_3\in(0, k) \}
\end{equation*}
and to denote by $\overline\Omega_{k,\infty}$ the union of all (closed) polyhedra of $\mathcal T_{\rho}$ that have a nonempty intersection with $\Omega_{k,\infty}$.
We define the set of admissible deformations on the rescaled infinite domain
as
\begin{equation*}
\begin{split}
{\mathcal A}_{\rho,\infty}(\Omega_{k,\infty}):= \big\{ u\in C^0(\overline\Omega_{k,\infty};{\mathbb R}^3) \colon & u \ \text{piecewise affine,}\\
& \nabla} %{\mathrm{D} u \ \text{constant on}\ \Omega_{k,\infty}\cap T\ \forall\, T\in\mathcal T_{\rho}\,, \\
& \det \nabla} %{\mathrm{D} u>0 \ \text{a.e.\ in}\ \Omega_{k,\infty}\,, \\
&u (\mathcal P) \ \text{is convex}\ \forall\, \mathcal P\in\mathcal T_{\rho} \big\} \,.
\end{split}
\end{equation*}
All definitions apply to each of the four cases presented above.
Correspondingly, we define the energy on the rescaled infinite domain and denote it by
${\mathcal E}_{\infty}^\lambda$.
Specifically, given a discrete deformation $v$ of the face-centred cubic lattice, ${\mathcal E}_{\infty}^\lambda$ is defined by
\begin{equation}\label{energia-infinita}
{\mathcal E}_{\infty}\fcc{,\,\lambda}(v,\rho,k) :=
\tfrac{1}{2}\!\!\! \sum_{\substack{\NN{x}{y}\\x\in \L_{1}\fcc{^-}(k) \\y\in\L_{\rho}\fcc{}(k)\\x,y\in\overline\Omega_{k,\infty}}}
\left(\big| v(x)-v(y) \big|-1\right)^2 +
\tfrac{1}{2}\!\!\! \sum_{\substack{\NN{x}{y}\\x\in \L_{\rho}\fcc{^+}(k)\\y\in\L_{\rho}\fcc{}(k)\\x,y\in\overline\Omega_{k,\infty}}}
\left(\big| v(x)-v(y) \big|-\lambda\right)^2 \,.
\end{equation}
Analogous definitions hold for $\L_\rho\hcp{}$, $\L_\rho\bcc{}$ and
$\L_\rho\dc{}$.
\section{Discrete rigidity in dimension three}\label{sec:rigidity}
A key tool in the analysis developed in \cite{LaPaSc2014} for two-dimensional heterogeneous nanowires as well as in the analysis of the three-dimensional setting is the following rigidity estimate.
\begin{theorem}\label{thm-rigidity}
\cite[Theorem 3.1]{fjm}
Let $N\geq 2$, and let $1< p < +\infty$.
Suppose that $U\subset{\mathbb R}^{N}$ is a bounded Lipschitz domain.
Then there exists a constant $C=C(U)$
such that for each $u\in W^{1,p}(U;{\mathbb R}^{N})$
there exists a constant matrix $R\in SO(N)$ such that
\begin{equation}\label{rigidity}
\|\nabla} %{\mathrm{D} u-R\|_{L^{p}(U;\M^{N\times N})} \leq C(U)
\|\mathrm{dist}(\nabla} %{\mathrm{D} u,SO(N))\|_{L^{p}(U)}\,.
\end{equation}
The constant $C(U)$ is invariant under dilation and translation of the domain.
\end{theorem}
In order to employ the above result, we need the discrete rigidity estimates of Lemmas \ref{lemma:tetra} and \ref{lemma:ottaedro}, which state that the energy of a lattice cell is bounded from below
by the distance of the deformation gradient from the set of rotations.
Similar rigidity estimates are used in \cite{BSV,FlaThe,Schm06,Th06}.
\par
We use the following notation for the vectors determined by the edges of the regular tetrahedron $\mathcal S$ of edge length one:
$\mathrm{w}_1:=(1,0,0)$, $\mathrm{w}_2:=(\frac{1}{2},\frac{\sqrt{3}}{2},0)$, $\mathrm{w}_3:=\mathrm{w}_2-\mathrm{w}_1$,
$\mathrm{w}_4:=(\frac{1}{2},\frac{\sqrt{3}}{6},\frac{\sqrt{6}}{3})$,
$\mathrm{w}_5:=\mathrm{w}_4-\mathrm{w}_2$, and $\mathrm{w}_6:=\mathrm{w}_4-\mathrm{w}_1$, cf.\ Figure~\ref{tetraedro}.
\par
\begin{lemma}\label{lemma:tetra}
There exists $C>0$ such that
\begin{equation}\label{rig-tetraedro}
\mathrm{dist}^2(F,SO(3)) \leq C \sum_{i=1}^6( |F\mathrm{w}_i|-1)^2
\end{equation}
for every $F\in GL^+(3)$.
\end{lemma}
\begin{proof}
\begin{figure}
\centering
\psfrag{w1}{$\mathrm{w}_1$}
\psfrag{w2}{$\mathrm{w}_2$}
\psfrag{w3}{$\mathrm{w}_3$}
\psfrag{w4}{$\mathrm{w}_4$}
\psfrag{fw1}{$F\mathrm{w}_1$}
\psfrag{fw2}{$F\mathrm{w}_2$}
\psfrag{fw4}{$F\mathrm{w}_4$}
\includegraphics[width=.5\textwidth]{tetraedro.eps}
\caption{The tetrahedron $\mathcal S$ and its image $F(\mathcal S)$.}
\label{tetraedro}
\end{figure}
Set $\delta_i:= |F \mathrm{w}_i| -1$ and $\delta:=(\delta_1,\dots,\delta_6)$, then $\sum_{i=1}^6( |F\mathrm{w}_i|-1)^2 = \sum_{i=1}^6 \delta_i^2=|\delta|^2$.
Without loss of generality we may assume that
$$
F \mathrm{w}_1 =(1+\delta_1)\mathrm{w}_1\,, \quad
F\mathrm{w}_2 \in {\rm{span}}\{\mathrm{w}_1,\mathrm{w}_2\}\,,\quad
F\mathrm{w}_2 \cdot e_2 >0 \,,
$$
as in Figure \ref{tetraedro}. Notice that the above assumptions imply $F\mathrm{w}_4 \cdot e_3 >0$.
We have
\begin{align}\label{dist-est-2}
\nonumber
\mathrm{dist}^2(F,SO(3)) \leq |F-I|^2 & \leq C \big( | (F-I) \mathrm{w}_1 |^2 + |(F-I)\mathrm{w}_2 |^2 + |(F-I)\mathrm{w}_4 |^2 \big) \\
&=C \big(\delta_1^2 + | (F-I)\mathrm{w}_2 |^2 + |(F-I)\mathrm{w}_4 |^2 \big) \,.
\end{align}
By a simple geometric argument one finds
\begin{equation}\label{dist-estimate}
|(F-I)\mathrm{w}_2 |^2 = 1 + (1+\delta_2)^2 -2(1+\delta_2)\cos\big(\theta_{12} - \tfrac{\pi}{3}\big) \,,
\end{equation}
where $\theta_{12}$ is the angle (measured anticlockwise) between $\mathrm{w}_1$ and $F\mathrm{w}_2$,
which is determined by
\begin{equation}\label{dist-estimate2}
\cos\theta_{12} =
\frac{ (1+\delta_1)^2 + (1+\delta_2)^2 - (1+\delta_3)^2 }{ 2 (1+\delta_1)(1+\delta_2)}
\quad \text{and} \quad \sin\theta_{12} >0 \,,
\end{equation}
cf.\ \cite[Proof of Lemma~2.2]{LaPaSc2014}.
Notice that the condition $\sin\theta_{12}>0$ follows from the assumptions $F\in GL^+(3)$
and $F\mathrm{w}_2 \cdot e_2 >0 $.
\par
Denote by $\theta_{ij}$ the acute angle formed by $F\mathrm{w}_i$ and $F\mathrm{w}_j$ and by $\eta_{44}$
that between $F\mathrm{w}_4$ and $\mathrm{w}_4$.
Since
\begin{equation}\label{formula:F-I}
|(F-I)\mathrm{w}_4 |^2 = 1+ (1+\delta_4)^2 - 2 (1+\delta_4)\cos\eta_{44}\,,
\end{equation}
in order to express
the right hand side of \eqref{dist-est-2} in terms of the $\delta_i$'s, we need to specialize
$\cos\eta_{44}$ in terms of the $\delta_i$'s.
Set
$$
\frac{F\mathrm{w}_4}{|F\mathrm{w}_4|}: =(a_1,a_2,a_3) \,,
$$
and remark that, by assumption, $a_3 >0$.
Thus
\begin{equation}\label{tildecos44}
\cos\eta_{44}= \frac{F\mathrm{w}_4}{|F\mathrm{w}_4|} \cdot \mathrm{w}_4 =
\tfrac{1}{2}a_1 + \tfrac{\sqrt{3}}{6}a_2 + \tfrac{\sqrt{6}}{3}\sqrt{1- a_1^2 -a_2^2} \,.
\end{equation}
On the other hand $a_1$ and $a_2$ are computed by solving
\begin{align}\label{coordinate}
& a_1 = \frac{F\mathrm{w}_4}{|F\mathrm{w}_4|} \cdot \mathrm{w}_1 = \cos\theta_{14} \,, \\
&
a_1 \cos\theta_{12} + a_2 \sin\theta_{12} =
\frac{F\mathrm{w}_4}{|F\mathrm{w}_4|} \cdot \frac{F\mathrm{w}_2}{|F\mathrm{w}_2|} = \cos\theta_{24}\,,
\end{align}
where
\begin{equation}
\cos\theta_{14} =
\frac{ (1+\delta_1)^2 + (1+\delta_4)^2 - (1+\delta_6)^2 }{ 2 (1+\delta_1)(1+\delta_4)}\,,
\end{equation}
\begin{equation}\label{cos24}
\cos\theta_{24} =
\frac{ (1+\delta_2)^2 + (1+\delta_4)^2 - (1+\delta_5)^2 }{ 2 (1+\delta_2)(1+\delta_4)}\,.
\end{equation}
Taking into account \eqref{dist-estimate}--\eqref{cos24}, one can express the right hand side of
\eqref{dist-est-2} as a function $f$ of $\delta$ and see that $f(0)=0$ and $\nabla f(0)=0$,
which implies $f(\delta)\leq C|\delta|^2$ for $|\delta|$ sufficiently small.
For larger $|\delta|$ the inequality readily follows from \eqref{dist-est-2}--\eqref{formula:F-I}.
\end{proof}
We will consider the octahedron $\mathcal O$ generated by the points
$P_1:=(0,0,0)$, $P_2:=(1,0,0)$, $P_3:=(0,1,0)$, $P_4:=(1,1,0)$, $P_5:=(\frac12,\frac12,\frac{\sqrt{2}}{2})$,
and $P_6:=(\frac12,\frac12,-\frac{\sqrt{2}}{2})$, see Figure \ref{ottaedro}.
\begin{figure}
\centering
\psfrag{1}{$P_1$}
\psfrag{2}{$P_2$}
\psfrag{3}{$P_3$}
\psfrag{4}{$P_4$}
\psfrag{5}{$P_5$}
\psfrag{6}{$P_6$}
\includegraphics[width=.35\textwidth]{ottaedrop.eps}
\caption{The octahedron $\mathcal O$.}
\label{ottaedro}
\end{figure}
We call $\mathcal T^{(1)}$ the triangulation determined by cutting $\mathcal O$ along the diagonal $P_1P_4$, further $\mathcal T^{(2)}$ denotes the triangulation corresponding to $P_2P_3$, and $\mathcal T^{(3)}$ the one corresponding to $P_5P_6$.
\par
Given a deformation $u$ of the six vertices of $\mathcal O$,
$u^{(i)}$ denotes the piecewise affine extension of $u$ corresponding to
the triangulation $\mathcal T^{(i)}$, $i=1,2,3$.
In the next lemma, $\mathcal Q$ denotes the interior of the (bounded) polyhedron determined by the images of the facets of $\mathcal O$ through any of the piecewise affine extensions $u^{(i)}$.
(Notice that the images of the facets do not depend on the chosen extension.)
\begin{lemma}\label{lemma:convex}
One has that $\det \nabla} %{\mathrm{D} u^{(1)}>0$ a.e.\ in $\mathcal O$ and $\mathcal Q$ is convex if and only if $\det \nabla} %{\mathrm{D} u^{(i)}>0$ for every $i=1,2,3$.
\end{lemma}
\begin{proof}
Assume that $\det \nabla} %{\mathrm{D} u^{(1)} > 0$ a.e. This implies that $\mathcal Q$ is connected, the diagonal $u(P_1)u(P_4)$ is contained in $\mathcal Q$, and the outer normal vectors to the facets of $\mathcal Q$ point towards the outside of $\mathcal Q$.
Consider now the tetrahedra of the triangulation $\mathcal T^{(2)}$: since the normals to the facets of $\mathcal Q$ point towards the outside of $\mathcal Q$,
it turns out that $\det \nabla} %{\mathrm{D} u^{(2)}> 0$ a.e. if and only if the diagonal $u(P_2)u(P_3)$ is contained in $\mathcal Q$. The same holds for $\mathcal T^{(3)}$ using the corresponding diagonal.
On the other hand, an octahedron is convex if and only if all the three diagonals are contained in the inner part of the octahedron itself.
\end{proof}
The octahedron satisfies an estimate corresponding to the one of Lemma \ref{lemma:tetra}.
\begin{lemma}\label{lemma:ottaedro}
There exists $C>0$ such that
\begin{equation}\label{rig-ottaedro}
\mathrm{dist}^2(\nabla} %{\mathrm{D} u,SO(3)) \leq C \!\!\sum_{\NN{P_i}{P_j}}( |\nabla} %{\mathrm{D} u(P_iP_j)|-1)^2 \quad
\text{ a.e.\ in }\mathcal O\,,
\end{equation}
for every $u\in C^0(\mathcal O;{\mathbb R}^3)$ such that $u$ is piecewise affine
with respect to the triangulation determined by cutting $\mathcal O$ along the
diagonal $P_1P_4$,
$\det\nabla} %{\mathrm{D} u>0$ a.e.\ in $\mathcal O$,
and $u(\mathcal O)$ is convex.
\end{lemma}
\begin{proof}
Let $\chi_i$, $i=1,\dots,4$, be the characteristic functions of the
four tetrahedra $T_1:=P_1P_2P_4P_5$, $T_2:=P_1P_2P_4P_6$,
$T_3:=P_1P_3P_4P_5$, $T_4:=P_1P_3P_4P_6$,
respectively.
Since $\nabla u = \sum_{i=1}^4 \chi_i F_i$ for some $F_i\in GL^+(3)$, it suffices to prove \eqref{rig-ottaedro} in each tetrahedron.
Notice that
$P_1$ and $P_4$ are not nearest neighbours
and therefore we cannot directly apply Lemma~\ref{lemma:tetra}.
On the other hand, the length of $u(P_1P_4)$, which is a common edge of the four deformed
tetrahedra, can be expressed as a function of all the edges of $u(\mathcal O)$, the latter being
a (possibly irregular) octahedron. Specifically, from the rigidity of convex octahedra, it follows that there exists a function $f$ such that
$$
|u(P_1P_4)| = f(l_1,\dots,l_{12})\,,
$$
where $l_i$, $i=1,\dots ,12$, are the lengths of the twelve edges of $u(\mathcal O)$.
In particular we set
$$
l_1:=|u(P_1P_2)|\,, \,\,
l_2:=|u(P_2P_4)|\,, \,\,
l_3:=|u(P_2P_5)|\,, \,\,
l_4:=|u(P_1P_5)|\,, \,\,
l_5:=|u(P_4P_5)|\,\,.
$$
The explicit formula of $f$ is not important.
Let $\delta_i:= l_i-1$ for
$i=1,\dots ,12$, and $\delta_0:=|u(P_1P_4)|- \sqrt{2}$. We claim that $f$ is differentiable at
$(1,\dots,1)$. Then
$$
f(1+\delta_1,\dots,1+\delta_{12}) = \sqrt{2} + O(|\delta|), \quad\text{ with }\delta=(\delta_1,\dots,\delta_{12})\,,
$$
which yields, in combination with Lemma \ref{lemma:tetra}, the following inequality for $\nabla u = F_1$ on the tetrahedron $T_1$:
\begin{align*}
\mathrm{dist}^2(F_1,SO(3)) &\leq C \left(\sum_{i=1}^5 (l_i-1)^2 + \left(|u(P_1P_4)|- \sqrt{2}\right)^2\right) = C \sum_{i=0}^5 \delta_i^2 \\
&= C \sum_{i=1}^5 \delta_i^2 +
C\Big(f(1+\delta_1,\dots,1+\delta_{12}) - \sqrt{2}\Big)^2 \leq
C|\delta|^2
\end{align*}
for $|\delta|\leq 1$. On the other hand, by the triangle inequality, we have for $|\delta|> 1$
$$
\delta_0^2 \leq
2(f^2(1+\delta_1,\dots,1+\delta_{12}) + 2)\leq 4(l_1^2 + l_2^2 + 2)
\leq C'|\delta|^2\,.
$$
The inequality for the other $T_i$'s is completely analogous.
\par
We are left to show that $f$ is differentiable at $(1,\dots,1)$. To this end, we prove
the existence and continuity of all its partial derivatives at $(1,\dots,1)$.
By a symmetry argument, it is enough to study the existence and continuity of
$\partial_3 f$ and $\partial_4 f$ (with reference to Figure~\ref{fig:piramide}).
\par
\begin{figure}
\centering
\psfrag{q1}{$Q_1$}
\psfrag{q2}{$Q_2$}
\psfrag{q3}{\hspace{-1mm}$Q_3 $}
\psfrag{q4}{$Q_4$}
\psfrag{q5}{$Q_5$}
\psfrag{q6}{$Q_6$}
\subfloat[]{
\includegraphics[width=.4\textwidth]{piramide.eps}
\label{fig:piramide}
}
\hspace{.12\textwidth}
\psfrag{o}{$O$}
\psfrag{g}{$\gamma$}
\subfloat[]{
\includegraphics[width=.39\textwidth]{proiezione.eps}
\label{fig:proiezione}
}
\caption{(a) The image of $\O$ through a piece-wise affine map $u$ such that
$l_i=1$ for each $i\neq 3$.
(b) The projection of $u(\O)$ on the plane $p$, where $O=\Pi(Q_1)=\Pi(Q_4)$.}
\label{fig:entrambi}
\end{figure}
We begin with $\partial_3 f$.
Let $u(P_i)=Q_i$ be such that $l_i=1$ for each $i\neq 3$ and $l_3\neq 1$.
Since
$f^2(1,1,l_3,1,\dots ,1)= 2-2\cos\alpha$,
where $\alpha$ is the acute angle formed by $Q_1Q_2$ and $Q_2Q_4$,
$f(1,1,l_3,1,\dots ,1)$ is a smooth function of $\alpha$ for $0<\alpha<\pi$.
Next remark that the points $Q_2,Q_3,Q_5,Q_6$ are coplanar;
let $p$ denote the plane containing them and $\Pi$ be the orthogonal projection onto $p$
(see Figure~\ref{fig:proiezione}).
Considering the projections we see that
$$
|Q_2\Pi(Q_1)|=|Q_3\Pi(Q_1)|=|Q_5\Pi(Q_1)|=|Q_6\Pi(Q_1)|=\cos\tfrac{\alpha}{2}\,.
$$
Let $\gamma$ denote the acute angle formed by $Q_2\Pi(Q_1)$ and $Q_6\Pi(Q_1)$ (which
is equal to that formed by $Q_6\Pi(Q_1)$ and $Q_3\Pi(Q_1)$ and that formed by
$Q_3\Pi(Q_1)$ and $Q_5\Pi(Q_1)$).
Then, by the cosine formula
\begin{equation}\label{formula:gamma}
\cos\gamma = 1- \frac{1}{2\cos^2\frac{\alpha}{2}}
\end{equation}
and
\begin{equation}\label{quasiformula:l3}
|Q_2Q_5|^2= 2\cos^2\tfrac{\alpha}{2}\Big(1- \cos(2\pi - 3\gamma) \Big).
\end{equation}
Combining \eqref{formula:gamma} and \eqref{quasiformula:l3} we obtain
\begin{equation*}
l_3:= |Q_2Q_5|=
8\cos^2\tfrac{\alpha}{2} - 3 - \frac{\Big( 2\cos^2\frac{\alpha}{2} - 1\Big)^3}{\cos^4\frac{\alpha}{2}}
= 3 -\frac{1}{\cos^2\frac{\alpha}{2}} \,.
\end{equation*}
Remark that since $\alpha<\pi$, $l_3$ is a smooth function of $\alpha$.
Moreover, since the derivative $\displaystyle \frac{\d l_3}{\d\alpha}$ is not zero at the point $\alpha = \pi/2$, by the implicit function theorem it follows that $l_3$ is an invertible function of $\alpha$ in a
neighbourhood of $\alpha = \pi/2$, and its inverse is smooth in a neighbourhood of $l_3=1$.
The differentiability of $f$ with respect to $l_3$ at $(1,\dots,1)$ then follows from its smooth dependence on $\alpha$.
\par
Finally, proving the existence and continuity of $\partial_4 f$ at $(1,\dots,1)$ is equivalent to
proving that the length of $Q_3Q_2$ is a smooth function of $l_3$ in a neighbourhood of $(1,\dots,1)$.
The latter follows equivalently to the previous argument taking into account that
$\displaystyle|Q_3Q_2|=2\cos\tfrac{\alpha}{2}$.
\end{proof}
\begin{remark}
Estimates \eqref{rig-tetraedro} and \eqref{rig-ottaedro} are crucial in the proof of the compactness of sequences of deformations with equibounded energy, as well as in the study of the
$\Gamma$-limit and its scaling properties (see Theorem \ref{thm3} and
Proposition \ref{lowerbound}).
Indeed, as already remarked, each of the lattices introduced in Section \ref{3d} defines a tessellation
of the space into tetrahedra and octahedra.
This allows us to deduce the following lower bounds on the energy ${\mathcal E}_{\varepsilon}^{\lambda}$:
\begin{align}
\label{ineq1}
\mathrm{dist}^2(\nabla u_\varepsilon, SO(3)) & \leq C {\mathcal E}_{\varepsilon}^{\lambda}(u_\varepsilon) \text{ a.e. in }\Omega_{k\varepsilon}^- \\
\label{ineq2}
\mathrm{dist}^2\Big(\nabla u_\varepsilon, \tfrac{\lambda}{\rho}SO(3)\Big) & \leq C {\mathcal E}_{\varepsilon}^{\lambda}(u_\varepsilon)
\text{ a.e. in }\Omega_{k\varepsilon}^+\,,
\end{align}
for each admissible deformation $u_\varepsilon$.
Observe, in particular, that in the case of the diamond cubic lattice the above inequalities
are obtained by first neglecting in the energy the bonds between atoms of the sublattice $\L\dc{_2}$
and then applying \eqref{rig-tetraedro} and \eqref{rig-ottaedro} on the tessellation of the space thus
defined.
\par
Inequalities \eqref{ineq1}--\eqref{ineq2} imply,
via the rigidity estimate \eqref{rigidity}, that $\nabla u_\varepsilon$ is locally
close to $SO(3)$ in $\Omega_{k\varepsilon}^- $ and to $ \frac{\lambda}{\rho}SO(3)$ in $\Omega_{k\varepsilon}^+ $.
\end{remark}
\section{Dimension reduction and scaling properties of the $\Gamma$-limit} \label{sec:4}
In the present section we show the results in the three dimensional setting that
were obtained in two dimensions in our previous paper \cite{LaPaSc2014}.
The proofs of these results follow the lines of those in \cite{LaPaSc2014} by application of
Lemmas \ref{lemma:tetra} and \ref{lemma:ottaedro}. We will therefore omit further details of the proofs here.
\par
Given $R\in SO(3)$, $\rho\in(0,1]$, and $k\in{\mathbb N}$, we define the
minimum cost of a transition from an equilibrium in $\Omega_{k,\infty}$ with $\xi_1 < 0$ to an equilibrium in $\Omega_{k,\infty}$ with $\xi_1>0$ as
\begin{equation*
\begin{split}
\gamma^\lambda(\rho,k,R):=\inf\big\{ & {\mathcal E}_{\infty}^\lambda(v,\rho,k) \colon M>0\,,
\ v\in {\mathcal A}_{\rho,\infty}(\Omega_{k,\infty})\,, \\
& \nabla} %{\mathrm{D} v=I \ \text{for} \ x_1\in(-\infty,-M)\,,\, \ \nabla} %{\mathrm{D} v=\tfrac\lambda\rho R \ \text{for} \ x_1\in(M,+\infty)
\big\}\,,
\end{split}
\end{equation*}
where ${\mathcal E}_{\infty}^\lambda$ is defined in \eqref{energia-infinita}.
\begin{remark}
In fact it can be proved that $\gamma^\lambda(\rho,k,R)$ does not depend on the rotation $R$
(see \cite[Proposition 2.4]{LaPaSc2014}). We therefore set
$$
\gamma^\lambda(\rho,k):=\gamma^\lambda(\rho,k,I)\,.
$$
\end{remark}
The function $\gamma^\lambda(\rho,k)$, which depends on the number of planes of atoms of
the two lattices $\L_1^-$ and $\L_\rho^+$ contained in the domain $\overline\Omega_{k,\infty}$,
is in fact the relevant quantity that describes the system when $\varepsilon$ tends to zero. More precisely,
our goal is to show that for $k$ sufficiently large, there holds
$$
\inf_{\rho\in(0,1)} \gamma^\lambda(\rho,k)< \gamma^\lambda(1,k)\,,
$$
i.e., the system displays dislocations.
In order to prove this,
we perform a dimension reduction with respect to the directions $\mathrm{v}_2$, $\mathrm{v}_3$.
To this end, for each $u_\varepsilon\in {\mathcal A}_{\rho,\varepsilon}(\Omega_{k\varepsilon})$ we define the rescaled deformation
$$
\tilde u_\varepsilon(x):=u_\varepsilon (A_\varepsilon x) \,,
$$
where $A_\varepsilon$ is the matrix defined by
\begin{equation*}
\begin{cases}
A_\varepsilon \mathrm{v}_1 = \mathrm{v}_1\\
A_\varepsilon \mathrm{v}_2 = \varepsilon \mathrm{v}_2\\
A_\varepsilon \mathrm{v}_3 = \varepsilon \mathrm{v}_3 \,.
\end{cases}
\end{equation*}
This yields a scaling of the domain $\Omega_{k\varepsilon}$ to $\Omega_{k}$, which is independent of $\varepsilon$.
For fixed $\rho\in(0,1]$ and $k\in{\mathbb N}$ we address the question of the $\Gamma$-convergence of the sequence of functionals
$\{{\mathcal I}_\varepsilon\}$ defined by
$$
{\mathcal I}_{\varepsilon}(\tilde u_\varepsilon) := {\mathcal E}^\lambda_\varepsilon(u_\varepsilon,\rho,k) \quad \text{for}\ \tilde u_\varepsilon
\in \widetilde{\mathcal A}_{\rho,\varepsilon}(\Omega_{k}) \,,
$$
where $ \widetilde{\mathcal A}_{\rho,\varepsilon}(\Omega_{k})$ is the corresponding set of admissible deformations, i.e.
\begin{equation*}
\begin{split}
\widetilde{\mathcal A}_{\rho,\varepsilon}(\Omega_{k}):= \big\{ \tilde u_\varepsilon\in C^0(\overline\Omega_{k}; {\mathbb R}^3) \colon & \tilde u_\varepsilon \ \text{piecewise affine,}\\
& \nabla} %{\mathrm{D} \tilde u_\varepsilon \ \text{constant on}\ \Omega_{k}\cap (A_\varepsilon^{-1}T)\ \forall\, T\in\mathcal T_{\rho,\varepsilon}^{(1)}\,, \\
& \det \nabla} %{\mathrm{D}\tilde u_\varepsilon>0 \ \text{a.e.\ in}\ \Omega_{k}\,,\\
& \tilde u_\varepsilon(A_\varepsilon^{-1}\mathcal P) \ \text{is convex}\ \forall\, \mathcal P\in\mathcal T_{\rho,\varepsilon} \big\} \,.
\end{split}
\end{equation*}
\begin{theorem}\label{thm3}
Let $\{\tilde u_\varepsilon\}\subset \widetilde{\mathcal A}_{\rho,\varepsilon}(\Omega_{k})$ be a
sequence such that
\begin{equation*}
\limsup_{\varepsilon\to 0^{+}} {\mathcal I}_{\varepsilon}(\tilde u_\varepsilon) \leq C \,.
\end{equation*}
Then there exists a subsequence (not relabeled) such that
\begin{equation*}
\nabla} %{\mathrm{D} \tilde u_\varepsilon A_\varepsilon^{-1}\stackrel{\ast}{\rightharpoonup} (\partial_{\mathrm{v}_1} \tilde u \, | \, d_2 \, | \, d_3) \quad \text{weakly* in}\ L^{\infty}(\Omega_k;{\mathbb M}^{3\times 3})\,,
\end{equation*}
where the functions $\tilde u \in W^{1,\infty}(\Omega_k;{\mathbb R}^3)$, $d_2, d_3\in L^{\infty}(\Omega_k;{\mathbb R}^3)$
are independent of $\mathrm{v}_2$ and $\mathrm{v}_3$, i.e.,
$\partial_{\mathrm{v}_i}\tilde u = \partial_{\mathrm{v}_i} d_2 = \partial_{\mathrm{v}_i} d_3= 0$, for $i=2,3$.
Moreover,
\begin{equation*}
(\partial_{\mathrm{v}_1} \tilde u \, | \, d_2 \, | \, d_3)
\in
\begin{cases}
{\rm co}(SO(3)) & \text{a.e.\ in}\ \Omega_k^-\,,\\
{\rm co}(\frac\lambda\rho SO(3)) & \text{a.e.\ in}\ \Omega_k^+\,.
\end{cases}
\end{equation*}
The sequence of functionals $\{ {\mathcal I}_{\varepsilon}\}$
$\Gamma$-converges, as $\varepsilon\to 0^{+}$, to the functional
\begin{equation*
{\mathcal I}(u)=
\begin{cases}
\gamma^\lambda(\rho,k) & \text{if}\ u\in{\mathcal A}\,,\\
+\infty & \text{otherwise,}
\end{cases}
\end{equation*}
with respect to the weak* convergence in $W^{1,\infty}(\Omega_k;{\mathbb R}^3)$, where
\begin{equation*}
\begin{split}
{\mathcal A}:=\big\{
u\in W^{1,\infty}(\Omega_k;{\mathbb R}^3)\colon &
\partial_{\mathrm{v}_2} u = \partial_{\mathrm{v}_3} u =0 \ \text{a.e.\ in}\ \Omega_k\,, \\
& |\partial_{\mathrm{v}_1} u |\leq 1 \ \text{a.e.\ in}\ \Omega_k^- \,,
|\partial_{\mathrm{v}_1} u |\leq \tfrac\lambda\rho \ \text{a.e.\ in}\ \Omega_k^+
\big\} \,.
\end{split}
\end{equation*}
\end{theorem}
\begin{remark} \label{rmk5}
As in the two-dimensional case, the limit of a sequence of discrete deformations with equibounded energy
does not depend on the triangulation chosen for the octahedra, see \cite[Remark 3.4]{LaPaSc2014}.
Moreover, the $\Gamma$-limit does not depend on the choice of the triangulation $\mathcal T_{\rho,\varepsilon}^{(1)}$ in \eqref{ad-3d},
since its formula only depends on the discrete values of the deformation, and not on its extension to the three-dimensional continuum. Similarly, it does not depend on the choice of the tessellation of
$\L\dc{}$.
\end{remark}
The next two results characterise the behaviour of the $\Gamma$-limit in the
dislocation-free case and in the case when dislocations are present.
\begin{proposition}[Estimate in the defect-free case, $\rho=1$]\label{lowerbound}
There exist $C_1,C_2>0$ such that for every $k\in{\mathbb N}$
\begin{equation*}
C_1 k^3 \leq \gamma^\lambda(1,k) \leq C_2 k^3 \,.
\end{equation*}
\end{proposition}
\begin{proposition}[Estimate for $\rho=\lambda$] \label{prop:lambda}
There exist positive constants $C_1',C_2'$ such that for every $k$
\begin{equation*}
C_1' k^2\le \gamma^\lambda(\lambda,k)\le C_2' k^2 \,.
\end{equation*}
\end{proposition}
\begin{remark}
The proof of Proposition \ref{lowerbound} is a generalisation of its two-dimensional
counterpart (see \cite[Proposition 2.5]{LaPaSc2014}) given the rigidity estimates of the previous section.
In contrast,
the proof of Proposition \ref{prop:lambda} is straightforward: it follows
by testing ${\mathcal E}_{\infty}^\lambda$ on the identical deformation $v(x)=x$ and
taking into account that
each interfacial atom has a number of bonds that is uniformly bounded in $k$.
\end{remark}
\begin{remark}\label{finalrem}
Propositions \ref{lowerbound}--\ref{prop:lambda} in combination with Theorem \ref{thm3}
prove that dislocations
are energetically preferred if the thickness of the nanowire modeled by $k$ is sufficiently large.
\end{remark}
\section*{Acknowledgements} \noindent
This work was partially supported by the DFG grant SCHL 1706/2-1.
The research of G.L.\ was supported by the ERC grant No.\ 290888.
|
\section{Introduction}
High space velocities of OB runaway stars are explained by two independent mechanisms:
dynamical ejection due to gravitational interactions of massive stars in cluster cores
\citep{1967BOTT....4...86P}
and binary disruption as a result of a supernova explosion of the initially more massive component
\citep{1961BAN....15..265B}.
Both scenarios are viable, but whether one of the mechanisms is dominant is still uncertain.
According to the virial theorem, through a symmetric supernova explosion in a binary system, if more than half of the total mass of the system is released,
then the new born neutron star (or black hole) and the non--degenerate component are no more gravitationally bound
\citep{1961BAN....15..265B}.
However, the energy stored in the orbit, in most cases, is not sufficient to produce the neutron star kick velocities that are typically in the range of 300--500 \kms
\citep{%
1994Natur.369..127L,
1997ARep...41..257A,
2005MNRAS.360..974H,
1997MNRAS.291..569H}.
The asymmetry in supernova explosions is responsible for such high velocities
\citep{2013A&A...552A.126W}.
Therefore, there are no pulsar companions to many of the OB runaway stars
\citep{1996ApJ...461..357S}.
In some cases, the compact object does not receive a significant kick and/or the majority of the total mass is stored on the secondary through conservative mass transfer, hence, the compact object remains bound to the companion star
\citep{1993SSRv...66..309V}.
The runaway high mass X--ray binaries like 4U1700-37 \citep{2001A&A...370..170A} and Vela X--1 \citep{1997ApJ...475L..37K} are such examples.
Yet, the low rate of X--ray binaries and the high rate of isolated neutron stars
by taking into consideration the selection effects, the binary disruption is likely to occur in most cases
\citep{2005Ap.....48..330G} (hereafter, G05).
The kinematics of the binary disruption due to an asymmetric supernova explosion is widely discussed in
\citet{1998A&A...330.1047T}.
However, a sample of observationally confirmed OB runaway--NS couples is needed for a better understanding of the problem.
The importance of searching for OB runaways inside SNRs was first mentioned in
\citet{1980JApA....1...67V}.
However, the kinematical study of known OB stars inside SNRs was concluded with a lack of OB runaways due to the poor sample of SNRs and OB stars.
Still, there is no known O or B--type runaway star that can be directly linked to an SNR given in the literature.
In G05, the outcome of exploring runaway pairs from binary supernova disruption is broadly discussed.
Firstly, identifying the explosion centers more precisely will be useful for determining the velocities of young NSs
of which proper motion measurements have high uncertainties.
Thus, the kick that is gained by the NS due to the asymmetry of the SN can be determined more precisely.
Secondly, the distance to the remnant can be measured more accurately by studying the runaway star,
as it cannot move far away from the explosion center in the observational lifetime of the SNR.
Finally, possible effects of a close binary system on the asymmetry of the SNe can also be examined.
Additionally, it would be a direct evidence of the binary supernova scenario (BSS).
The observational efforts are concentrated on runaway--NS coupling and abundance investigations of runaway stars.
Some examples of the runaway--NS pairs (components are separated) are PSR\,B1929+10 -- $\zeta$ Oph
\citep{%
2001A&A...365...49H,
2008AstL...34..686B,
2010MNRAS.402.2369T},
PSR\,J0630-2834 -- HIP\,47155
\citep{2013MNRAS.435..879T}
and PSR\,J0826+2637 -- HIP13962
\citep{2014MNRAS.438.3587T}.
Based on their motions in space, the pulsars and the corresponding runaway stars were traced back in time by using 3--D Monte Carlo simulations.
They were found at the same position and the same time inside a young open cluster.
The PSR\,J0630-2834 -- HIP\,47155 pair is thought to be ejected from very old SNR Antlia.
There are considerable uncertainties in these cases as the supernova events took place more than \dex{5} yr ago.
The separation between the objects is very large and the SNR has faded away long ago and/or the components are outside of the SNR (the Antlia case).
Other important observational evidence is the enhancement of $\alpha$--process elements in the hyper--velocity star HD\,271791.
The star is proposed to be ejected from a massive close binary system due to an SN
that enriches its photosphere in elements which can be synthesized in large amounts during the evolution of the progenitor
\citep{2008ApJ...684L.103P}.
The method followed in our work is a direct study of possible runaway stars inside SNRs as described in G05.
Briefly, assuming that the massive binary mass ratio is greater than 1:4,
OB--type star candidates are determined by a careful study of their BVJHK magnitudes obtained from the UCAC4 catalog
\citep{2012yCat.1322....0Z}.
The angular separation of the sources from the geometrical centers of the remnants are within the limits of the maximum angular distance
that a runaway star can have in a lifetime of an SNR.
The distance moduli are calculated for all sources within this region,
and the sources having extinctions and radial distances consistent with those of the SNRs are considered as "candidates".
Measuring the radial velocities and identifying the spectral types of these objects via spectroscopy
reveal their runaway nature, their youth and the exact spatial relations with the SNRs, in other words, the genetic connections.
Although the runaway stars arising from BSS may also be late type stars,
it is time consuming to check the possible runaway nature of all the stars inside each SNR.
Also, the late type stars are too faint to observe at large distances.
A star having the same age as a supernova progenitor must be young.
As OB type stars evolve faster, they automatically satisfy this condition.
Furthermore, high mass stars are rare objects.
An OB runaway star discovered inside an SNR can be explained by the BSS.
Considering the very short observable lifetimes of SNRs and relatively short later evolution stages of stars,
most probably, main sequence stars are expected as OB runaways connected to SNRs.
The space velocity of an OB runaway star is thought to be larger than 30$-$40 \kms
\citep{%
1965MNRAS.130..245F,
1996ApJ...461..357S}.
A more precise value is proposed in
\citet{2011MNRAS.410..190T}; 28 \kms in 3--D and 20 \kms in 2--D.
To summarize, a main--sequence OB type star of
which at least one component of the velocity vector is greater than {20} \kms is searched in selected SNRs.
The first criterion of the candidate selection is the restriction of the angular position.
Most of the OB runaways have peculiar velocities lower than 80 \kms
\citep{%
1986ApJS...61..419G,
1996AJ....111.1220P,
2011MNRAS.410..190T}.
As the shock wave velocity of the SNRs are decelerated from roughly 10000 \kms to several hundred \kms,
it is expected that the runaway star cannot exceed one tenth of the angular diameter ($\theta$) of the related SNR.
This value is somewhat relaxed to $\theta$/6 considering the uncertainties in the geometrical centers of the SNRs.
The stars in this region are expected to be consistent with the respective SNR in terms of distance and reddening.
For this comparison, the adopted distances of SNRs given in
\citet{%
2003SerAJ.167...93G,
2004SerAJ.168...55G,
2004SerAJ.169...65G},
and \Av values from
\citet{1980BICDS..19...61N}
were used.
48 SNRs within 5 \kpc from the Sun were selected for investigation.
\begin{figure*}
\includegraphics[width=10cm,height=9cm]{F41-S147.pdf}
\includegraphics[width=6cm,height=9cm]{F41-S147-RunawayPSR-Zoom.pdf}
\caption{
The $4.8\,^{\circ}\times4.1\,^{\circ}$
\halpha image of SNR S147 taken at the University Observatory Jena.
The green cross represents the pulsar, the cyan cross shows the GC of the SNR and the yellow circle is HD\,37424.
The yellow vectors show the proper motion of the objects.
The red box is the zoom--in area shown in the right panel.
The white arrows are the tracing back cones of the proper motion for 29300 yr.
The angle between the pm vectors is $139^{\circ}-148^{\circ}$
In 2--D calculations, both objects come towards each other as close as 0.126 \pc which means that they have a common origin; binary supernova disruption.}
\label{f:mosaic}
\end{figure*}
In this paper, the result of the runaway search in SNR G180.0-1.7 (S147) is given.
The kinematic relation between the runaway star HD\,37424 and the pulsar PSR\,J0538+2817 is shown,
the possible host OB association is discussed,
the SNR parameters are constrained and
the pre--supernova binary is constructed.
\section{S147, PSR\,J0538+2817 and HD\,37424}
S147 is a shell type SNR located in the Galactic anti--center direction.
It is 180 arcmin in diameter with a geometrical center (GC) at
$\alpha=05\mathrm{h} 39\mathrm{m} 00\mathrm{s}$, $\delta=+27\,^{\circ} 50\mathrm{'} 00\mathrm{''}$
\citep{2009yCat.7253....0G}.
It was first mentioned as an SNR candidate in
\citet{1958RvMP...30.1048M}.
The compact object related to the SNR is radio pulsar PSR\,J0538+2817
\citep{1996ApJ...468L..55A}.
In optical bands, the shell structure is well defined and dominated by filamentary emission in \halpha.
The emission is brighter in the north and south edges and mainly concentrated in the southern parts.
Despite of its old age, it conserves the spherical symmetry except for the blowout regions in east and west
(Figure \ref{f:mosaic}).
The total absorption in the V band is \Av=$0.7\pm0.2$ magnitude
\citep{1985ApJ...292...29F}.
Radio observations reveal that the spectral index is unusually varying.
The shell structure observed at radio wavelengths coincides with that in optical bands and is well defined
\citep{1986A&A...163..185F}.
But, no X--ray emission is observed from the remnant
\citep{1990A&A...227..183S}.
Suggested distances vary from 0.6 to 1.9 \kpc in various publications (Table \ref{t:snr_distance}).
Four of the measurements are based on \sigmad relation which is useful to estimate the distance by using the surface brightness of the SNR.
In \citet{1980PASJ...32....1S} and \citet{1980A&A....92..225K}, the model of
\citet{1979AuJPh..32...83M}
is used, while in \citet{1976MNRAS.174..267C} and \citet{2003A&AT...22..273G}, estimations are based on their own models.
The radius lower limit calculated through SNR dynamics based on the model of
\citet{1974ApJ...188..501C}
sets another constraint on the SNR distance.
The distance derived from the pulsar's parallax or dispersion measure is larger than the distance suggested for the background stars of which spectra show high velocity gas related to the SNR.
\begin{table}
\caption{Distance estimates for S147. (R$^{l}$) denotes the radius lower limit of the SNR.
}
\label{t:snr_distance}
\begin{center}
\begin{tabular}{l l l}
Distance (\kpc)
& Method
& Reference \\\hline
1.6$\pm$0.3 & \sigmad & {\citet{1980PASJ...32....1S}} \\
0.8--1.37 & R$^{l}$, \sigmad & {\citet{1980A&A....92..225K}} \\
0.6 & R$^{l}$ & {\citet{1979ApJ...229..147K}} \\
0.9 & \sigmad & {\citet{1976MNRAS.174..267C}} \\
0.8$\pm$0.1 & \Av & {\citet{1985ApJ...292...29F}} \\
1.06 & \sigmad & {\citet{2003A&AT...22..273G}} \\
0.88< & High Vel Gas & {\citet{2004A&A...426..555S}} \\
1.2 & Pulsar DM & {\citet{2003ApJ...593L..31K}} \\
1.47$^{+0.42}_{-0.27}$ & Pulsar Plx & {\citet{2007ApJ...654..487N}} \\
1.3$^{+0.22}_{-0.16}$ & Pulsar Plx & {\citet{2009ApJ...698..250C}} \\\hline
\end{tabular}
\end{center}
\end{table}
The age was estimated as more than 100 \kyr
from the Sedov solution by using a shock velocity $\sim$80--100 \kms
\citep{1980PASJ...32....1S,
1980A&A....92..225K}.
This velocity was derived from the RV of the filamentary knots in
\citet{1976SvA....20...19L,
1979ApJ...229..147K}.
However the pulsar--SNR relation indicates an age of $30\pm4$ \kyr (from the travel time from the GC to the present position)
\citep{2003ApJ...593L..31K,
2007ApJ...654..487N}.
For a distance of 1.3 \kpc, and an angular diameter of 200 arcmin, the radius (R) of the SNR is 38 \pc.
Then, assuming an age of 30 \kyr, the remnant is in its Sedov Phase having a blast wave velocity of 500 \kms.
The estimated explosion energy is between 1$-$3\tdex{51} erg and the corresponding inter--cloud gas density is 0.03$-$0.1 $\mathrm{cm^{-3}}$
\citep{2012ApJ...752..135K}.
The SNR is expanding in a low density medium, probably in the cavity generated by the progenitor.
The central source, PSR\,J0538+2817, is an extensively studied radio and X--ray pulsar located at
$\alpha=05\mathrm{h} 38\mathrm{m} 25.1\mathrm{s}$,
$\delta=+28\,^{\circ} 17\mathrm{'} 09.2\mathrm{''}$,
$\sim$28 arcmin away from the GC towards north.
The 143.16 $\mathrm{ms}$ period and 3.67\tdex{-15} $\mathrm{s\,s^{-1}}$ period derivative
\citep{1996ApJ...468L..55A}
imply a characteristic age of $\sim$620 \kyr which is $\sim$20 times larger than its kinematic age of 30 \kyr.
This discrepancy is explained by either a long initial spin period of P$_{0}=139$ \msec
\citep{2003ApJ...593L..31K}
or strong magnetic field decay
\citep{2004IJMPD..13.1805G}.
The parallax distance was measured as $1.47\pmpm{+0.42}{-0.27}$ \kpc
\citep{2007ApJ...654..487N}
and $1.3\pmpm{+0.22}{-0.16}$ \kpc
\citep{2009ApJ...698..250C}
and the dispersion measure (DM) distance is 1.2 \kpc
\citep{2003ApJ...593L..31K}
by using the NE2001 model of
\citet{2002astro.ph..7156C}
which are consistent with each other.
The most precisely measured proper motion is
$\mu_{\alpha}^{*}=-23.57\pmpm{+0.10}{-0.10}$ \masyr, $\mu_{\delta}=52.87\pmpm{+0.09}{-0.10}$ \masyr
corresponding to a transverse velocity of $357\pmpm{+59}{-43}$ \kms at $1.3\pmpm{+0.22}{-0.16}$ \kpc
\citep{2009ApJ...698..250C}.
Although it presents no clear $\gamma$--ray emission
\citep{2012ApJ...752..135K},
the pulsar is observable in soft X--rays due to thermal emission.
PSR\,J0538+2817 is famous for being one of the few thermal pulsars:
Based on a blackbody model, the surface temperature and the emitting radius are
given as T$=2.12\tdex{6}$ K and R$=1.68\pm0.05$ \km
\citep{2003ApJ...591..380M}.
The hydrogen atmosphere model with B$=1\tdex{12}$ G gives T$=2.12\tdex{6}$ K and R$\simeq$10 \km
\citep{2004MmSAI..75..458Z}.
A faint pulsar wind nebula is observed in X--rays but not in radio
\citep{2000MNRAS.318...58G}.
It is not known whether the observed elongated structure is the torus or the jets of the PWN.
But assuming that it is due to the torus, it provides valuable information on the spin--kick alignment.
HD\,37424 is a sound OB runaway star at 10.3 arcmin away from the GC to the west at
$\alpha=05\mathrm{h} 39\mathrm{m} 44.4\mathrm{s}$, $\delta=+27\,^{\circ} 46\mathrm{'} 51.2\mathrm{''}$.
Photometric and kinematic information on the star is given in Table \ref{t:star}.
While proper motion values were retrieved from the $\textit{UCAC4}$ catalog
\citep{2012yCat.1322....0Z},
the photometric magnitudes are obtained from the $\textit{ASCC-2.5 V3}$ catalog
\citep{2001KFNT...17..409K}.
Its spectral type was previously identified as B0.5/1 IV/V
\citep{1979RA......9..479C}.
As there was no public data nor a visualized spectrum, we established the spectral type again by taking a spectrum.
The star has a very high proper motion compared to other early type and possibly distant stars.
There is no reported variability or binarity in the literature.
Hence, HD\,37424 is a good OB runaway star candidate.
\begin{table}
\caption{BVJHK (in mag) and pm (in \masyr) values of HD\,37424.}
\label{t:star}
\begin{center}
\begin{tabular}{c r@{$\pm$}l}
Parameter
& \multicolumn{2}{c}{Value} \\\hline
B & 9.062&0.017 \\
V & 8.989&0.019 \\
J & 8.666&0.017 \\
H & 8.696&0.026 \\
K & 8.699&0.019 \\
$\mu_{\alpha}^{*}$ & 10.8&0.8 \\
$\mu_{\delta}$ & -10.2&0.6 \\\hline
\end{tabular}
\end{center}
\end{table}
\section{Observations}
HD\,37424 was observed at the \tubitak National Observatory (TUG)
with the TUG Faint Object Spectrograph and Camera (TFOSC) mounted on the 150\,cm Russian Turkish Telescope (RTT--150).
The observations were carried out on 2013 August 9 and 10.
Two spectra were obtained with a 600\,s exposure time for each.
Grism 14 was used with a 67 micron slit.
The wavelength range in this configuration is 3270--6120 \AA\ and the resolving power (R) is $\sim$1337.
5 He lamp spectra taken in the same night with the object were used for wavelength calibration,
and 5 halogen lamp spectra were obtained in the beginning and at the end of the night for flat--fielding.
Also, 2 more targets had been observed in the same way on 2012 September 13 and 14 for the search for the runaway star.
With the same purpose, 12 objects were observed at the Calar Alto Observatory on 2013 January 29
with Calar Alto Faint Object Spectrograph (CAFOS) on the 2.2 meter telescope.
The covered wavelength range of the grism B--200 is 3200--7000 \AA\ and R$\sim$550.
3 HgCdAr and 3 halogen spectra were taken in the beginning and in the end of the night for wavelength calibration and flat--fielding.
The high resolution observations of HD\,37424 were performed at the Fred L. Whipple Observatory on 2013 September 16 and 21.
The Tillinghast Reflector Echelle Spectrograph (TRES) mounted on the 1.5 meter telescope was used.
Two spectra were taken at R$\sim$44000 in a broad wavelength range; 3800--9100 \AA\ for 300\,s exposure time.
HD\,36665 which is supposed to be a background object to the SNR was observed on 2014 February 11 with the
Fibre Linked Echelle Astronomical Spectrograph (FLECHAS)
\citep{2014AN....335..417M}
at the University Observatory Jena.
The covered wavelength range is 3900--8100 \AA\ and R$\sim$9200.
Three spectra with 600 s exposure time were obtained.
Three tungsten and 3 Th-Ar lamp spectra were taken immediately before this set for flat--fielding and wavelength calibration.
The observations of SNR\,S147 were performed at the University Observatory Jena with the Schmidt-Teleskop-Kamera (STK),
operated in the Schmidt-focus of the 90\,cm telescope of the observatory (see \citet{2010AN....331..449M}).
The observations were carried out with the narrow--band \halpha filter of the STK, which spans a full--width--half--maximum of 5 nm.
In total, 22 fields on the sky were observed during 5 nights, each with a field of view of $52.8'\times52.8'$.
At each position two 600\,s integrations were taken with the STK, i.e. about 7.3 hours of integration time in total.
The individual raw images per field were dark- and flatfield--corrected and finally averaged.
Point sources detected in the overlap regions of the fully processed images were then used to create the mosaic image of SNR\,S147, shown in Figure \ref{f:mosaic},
which covers a total field of view of $4.8^{\circ}\times 4.1^{\circ}$.
FLECHAS, TFOSC and CAFOS data were reduced with IRAF.
All raw frames from TFOSC and CAFOS were over--scan stripped
and no further subtraction was done as long as the over--scanned dark frames yields null count.
Yet, FLECHAS images were subtracted by the dark frames taken in the previous night.
The exposure time of a dark frame is the same as that of the corresponding science or calibration frame.
Hence, the illumination effect was removed.
Then, the flat images were combined while rejecting the frames having minimum and maximum counts.
Next, the object spectra were extracted by assigning an appropriate background region to each.
The arc spectra were extracted as traced by the corresponding aperture of the object spectra.
After wavelength calibration was performed, all spectra were normalized.
The final spectra were compared with MK standard star spectra retrieved from
\citet{1990PASP..102..379W}.
To maintain equally resolved spectra, the standard spectra were boxcar smoothed by 13.
Figure \ref{f:spt} shows the TFOSC spectrum of HD\,37424 comparing with main sequence B type stars from various subclasses.
HD\,37424 shows
\ion{He}{ii} $\lambda$4686
absorption indicating that the spectral type is earlier than B1.
This feature is not seen in stars with a spectral type later than B0.7.
On the other hand,
\ion{He}{i} lines (\ie \ion{He}{i} $\lambda$4713)
are not as weak as in O type stars compared to
\ion{He}{ii} lines; $\lambda$4200, $\lambda$4686.
The
\ion{Si}{iv} $\lambda$4116
line is totally blended and unmeasurable.
\ion{Si}{iii} $\lambda$4552
is in low strength compared to
\ion{Si}{iv} $\lambda$4089
like in the B0V type spectrum.
This ratio decreases towards higher temperatures, but
\ion{He}{i} $\lambda$4541
is not stronger than
\ion{Si}{iii} $\lambda$4552
as seen in B0V and O9.5V spectra.
Given the equal strengths of
\ion{O}{ii} $\lambda$4640,
\ion{He}{ii} $\lambda$4686,
\ion{He}{i} $\lambda$4713
and more intense
\ion{C}{iii}+\ion{O}{ii} $\lambda$4650 blends,
HD\,37424 matches best with B0.2V.
Considering the slightly stronger
\ion{He}{i} $\lambda$4713,
B0.5 would be the best approach to the temperature class of the star.
\ion{Si}{iv} $\lambda$4089
strength increases against
\ion{He}{i} $\lambda$4026--4144
as the luminosity class goes higher.
However, neutral helium lines are strong enough to judge that HD\,37424 is a main sequence star;
B0.5V.
The spectral type is between B0.2V--B0.7V, but it is assumed as B0.5$\pm$0.5V in the distance and extinction calculations to avoid underestimation of the errors.
\begin{figure}
\includegraphics[width=8cm]{F31-spt}
\caption{TFOSC spectrum of HD\,37424 compared to main sequence B type stars.
Distinctive features are marked. From left to right;
\ion{Si}{iv} $\lambda$4089,
\ion{He}{ii} $\lambda$4200,
\ion{Si}{iii} $\lambda$4552,
\ion{O}{ii} $\lambda$4640,
\ion{C}{iii}+\ion{O}{ii} $\lambda$4650 blend,
\ion{He}{ii} $\lambda$4686,
\ion{He}{i} $\lambda$4713}
\label{f:spt}
\end{figure}
The sources observed by CAFOS are foreground stars with a maximum distance of $671\pmpm{+96}{-56}$ \pc.
Their spectral properties were identified by the same procedure as done for HD\,37424.
However, due to the lower resolution CAFOS spectra have higher uncertainty (Table \ref{t:obs_result}).
We used FLECHAS to identify the spectral types of the background star HD\,36665.
This is a B1$\pm$0.5V type star (Figure \ref{f:hd36665}).
\begin{figure}
\includegraphics[width=8cm]{F32-hd36665}
\caption{
The FLECHAS spectrum of HD\,36665 (black) in comparison with a B1V template.
}
\label{f:hd36665}
\end{figure}
\begin{table}
\caption{Observed stars in the region.
Names, angular separations to the SNR GC (in arcmin), spectral types, instruments used and
distances with upper and lower errors (in \pc) are given.
The error in temperature subclasses is $\pm$1 for those with integer value and $\pm$0.5 for those with half integer value.
The resolution of CAFOS is not enough to distinguish the luminosity classes IV and V.
Yet, the stars are assumed to be in luminosity class V.}
\label{t:obs_result}
\begin{center}
\begin{tabular}{@{}c@{~~}c@{~~}c@{~~}c@{~~}c@{~}r@{~/~}l@{}}
Name
& Angular
& Spec.
& Inst.$^1$
& Distance
& \multicolumn{2}{c}{Distance} \\
& Separation
& Type
&
&
& \multicolumn{2}{c}{(Error)} \\\hline
TYC 1869-01281-1 & 19.4 & A2V & C & 574 & +73& -41 \\
TYC 1869-01317-1 & 5.5 & B9.5V & T &1013 &+183&-174 \\
TYC 1869-01334-1 & 25.5 & F2V & C & 289 & +38& -34 \\
TYC 1869-01376-1 & 27.9 & A7V & C & 428 & +18& -32 \\
TYC 1869-01505-1 & 23.5 & F1V & C & 380 & +50& -44 \\
TYC 1869-01610-1 & 17.6 & F7V & C & 245 & +18& -16 \\
TYC 1869-01632-1 & 27.2 & F8V & C & 316 & +37& -26 \\
TYC 1869-01642-1 & 5.61 & B9.5V & T & 952 &+172&-163 \\
TYC 1869-01679-1 & 25.6 & G2V & C & 170 & +11& -9 \\
TYC 1869-01749-1 & 20.1 & F8V & C & 166 & +19& -14 \\
TYC 1873-00145-1 & 25.0 & A3V & C & 603 & +46& -50 \\
TYC 1873-00307-1 & 29.2 & F8V & C & 326 & +40& -29 \\
TYC 1873-00347-1 & 19.8 & A0V & C & 671 & +96& -56 \\
UCAC4-589-020390 & 22.1 & F4V & C & 381 & +14& -16 \\\hline
\multicolumn{7}{l}{$^1$ Instruments: C: CAFOS, T: TFOSC} \\
\end{tabular}
\end{center}
\end{table}
TRES data were used for radial velocity (RV) measurements of HD\,37424.
Due to the low signal to noise, 10 absorption features could be used (Table \ref{t:rv}).
Doublets or triplet lines are avoided as these features increased the dispersion significantly.
The lines were fitted by Gaussian functions and the best shifts were determined.
Two more Gaussian fits with different centers were applied for each line;
one fits the outer edge of the noise of the blue side while fitting the inner edge of the red.
The other fits the outer edge of the red and the inner edge of the blue.
By doing this, underestimating the errors is avoided.
As the features are wide and the data are noisy,
the centers of these secondary fits are far away from the best fits, 16 \kms on average.
The difference between the best fit and the upper limit (towards red) fit is distinct from
the best fit and the lower limit (towards blue) difference.
To avoid underestimating the error, the greater difference is assigned as the final error.
The average heliocentric velocities of the first and the second spectra are -9.2 and -9.1 \kms
with line to line scattering standard deviations being 6.5 and 5.5 \kms respectively.
\begin{table}
\caption{Radial velocity measurements of HD\,37424.
The laboratory wavelength of the features, the Gaussian width ($\sigma$) of the fit,
the wavelength shift obtained by the best fitting Gaussian,
upper and lower limits of the shift from different Gaussian fits and the final error are presented.
The features are given in \AA\ while all other columns are in \kms}
\label{t:rv}
\begin{center}
\begin{tabular}{@{}c c c c c c@{}}
& & \multicolumn{4}{c}{Observations on September 16th, 2013} \\\cline{3-6}
\noalign{\smallskip}
Feature & $\sigma$ & BF & UL & LL & E \\\hline
\noalign{\smallskip}
3889.051 & 164 & -10.1 & 19.3 & -29.2 & 29.4 \\
3970.074 & 137 & -0.5 & 21.4 & -22.3 & 21.9 \\
4101.737 & 144 & -9.4 & 12.2 & -32.3 & 22.9 \\
4143.759 & 98 & +4.0 & 17.1 & -10.9 & 14.9 \\
4340.468 & 142 & -10.6 & 6.2 & -28.8 & 18.2 \\
4387.928 & 81 & -13.1 & -7.1 & -24.1 & 11 \\
4861.332 & 139 & -13.8 & 0.9 & -18.9 & 14.7 \\
4921.929 & 89 & -16.5 & -1.6 & -32.4 & 15.9 \\
6562.817 & 115 & -14.7 & 0.3 & -21.5 & 15 \\
6678.149 & 82 & -7.5 & 0.3 & -14.5 & 7.8 \\\hline
\noalign{\smallskip}
& & \multicolumn{4}{c}{Observations on September 21st, 2013} \\\cline{3-6}
\noalign{\smallskip}
Feature & $\sigma$ & BF & UL & LL & E \\\hline
\noalign{\smallskip}
3889.051 & 164 & -12.6 & 6.2 & -28.2 & 18.8 \\
3970.074 & 137 & -1.4 & 18.0 & -19.8 & 19.4 \\
4101.737 & 144 & -5.1 & 14.9 & -21.2 & 20 \\
4143.759 & 98 & -1.9 & 7.2 & -8.7 & 9.1 \\
4340.468 & 142 & -11.5 & 11.9 & -36.6 & 25.1 \\
4387.928 & 81 & -8.8 & -4.1 & -19.3 & 10.5 \\
4861.332 & 139 & -11.7 & -3.0 & -19.5 & 8.7 \\
4921.929 & 89 & -17.2 & -4.9 & -31.6 & 14.4 \\
6562.817 & 115 & -14.4 & 1.4 & -29.7 & 15.8 \\
6678.149 & 82 & -6.7 & 3.3 & -15.1 & 10 \\\hline
\noalign{\smallskip}
\multicolumn{6}{l}{BF: Best fit, UL: Upper Limit, LL: Lower Limit, E: Error}\\
\end{tabular}
\end{center}
\end{table}
As long as the star is inside the supernova remnant,
high velocity gas accelerated by the SNR is expected to reveal itself as
blue--shifted components to \ion{Ca}{ii}--K and H and/or \ion{Na}{i}--D1 and D2 absorption features.
The spectra show no clear high velocity features (Figure \ref{f:3933}).
The average heliocentric velocity of the interstellar gas is 12.1$\pm$0.5 \kms (Table \ref{t:ca_vel}).
This is typical for the neighboring stars
\citep{1973ApJ...181..799S}.
The star has no positional correspondence with bright filaments, but with fainter \halpha emitting regions.
It is located at the vicinity of the eastern cavity seen in the radio and $\gamma$--ray images
\citep{1980A&A....92..225K,
2012ApJ...752..135K}.
As the star can hardly be a foreground source,
we suggest that, in this direction, the shocked ejecta has not reached the dense interstellar gas yet.
The reader must note that, the background stars displaying high velocity \ion{Ca}{ii} lines in
\citet{2004A&A...426..555S}
are located behind bright filament knots.
The high velocity gas is found where the filaments are concentrated
\citep{1976SvA....20...19L}.
\begin{figure*}
\includegraphics[width=5.5cm]{F33-3933}
\includegraphics[width=5.5cm]{F33-3968}
\includegraphics[width=5.5cm]{F33-5890}
\caption{The strong interstellar lines of HD37424.
\textit{Left}: \ion{Ca}{ii}--K line from TRES spectrum.
Despite a weak blended feature at -13 \kms, there is no clear high velocity component.
\textit{Middle}: \ion{Ca}{ii}--H line (H$\epsilon$ as a background) from the same spectrum.
Again there is no clear high velocity gas component.
\textit{Right}: Interstellar \ion{Na}{i}--D1 and D2 lines.
The feature with high FWHM is the \ion{He}{i} $\lambda$5875 triplet. Those around Sodium doublet are tellurics.
}
\label{f:3933}
\end{figure*}
\begin{table}
\caption{Measured velocities for interstellar \ion{Ca}{ii}--K and H and \ion{Na}{i}--D1 and D2 lines.
These lines have a low Gaussian width unlike the intrinsic features of the star.
The average velocity is 12.1 \kms with a standard deviation 0.5 \kms in each observation.
The data in the first column are in (\AA) while the others are in \kms.}
\label{t:ca_vel}
\begin{center}
\begin{tabular}{c c c c}
Feature & $\sigma$ & Velocity & Velocity \\
& & (Day 1) & (Day 2) \\\hline
\noalign{\smallskip}
3933.664 & 7.5 & 12.3 & 11.8 \\
3968.47 & 5.9 & 12.8 & 12.7 \\
5889.953 & 9.0 & 11.6 & 11.6 \\
5895.923 & 8.3 & 11.9 & 12.0 \\\hline
\noalign{\smallskip}
\multicolumn{4}{l}{Day 1: Sep. 16th, 2013; Day 2: Sep 21st, 2013.}\\
\end{tabular}
\end{center}
\end{table}
Using the absolute visual magnitude from \citet{1982lbg6.conf.....A},
for a spectral type B0.5V$\pm$0.5, the distance modulus yields $1868\pmpm{+410}{-403}$ \pc for HD\,37424.
Using the luminosity of the same spectral type from
\citet{2010AN....331..349H}, we find the distance as 1318$\pm$119 \pc which is well consistent with the distance of the pulsar.
The total visual absorption was taken into account in both calculations.
As the absolute magnitudes for early type stars can range from star to star in the same temperature,
interstellar \ion{Ca}{ii} lines are also used in distance determination.
By using the following relation from \citet{2009A&A...507..833M},
\begin{equation}
D = 77+\left(2.78+\frac{2.60}{\frac{EW(K)}{EW(H)}-0.932}\right)EW(H)
\end{equation}
the distance to HD\,37424 is found to be $1288\pmpm{+304}{-193}$ \pc which is almost in the same range as the pulsar.
The equivalent widths are 243$\pm$7 and 160$\pm$15 m\AA\ for \ion{Ca}{ii}--K and \ion{Ca}{ii}--H lines respectively.
The extinction towards the star was also derived from 4430 and 4502 \AA\ DIB's.
The measured equivalent widths have high error due to the blending.
Using the EW's, the E(B-V) was calculated based on the relation mentioned in
\citet{1975ApJ...196..129H}.
It yields 0.42$\pm$0.08 mag; points to somewhat larger color excess but does not exclude 0.35$\pm$0.04 mag derived from photometry.
\section{Kinematics}
HD\,37424 and PSR\,J0538+2817 have proper motions receding from each other and departing from the same location on the sky (Figure \ref{f:mosaic}).
The proper motion of both objects are corrected for Galactic rotation and solar motion.
The Galactocentric distance to the Sun is taken as 8.5 \kpc and the solar rotational velocity as 220 \kms.
The local standard of rest was taken from \citet{2011MNRAS.410..190T} as
(U$_{\odot}$,V$_{\odot}$,W$_{\odot}$) =
(10.4$\pm$0.4, 11.6$\pm$0.2, 6.1$\pm$0.2) \kms.
The resultant values for HD\,37424 are
$\mu_{\alpha}^*=10.0\pm0.8$ \masyr,
$\mu_{\delta}=-5.9\pm0.6$ \masyr and
for the pulsar,
$\mu_{\alpha}^*=-24.4\pm0.1$ \masyr,
$\mu_{\delta}=57.2\pm0.1$ \masyr.
Together with the $-20.0\pm6.5$ \kms peculiar radial velocity, HD\,37424 has a space velocity of 74$\pm$8 \kms at 1.3 \kpc.
HD\,37424 is a runaway star of which velocity is higher than typical runaway velocities 40--50 \kms.
The 2--D space velocity of the pulsar at the same distance is 382.2$\pm$0.8 \kms.
We constructed the past 3D trajectories of PSR\,J0538+2817 and the runaway star HD\,37424
to evaluate whether these two objects could have been at the same place at the same time in the past.
Since we applied the same method already in preceding papers \citep{2010MNRAS.402.2369T,2011MNRAS.417..617T,2012PASA...29...98T},
we refer to these publications for details.
We construct three million past trajectories of PSR\,J0538+2817 and HD\,37424
throughout Monte Carlo simulations by varying the observables (parallax, proper motion, radial velocity) within their error intervals.
For the radial velocity of the NS, we assume a uniform distribution in the range of -1500 to +1500 \kms.
From all pairs of trajectories, we evaluate the smallest separation $d_{min}$ and the past time $\tau$ at which it occurred.
The distribution of separations $d_{min}$ is supposed to obey the distribution of absolute differences of
two 3D Gaussians (see e.\,g. \citealt{2001A&A...365...49H}, equations A3 and A4; \citealt{2012PASA...29...98T}, equations 1 and 2),
if it is assumed that the stellar 3D positions are Gaussian distributed.
Since the actual (observed) case is different from this simple model
(no 3D Gaussian distributed positions, due to e.\,g. the Gaussian distributed parallax that goes into the position reciprocally,
uniform radial velocity distribution of the NS, etc.), we adapt the theoretical formulae (we use equations 1 and 2 in
\cite{2012PASA...29...98T}
with the symbols $\mu$ and $\sigma$ for the expectation value and standard deviation, respectively)
only to the first part of the $d_{min}$ distribution (up to the peak plus a few more bins, see \citealt{2012PASA...29...98T}).
The derived parameter $\mu$ then gives the positional difference between the two objects.
\footnote{The uncertainties on the separation are dominated by the kinematic uncertainties of the NS
that are typically of the order of a few hundred \kms (because of the assumed radial velocity distribution).
As a consequence, the distribution of separations $d_{min}$ shows a large tail for larger separations.
However, the first part of the $d_{min}$ distribution (slope and peak) can still be explained well with the theoretical curve
(here equation 2 in \citealt{2012PASA...29...98T}) since the kinematic dispersion for only those runs are much smaller, a few tens of \kms for the NS, i.e. a few tens of pc after 1 Myr.}
We find that both stars were at the same position in the past
\ie
$\mu=0$ at
$\left(l, b\right)=\left(84.82\pm0.01, 27.84\pm0.01\right)$ deg at
$30\pm4$ \kpc (Figure \ref{f:nina}).
This predicted position of the supernova is $4.2\pmpm{+0.8}{-0.6}$ arcmin offset from the nominal geometric center.
The predicted distance of the supernova (as it is seen from the Earth today) is $1333\pmpm{+103}{-112}$ \pc.
\begin{figure}
\includegraphics[width= 8 cm]{F42-HistogramNina.pdf}
\caption{Distributions of minimum separations $d_{min}$ and corresponding flight times $\tau$ for encounters between PSR\,J0538+2817
and HD\,37424. The solid curves drawn in the $d_{min}$ histogram (bottom panel) represent the theoretically expected distribution with $\mu=0$ and $\sigma=0.35$ \pc, adapted to the first part of the histogram.}
\label{f:nina}
\end{figure}
\section{Associations}
There is no known OB association within 4.5$^{\circ}$ ($\sim$100 \pc at 1300 \pc) from SNR S147.
Considering the angular separations and radial distances, the progenitor star cannot be linked to any of the young open cluster around.
Hence, it is considered to be a runaway star ejected \eg from the cluster NGC 1960, which is 217 \pc away from the SNR, several millions years ago
\citep{2007ApJ...654..487N}.
However, binarity among cluster ejected runaway stars are rare
\citep{1986ApJS...61..419G}.
So, as the progenitor was the previous binary companion to HD\,37424, it may be unlikely that they were ejected from a cluster.
Therefore, the neighboring stars are investigated to search for any association.
All of the OB type stars within 4.5$^{\circ}$ (100 pc at 1.3 kpc distance) of the geometrical center (GC) of the SNR
were selected from \textit{The Catalogue of Stellar Spectral Classifications}
\citep{2013yCat....1.2023S}.
The spectral types are known through spectroscopic observations.
There are some discrepancies in spectral types reported by different papers.
So, always the latest reference was chosen.
Together with 14 stars from our observations at TUG and Calar Alto, 99 stars are used.
The photometric data were obtained from the catalog \textit{ASCC-2.5 V3}
\citep{2001KFNT...17..409K} except for UCAC4--589--020390 of which photometric
data was retrieved from the \textit{UCAC4} catalog.
For each star having an integer spectral subclass, distance and extinction are calculated in an interval
between one spectral subclass above and below, \eg B0V--B2V for a B1V type star.
For those having half integer subclass, half spectral subclass above and below are used \eg B0V--B1V for a B0.5V type star.
The total visual extinction, \Av, was determined by using BVJHK colors and intrinsic color differences of corresponding spectral types.
Using the following relation, \Av values were derived for each color difference; B-J, B-H, B-K, V-J, V-H, V-K.
\begin{equation}
\Av =
\frac{(\lambda_{1}-\lambda_{2}) - (\lambda_1-\lambda_2)_0}%
{\mathrm{A_{\lambda_1}}/\Av-\mathrm{A_{\lambda_2}}/\Av},
\end{equation}
The intrinsic colors are obtained from
\citet{1995ApJS..101..117K} and
\citet{1994MNRAS.270..229W},
while $\mathrm{A_{\lambda}}$/$\mathrm{A_{V}}$ ratios are from
\citet{1985ApJ...288..618R}.
Color differences of short intervals, \ie H-K have high errors as they are multiplied by higher coefficients.
The color excess E(B-V) is mentioned separately.
(Table \ref{t:26_stars})
\Av values obtained from six color differences were averaged and
their standard deviation were calculated.
This was applied for all three possible spectral types of the source.
As long as the error due to the spectral type uncertainty is larger than the standard deviation of the color differences,
it was assigned to be the final error.
When the error is asymmetric due to the intrinsic colors of different spectral types, the larger one was accepted.
In some cases, the uncertainty is dominated by the error in color differences.
Then, these were preferred to be the final error for such sources.
E(B-V) versus \Av fit
yields a total to selective absorption ratio is 3.24$\pm$0.06.
In individual cases, E(B-V) deviates strongly from \Av.
However, the large sample reveals that the ratio of total to selective absorption has a usual value.
The distances were derived from distance moduli using the absolute magnitudes from
\citet{1982lbg6.conf.....A}.
Errors in distances are due to the uncertainty in spectral type and the error in $\mathrm{A_{V}}$ were also taken into account.
The spectro--photometric distance alone is not enough
due to the high dispersion in the brightness of OB--type stars
\citep{2006MNRAS.371..185W} (Thereafter W06).
Assuming the stars beyond 1 \kpc are within 100 \pc from the SNR GC, the absolute visual magnitudes were calculated.
Although the dispersion mentioned in W06 is high, 24 of the stars fit well in the comparison with the W06 values.
Hence, we suggest that these stars are members of an OB association.
(Table \ref{t:26_stars}, Table \ref{t:M_stars},
19 of them have similar pm values.
The average pm in right ascension is -1.39 \masyr and -4.17 \masyr
in declination with 0.99 \masyr and 1.45 \masyr standard deviation respectively (Table \ref{t:13_stars}).
At 1.3 \kpc, 5 of them are runaway stars exceeding 20 \kms 2--D peculiar velocity.
This is consistent with the general ratio of the runaway stars to the normal stars which is 10--30 percent
\citep{1987ApJS...64..545G}.
\begin{table*}
\caption{24 stars beyond 1 \kpc are presented. Angular distances are given in degrees, distances in parsecs, extinctions and brightness in visual band in magnitudes.
In the column; SpT adopted, the average spectral types that are used in distance calibrations are given.
}
\label{t:26_stars}
\begin{center}
\begin{tabular}{@{}c@{~~~}c@{~~~}c@{~~~}c@{~~~}c@{~~~}c@{~~~}c@{~~~}c@{~~~}c@{~~}r@{~/~}lc@{~~~}c@{~~~}c@{~~~}c@{}}
Ang. Sep.
& Name
& SpT
& SpT
& $\mathrm{A_{V}}$
& $\mathrm{A_{V}}$
& E(B-V)
& E(B-V)
& Distance
& \multicolumn{2}{c}{Distance}
& V
& Ref\#$^*$ \\
&
&
& (adopted)
&
& (Err)
&
& (Err)
&
& \multicolumn{2}{c}{(Err)}
&
& \\\hline
0.09 & TYC 1869-01317-1 & B9.5V & B9.5V & 0.83 & 0.1 & 0.26 & 0.090 & 1013 & +183&-174 & 11.258 & tw \\
0.09 & TYC 1869-01642-1 & B9.5V & B9.5V & 1.00 & 0.1 & 0.3 & 0.090 & 952 & +172&-163 & 11.293 & tw \\
0.17 & HD 37424 & B0.5V & B0.5V & 1.28 & 0.06 & 0.35 & 0.036 & 1868 & +410&-403 & 8.989 & tw \\
0.52 & HD 36993 & B0.5/1III/IV & B0.5III-IV & 1.35 & 0.06 & 0.39 & 0.028 & 1961 & +636&-628 & 8.248 & 1 \\
0.63 & HD 37318 & B0.5Ve & B0.5V & 2.29 & 0.15 & 0.59 & 0.026 & 903 & +231&-228 & 8.399 & 1 \\
1.00 & BD+27 797 & B0.5Ve & B0.5V & 2.50 & 0.14 & 0.62 & 0.094 & 1788 & +485&-465 & 10.095 & 2 \\
1.35 & HD 37696 & B0.5IV/V & B0.5IV-V & 1.12 & 0.06 & 0.29 & 0.023 & 1549 & +565&-558 & 7.973 & 1 \\
1.35 & BD+27 850 & B1.5IVe & B1.5V & 1.31 & 0.04 & 0.35 & 0.056 & 1525 & +313&-305 & 9.362 & 2 \\
1.52 & HD 245770 & O9/B0III/Ve & B0III-V & 2.11 & 0.13 & 0.79 & 0.053 & 2595 & +1090&-1066 & 9.187 & 3 \\
1.76 & HD 36441 & B0.5/1.5V & B1V & 1.13 & 0.13 & 0.33 & 0.058 & 1153 & +628&-389 & 8.239 & 1 \\
1.97 & BD+26 943 & B2V & B1.5V & 1.41 & 0.14 & 0.37 & 0.040 & 1357 & +723&-511 & 9.623 & 8 \\
2.61 & HD 38010 & B1III & B1III & 1.29 & 0.1 & 0.22 & 0.028 & 967 & +437&-236 & 6.817 & 4 \\
2.95 & BD+30 938 & B3III & B3III & 1.44 & 0.04 & 0.45 & 0.030 & 1332 & +741&-255 & 9.062 & 7 \\
2.98 & Sh 2-242 1 & B0V & B0V & 2.17 & 0.13 & 0.78 & 0.056 & 2318 & +853&-842 & 9.996 & 5 \\
3.06 & HD 37366 & O9.5V & O9.5V & 1.22 & 0.06 & 0.35 & 0.016 & 1375 & +201&-199 & 7.640 & 6 \\
3.07 & BD+30 987 & B5III: & B5III & 0.73 & 0.08 & 0.25 & 0.044 & 1523 & +398&-287 & 9.444 & 7 \\
3.24 & BD+30 976 & B7V & B7V & 0.89 & 0.17 & 0.18 & 0.020 & 994 & +264&-226 & 10.277 & 7 \\
3.45 & BD+25 989 & B1Vn & B1V & 1.59 & 0.13 & 0.43 & 0.082 & 1872 & +1061&-649 & 9.751 & 8 \\
3.61 & BD+27 909 & B2III & B2III & 1.87 & 0.1 & 0.55 & 0.076 & 2004 & +683&-761 & 9.479 & 8 \\
4.05 & BD+31 1065 & B3III & B3III & 0.67 & 0.04 & 0.27 & 0.043 & 2304 & +1286&-442 & 9.482 & 7 \\
4.07 & BD+31 1050 & B3III & B3III & 1.01 & 0.03 & 0.35 & 0.047 & 1791 & +988&-338 & 9.276 & 7 \\
4.09 & HD 38909 & B3II-III & B3II-III & 0.56 & 0.03 & 0.17 & 0.033 & 2057 & +1007&-992 & 8.145 & 9 \\
4.28 & BD+31 1021 & B7V & B7V & 1.05 & 0.12 & 0.14 & 0.106 & 1007 & +250&-218 & 10.465 & 7 \\
4.46 & HD 40297 & B9.5Ib/II & B9.5Ib/II & 1.07 & 0.13 & 0.25 & 0.015 & 1337 & +693&-688 & 7.270 & 1 \\\hline
\multicolumn{13}{p{15cm}}{%
$^*$ Ref\#: %
1: \cite{1979RA......9..479C}; %
2: \cite{1999A+AS..137..147S}; %
3: \cite{1998PASP..110.1310W}; %
4: \cite{1976A+AS...26..241C}; %
5: \cite{1990AJ.....99..846H}; %
6: \cite{1971ApJS...23..257W}; %
7: \cite{1961AnTou..28...33B}; %
8: \cite{1977ApJ...217..127C}; %
9: \cite{1955ApJS....2...41M}; %
10: \cite{1977ApJ...217..127C}; %
tw: This work.} \\
\end{tabular}
\end{center}
\end{table*}
\begin{table}
\caption{Visual absolute magnitudes at 1.3$\pm$0.1 \kpc for the stars of the possible OB association of which HD 37424 is a member.
Errors are mainly due to the distance range 1.2$-$1.4 \kpc.
The expected M$_{V}$ for the corresponding spectral types and its dispersion from
\citet{2006MNRAS.371..185W} (W06)
are given. All values are in mag.}
\label{t:M_stars}
\begin{center}
\begin{tabular}{@{}c@{~~}c@{~}r@{~/~}l@{~~}c@{~~}c@{}}
& $\mathrm{M_{V}}$
& \multicolumn{2}{c}{error}
& $\mathrm{M_{V}}$
& Disp. \\
Name
& (@1.3$\pm$0.1 \kpc)
& \multicolumn{2}{c}{~}
& (W06)
& (W06) \\\hline
TYC 1869-01317-1 & -0.14 & +0.26&-0.27 & +0.29 & 1.40 \\
TYC 1869-01642-1 & -0.28 & +0.26&-0.27 & +0.29 & 1.40 \\
HD 37424 & -2.86 & +0.22&-0.23 & -3.34 & 2.40 \\
HD 36993 & -3.67 & +0.22&-0.23 & -4.02 & 2.30 \\
HD 37318 & -4.46 & +0.31&-0.32 & -3.34 & 2.40 \\
BD+27 797 & -2.97 & +0.30&-0.31 & -3.34 & 2.40 \\
HD 37696 & -3.72 & +0.22&-0.23 & -3.34 & 2.40 \\
BD+27 850 & -2.52 & +0.20&-0.21 & -2.95 & 1.16 \\
HD 245770 & -3.49 & +0.29&-0.30 & -3.34 & 2.40 \\
HD 36441 & -3.46 & +0.29&-0.30 & -2.95 & 1.16 \\
BD+26 943 & -2.36 & +0.30&-0.31 & -2.64 & 1.40 \\
HD 38010 & -5.04 & +0.26&-0.27 & -4.10 & 2.20 \\
BD+30 938 & -2.95 & +0.20&-0.21 & -2.32 & 1.50 \\
Sh 2-242 1 & -2.74 & +0.29&-0.30 & -3.34 & 2.40 \\
HD 37366 & -4.15 & +0.22&-0.23 & -4.49 & 2.27 \\
BD+30 987 & -1.86 & +0.24&-0.25 & -1.49 & 2.00 \\
BD+30 976 & -1.18 & +0.33&-0.34 & -0.63 & 1.40 \\
BD+25 989 & -2.41 & +0.29&-0.30 & -2.95 & 1.16 \\
BD+27 909 & -2.96 & +0.26&-0.27 & -2.63 & 2.20 \\
BD+31 1065 & -1.76 & +0.20&-0.21 & -2.32 & 1.50 \\
BD+31 1050 & -2.30 & +0.19&-0.20 & -2.32 & 1.50 \\
HD 38909 & -2.98 & +0.19&-0.20 & -2.32 & 1.50 \\
BD+31 1021 & -1.15 & +0.28&-0.29 & -0.63 & 1.40 \\
HD 40297 & -4.37 & +0.29&-0.30 & -3.75 & 1.30 \\\hline
\end{tabular}
\end{center}
\end{table}
\begin{table}
\caption{19 stars with common proper motion are presented.
The average PM$_{\text{RA}}$ is -1.39 \masyr and PM$_{\text{Dec}}$ is -4.17 \masyr.
The last three column represents the 2--D space velocity of
the stars with respect to the average motion of the group with maximum and minimum values in \kms.
All the values of proper motions are in \masyr.}
\label{t:13_stars}
\begin{center}
\begin{tabular}{c c c c c c c c}
Name & $\mu_{\alpha}^{*}$ & err & $\mu_{\delta}$ & err & $\mathrm{V_{REL}}$ & max & min \\\hline
HD 37318 & -0.8 & 0.6 & -5.9 & 0.6 & 11.2 & 16.1 & 6.9 \\
TYC 1869-01642-1 & -1.5 & 1.1 & -2.4 & 1.2 & 11.0 & 19.8 & 7.1 \\
HD 38010 & -1.7 & 1.0 & -3.3 & 1.0 & 5.7 & 14.1 & 0.0 \\
BD+30 976 & -3.5 & 0.9 & -5.3 & 0.5 & 14.7 & 21.1 & 8.4 \\
BD+31 1021 & -2.8 & 0.7 & -5.5 & 0.8 & 11.9 & 18.4 & 5.4 \\
TYC 1869-01317-1 & -0.9 & 0.9 & -5.1 & 0.9 & 6.5 & 14.2 & 0.0 \\
HD 36441 & -2.0 & 0.6 & -6.6 & 0.6 & 15.4 & 20.1 & 11.2\\
BD+30 938 & -1.7 & 0.8 & -3.6 & 0.6 & 4.0 & 9.9 & 0.0 \\
HD 40297 & -1.4 & 1.0 & -3.7 & 1.0 & 2.9 & 11.0 & 0.0 \\
BD+26 943 & -0.9 & 0.8 & -3.7 & 0.8 & 4.2 & 11.2 & 0.0 \\
BD+30 987 & -2.0 & 0.6 & -4.7 & 1.2 & 4.9 & 13.0 & 0.0 \\
BD+27 797 & -0.1 & 0.7 & -3.1 & 1.4 & 10.4 & 19.6 & 4.2 \\
BD+31 1050 & -1.0 & 0.6 & -4.0 & 0.6 & 2.7 & 7.8 & 0.0 \\
BD+25 989 & -1.5 & 0.9 & -2.5 & 1.6 & 10.3 & 21.1 & 4.9 \\
HD 36993 & -0.2 & 0.9 & -5.8 & 0.6 & 12.4 & 18.8 & 6.6 \\
HD 38909 & +0.7 & 0.5 & -4.2 & 1.5 & 12.9 & 18.4 & 0.0 \\
BD+31 1065 & -1.0 & 0.6 & -1.3 & 0.5 & 17.9 & 21.7 & 14.7\\
Sh 2-242 1 & -0.1 & 0.7 & -1.8 & 1.0 & 16.7 & 24.2 & 9.2 \\
HD 245770 & -2.0 & 0.5 & -4.3 & 1.1 & 3.8 & 10.2 & 0.0 \\\hline
\end{tabular}
\end{center}
\end{table}
\section{Discussion}
The runaway nature of HD\,37424 is clear.
The chance projection of such a massive runaway star moving away from the GC of an SNR in the Galactic anti--center direction must be very low.
In addition, combining with the central compact object after tracing back both objects at the same time
shows that HD\,37424 is clearly the pre--supernova binary companion of the progenitor of SNR S147 and PSR\,J0538+2817.
BSS is the favored explanation for its runaway nature.
HD\,37424 is a B0.5V type star with a mass of $\sim$13 \msun
\citep{2010AN....331..349H}.
So, the progenitor of the pulsar must have a higher mass.
Based on the lack of O type stars in the field (see Table \ref{t:26_stars})
we set an upper mass limit of 20--25 \msun.
It may even imply a twin binary.
The Roche Lobe radii calculated for 15, 20 and 25 \msun vary between 91 and 311 \rsun which shows that the system might have been an interacting binary.
Hence, the progenitor star should be a naked helium star at the final stage of its evolution with a mass even as low as 2 \msun
\citep{1993SSRv...66..309V},
\citep{1995ApJ...448..315W}.
However, how conservative the mass transfer was, will be understood after further observations.
Assuming a circular orbit, pre--supernova binary parameters are calculated for 2, 5, 10, 15, 20 and 25 \msun (Table \ref{t:binary}) progenitor masses.
\begin{table*}
\caption{The binary separations, the orbital velocities of the progenitor, the orbital periods
and the Roche Lobe radii of the pre--supernova binary for various final masses of the progenitor with 13 \msun HD\,37424 are given.
Roche Lobe radii are calculated based on
\citet{1983ApJ...268..368E}.}
\label{t:binary}
\begin{center}
\begin{tabular}{c r@{~}l r@{~}l r@{~}l r@{~}l r@{~}l r@{~}l}
Progenitor Mass (\msun)
& \multicolumn{2}{c}{ 2}
& \multicolumn{2}{c}{ 5}
& \multicolumn{2}{c}{10}
& \multicolumn{2}{c}{15}
& \multicolumn{2}{c}{20}
& \multicolumn{2}{c}{25} \\\hline
\noalign{\smallskip}
Binary Separation (\rsun)
& 9 & $\pmpm{-1}{+3}$
& 49 & $\pmpm{-9}{+11}$
& 152 & $\pmpm{-26}{+37}$
& 281 & $\pmpm{-49}{+68}$
& 425 & $\pmpm{-75}{+101}$
& 576 & $\pmpm{-101}{+137}$ \\
\noalign{\smallskip}
Orbital Velocity (\kms)
& 481 & $\pm$49
& 192 & $\pm$20
& 96 & $\pm$10
& 64 & $\pm$7
& 48 & $\pm$5
& 38 & $\pm$4 \\
\noalign{\smallskip}
Orbital Period (days)
& 0.85 & $\pmpm{-0.22}{+0.32}$
& 9 & $\pmpm{-2}{+4}$
& 45 & $\pmpm{-11}{+17}$
& 103 & $\pmpm{-26}{+39}$
& 176 & $\pmpm{-44}{+66}$
& 259 & $\pmpm{-65}{+98}$ \\
\noalign{\smallskip}
Roche Lobe Radius (\rsun)
& &
& &
& &
& 110 & $\pmpm{-19}{+27}$
& 177 & $\pmpm{-31}{+42}$
& 251 & $\pmpm{-44}{+60}$ \\\hline
\noalign{\smallskip}
\end{tabular}
\end{center}
\end{table*}
As discussed in the previous section, the OB stars around might be members of an unidentified old OB association
of which all of the O type stars underwent supernova explosions.
This also makes a plausible explanation for the low density medium in which the SNR expands symmetrically.
But, an ejection of the pre-SN system is also possible.
A membership to an OB association or to a cluster is important also regarding the distance determination.
In this work, the distance derived from pulsar parallax ($1.3\pmpm{+0.20}{-0.16}$ \kpc) is accepted as the most reliable estimation.
Also, the distance to the star measured from interstellar lines is in the same range, $1288\pmpm{+304}{-193}$ (see section 3).
The spectro--photometric distance is much larger by using absolute magnitudes from \citet{1982lbg6.conf.....A}.
Yet,
by using typical luminosities for B0.5V type suggested in
\citet{2010AN....331..349H},
it is 1318$\pm$119 \pc.
Hence, the distance to the star and the SNR can be assumed to be 1.3 \kpc.
However, the \Av measured directly towards S147 is much lower than the \Av towards the stars beyond 1 \kpc.
Furthermore, two stars, HD\,36665 and HD\,37318, show highly shifted interstellar \ion{Ca}{ii} and \ion{Na}{i} lines
related to the SNR implying that these objects are background sources
\citep{2004A&A...426..555S}.
Their distances based on the reported spectral types are closer to the Sun than HD\,37424 is.
HD\,36665 has $837\pmpm{+347}{-285}$ \pc for B1V type and HD\,37318 is $903\pmpm{+231}{-228}$ \pc far away adopting B0.5V.
On the other hand, the distance for HD\,36665 is identified as 1860 \pc in
\citet{1980BICDS..19...61N} through H$_{\beta}$ measurements.
It has a high \Av of 1.74 mag.
If we assume that HD\,36665 is also at 1.3$\pm$0.1 \kpc,
then it must be a bright B1V type star which is 1--1.5 mag brighter than the average value given in W06.
HD\,37318 also can be a member of the possible OB association.
The supernova event had occurred quite nearby in a fairly reddened medium.
Assuming a very faint SN (\Mv = $-14$ mag), the apparent visual magnitude is $-2.1$ mag, and
for a bright SN with \Mv$ = -21$ mag, the apparent visual magnitude is $-9.1$ mag; as bright as SN\,1006.
The event might have also caused the $^{10}\mathrm{Be}$ peak at 35$\pm$2 \kyr measured in
\citet{1987Natur.326..273R}
from the deep ice cores from Dome C and Vostok Antarctica.
\section{Conclusion}
HD\,37424 is a B0.5V type runaway star from a binary ejection due to the supernova which gave birth to SNR S147 and PSR\,J0538+2817.
The star has $-9.2\pm6.5$ \kms heliocentric radial velocity and 74$\pm$8 \kms 3--D peculiar velocity.
The distance calculated from interstellar \ion{Ca}{ii} lines is $1288\pmpm{+304}{-193}$ \pc and from spectro--photometry 1318$\pm$119 \pc.
No high velocity gas related with the SNR is detected in the spectra.
The past trajectories of the pulsar and HD\,37424 are reconstructed throughout Monte Carlo simulations.
It is found that both stars were at the same position at
$\left(l, b\right)=\left(84.82\pm0.01, 27.84\pm0.01\right)$ deg at $30\pm4$ \kyr in the past.
The position of the explosion is $4.2\pmpm{+0.8}{-0.6}$\,arcmin away from the geometrical center.
The distance of the SNR is found as 1333$\pmpm{+103}{-112}$ \pc.
\Av towards the SNR is 1.28$\pm$0.06 mag.
Today's kinematics of the stars are well known,
further detailed calculations should be done to find the true kick vectors of the pulsar and a possible spin-kick alignment.
There is no known OB association reported close to the SNR.
19 OB type stars within 100\pc from the SNR geometrical center have very similar 2--D velocities with the 5 runaway stars including HD\,37424.
This might be an old OB association having no bright O type stars.
The low density medium in which the SNR is expanding is probably due to the previous SNe and H II regions
driven by this old association.
The progenitor was a massive star with a zero age main sequence mass greater than 13 \msun.
Considering the lack of O type stars in the field, a progenitor mass much larger than 20 \msun is not expected.
Assuming a circular orbit, the pre--supernova binary separation is in the range of 8 to 711 \rsun for 2 to 25 \msun final mass of the progenitor.
The corresponding Roche Lobe radii for 15 to 25 \msun masses vary from 91 to 311 \rsun.
It must have been an interacting binary.
For 1.3 kpc distance and 1.3 mag extinction, the SN which happened 30$\pm$4 kyr ago had an apparent brightness of $-2.1$ to $-9.1$ mag.
The source will be investigated regarding the elemental signatures of binary accretion and
the possible supernova debris on its photosphere through high resolution and high S/N spectra.
\section*{Acknowledgements}
This work was conceived by the late Oktay H. GUSE\.{I}NOV (1938-2009).
The authors acknowledge the leading role he had in all stages of this work and other related topics.
His contribution to Galactic Astrophysics and his scholarship will be greatly missed.
We thank Janos Schmidt and Christian Ginski for their observational support and Ronny Errmann for useful discussions.
BD, NT and RN acknowledge support from DFG in the SFB/TR-7 Gravitational Wave Astronomy.
This research has made use of Aladin.
We would like to thank the Calar Alto observatory staff
for their help in our observations;
and we would like to thank DFG in NE 515 / 53-1 for
financial support for the Calar Alto observing run.
We also thank \tubitak National Observatory for their supports in using RTT--150 with project number 12ARTT150-267.
|
\section{Introduction}
Online social networks such as Facebook foster the aggregation of people around common interests, narratives, and worldviews. Indeed, the World Wide Web caused a shift of paradigm in the production and consumption of contents that increased volume and heterogeneity of available contents. Users can express their attitudes by producing and consuming heterogeneous information --- e.g. conspiracists avoid mainstream news and follow their own information sources, whereas debunkers try to inhibit the diffusion of false claims.
Images of kittens and pets, political memes, gossip, scandals spread on Facebook.
By liking, commenting, and sharing their preferred contents, users can express their passions and emotions --- and, among these latter, sarcasm in not an exception.
Indeed, not rarely we can find pages promoting parodistic and sarcastic imitations of online social dynamics --- e.g., {\em Ebola and Kittens} \cite{kittens} or {\em In favor of chem-trails} \cite{chemtrails}.
An interesting case in the Italian Facebook is a page \cite{TotoCutugno} with more than $40K$ followers that posts everyday the exactly alike picture of Toto Cutugno, a famous Italian pop-singer.
In this work, we use the intriguing case of that page as a baseline to study and model the effect of content diversity on popularity.
Specifically, we analyze user activity and post consumption patterns on the baseline page for a timespan of about $4$ months.
Through a comparative analysis between two sets of pages producing heterogeneous contents, we show that there are no remarkable differences in user activity patterns, whereas significant dissimilarities between post consumption patterns emerge.
Such a comparative analysis allows to derive a model of information consumption accounting for the heterogeneity of contents.
Hence, we show that the proposed model is able to reproduce the phenomenon observed from empirical data. In particular, we show the effects of different levels of contents' heterogeneity on posts consumption patterns.
\section{Background and Related Works}
A large body of literature addresses the study of social dynamics on socio-technical systems from social contagion up to social reinforcement \cite{Onnela2010,Ugander2012,Lewis2012,Mocanu2012,Adamic2005,Kleinberg2013,eRep,Quattrociocchi2011,Quattrociocchi2009,Bond2012,Moreno2011,Centola2010,Castellano2007,Quattrociocchi2014,Bennaim2003,Adamic2013,Hannak2012,Cheng2014, Goncalves2012}.
Among these, one of the most defining topic of computational social science is the understanding of driving forces behind the popularity of contents \cite{tatar2014survey}.
Such a challenge has been addressed looking at the sentiment of comments, contents, or users' attention \cite{tatar2011predicting,figueiredo2014does,zadeh2014modeling,ratkiewicz2011detecting,bandari2012pulse,gomez2008statistical,szabo2010predicting,lerman2010using,lerman2010information}.
However, the mechanisms behind popularity remain largely unexplored \cite{Goldhaber1997,Watts2002,Leskovec2009}.
In \cite{salganik2006experimental} the authors address such a challenge experimentally by measuring the
impact of content quality and social influence on the eventual popularity or success of cultural artifacts.
The effects of specific contents on the formation of communities of interest, their permeability to false information, and the resistance to changes have been recently characterized in \cite{JTM2014,Rojecki2014,SOCINFO2014, PLOS2014}. In particular, in \cite{WWW2015} the authors point out that connectivity patterns of the Facebook social network are prominently driven by homophily of users --- i.e., the tendency of individuals to associate with similar others --- towards specific kinds of contents. Microblogging platforms such as Facebook and Twitter \cite{Tapscott2006} have lowered the cost of information production and broadcasting, boosting the potential reach of each idea or meme \cite{Dawkins1989,Bauckhage2011} --- i.e., content or concepts that spread rapidly on the Web. Still, the abundance of information to which we are exposed through online social networks and other socio-technical systems is exceeding our capacity to consume it \cite{Weng2012}.
As a result, the dynamics of information is driven more than ever before by the economy of attention \cite{Simon1971,dukas2001limited, Lehmann2012}.
We address this challenge by studying the interlink between contents diversity and popularity. More specifically, we investigate the effects of sources producing always the same information on users' activity, consumption, and attention patterns.
\section{Data Description}
In this work, we aim at investigating the role of content diversity on the dynamics of information consumption in online social networks.
To this end, we use a set of Facebook pages promoting heterogeneous contents and a Facebook page promoting always the same picture. The set of pages promoting heterogeneous contents is composed by $73$ public Facebook pages, whereof $34$ are about scientific news and $39$ about conspiratorial news; we refer to the former as \emph{science pages} and to the latter as \emph{conspiracy pages}. The page promoting homogeneous contents is called "La stessa foto di Toto Cutugno ogni giorno" ("Everyday the same photo of Toto Cutugno", a well-known Italian pop singer-songwriter); such a page, by publishing everyday the same picture of the Italian singer --- and nothing else ---, represents the perfect control for studying content diversity; we refer to this page as the \emph{baseline page}.
Starting from these pages, we downloaded all the posts, we collected all the \emph{likes} and \emph{comments} to the posts, and we counted the number of \emph{shares}. Data related to science and conspiracy pages have been collected from August 22, 2013 to December 31, 2013, whereas data related to the baseline page have been collected from August 22, 2014 (birthdate of the page) to December 31, 2014.
In total, we collected around $2M$ likes and $190K$ comments, made by about $340K$ and $65K$ users, respectively.
In Table \ref{tab:1} we summarize the details of our data collection.
Likes, shares, and comments have a different meaning from the user viewpoint. Most of the times, a like stands for a positive feedback to the post; a share expresses the will to increase the visibility of a given information; and a comment is the way in which online collective debates take form. Comments may contain negative or positive feedbacks with respect to the post.
\begin{center}
\begin{table}[!h]
\centering
\tiny
\begin{tabular}{l|c|c|c|c}
\hline\bf { } & \bf {Total} & \bf {Science} & \bf {Conspiracy} & \bf {Baseline} \\ \hline
Pages & $ 74 $ & $ 34 $ & $ 39 $ & $1$ \\
Posts & $ 49,354 $ & $ 13,028 $ & $ 36,169 $ & $157$ \\
Likes & $ 2,095,677 $ & $ 614,078 $ & $ 1,184,084$ & $297,515$ \\
Comments & $192,967 $ & $ 40,608 $ & $ 138,138 $ & $14,221$ \\
Shares & $3,782,480$ & $477,457$ & $3,297,687$ & $7,336$ \\
Likers & $ 344,367 $ & $ 162,146 $ & $ 159,524 $ & $22,697$ \\
Commenters & $ 64,903 $ & $ 18,358 $ & $ 41,666 $ & $4,875$ \\
\end{tabular}\newline
\caption{ \textbf{Dataset breakdown.} The number of pages,
posts, likes, comments, shares, likers, and commenters for science pages, conspiracy pages, and the baseline page.}
\label{tab:1}
\end{table}
\end{center}
\section{Results and Discussion}
In this section, we first present the statistical signatures characterizing users activity on pages with diversified content on specific topics (science and conspiracy news) against the case of the page posting every day the same picture (baseline). Then we derive a model of information consumption mimicking user preferences with respect to contents.
\subsection{Content and Users Activity}
Let us focus some regularities concerning users' activity on science pages and conspiracy pages compared with the baseline page.
Figure \ref{fig:2} shows the probability density function (PDF) for the normalized\footnote{We performed the unity--based normalization to bring all values in the range $[0,1]$.} number of likes for each user. We find that the activity of users presents an heavy--tailed distribution.
\begin{figure}[h]
\centering\includegraphics[width = 0.45\textwidth]{1.jpg}
\caption{\textbf{Users' activity patterns.} Probability density function (PDF) for the normalized number of likes by each user.}
\label{fig:2}
\end{figure}
In Figure \ref{fig:3} we show the PDF of the users' lifetime in terms of their liking activity --- i.e. the temporal interval between the first and the last like of the user on a given page. We find a slight difference in the lifetime of the baseline users with respect to science and conspiracy users.
\begin{figure}[h]
\centering\includegraphics[width = 0.45\textwidth]{2.jpg}
\caption{\textbf{Users' lifetime.} Probability density function (PDF) of the users' lifetime in terms of their liking activity. The PDF shows a slight difference in the lifetime of the baseline users with respect to science and conspiracy users.
}
\label{fig:3}
\end{figure}
These figures show that users activity patterns are similar and present heavy--tailed distributions despite the different nature of the contents, and we can not find any significant difference between the users interaction patterns induced by heterogeneous or homogeneous contents.
Conversely, by analyzing consumption patterns related to posts, we find a significant difference in the information consumption dynamics. Figure \ref{fig:5} shows the PDF for the number of likes received by posts belonging to science pages, conspiracy pages, and the baseline page.
\begin{figure}[h]
\centering\includegraphics[width = 0.45\textwidth]{3.jpg}
\caption{\textbf{Posts' consumption patterns.} Probability distribution function (PDF) for the normalized number of likes received by posts belonging to science pages, conspiracy pages, and the baseline page. The PDFs show remarkable differences between consumption patterns' distributions related to pages promoting heterogeneous contents and those related to the page promoting homogeneous contents.}
\label{fig:5}
\end{figure}
The number of likes received by posts are heavy--tailed distributed if the posts belong to pages promoting heterogeneous contents (science and conspiracy pages); whereas they are approximately distributed according to a Gaussian if the posts belong to a page promoting homogeneous content (baseline page).
Summarizing, users' activities always present heavy--tailed distributions resolving in heavy--tailed distributed consumption patterns on posts in the heterogeneous contents case. Still, when the content promoted by a page is homogeneous -- i.e., always the same -- we find that the heavy--tailed distributed users' activities resolve in posts' consumption patterns that are approximately Gaussian.
\subsection{Modeling Contents Consumption}
Here we introduce a model of pattern consumption that exploits the Beta distribution properties to generate different levels of posts' attractiveness, thus varying content--heterogeneity in the simulated collection of posts.
The Beta distribution is a family of continuous probability distributions defined in the interval $[0,1]$ and characterized by two real parameters, $\alpha > 0$ and $\beta > 0$, which control the shape of the distribution. In particular, for $\alpha = 1$ and $\beta = 1$ the Beta distribution $\mathcal{B}e(\alpha,\beta)$ is equivalent to the Uniform distribution $\mathcal{U}(0,1)$.
Conversely, if $\alpha = 1$ and $\beta \gtrsim 20$, the Beta distribution $\mathcal{B}e(\alpha,\beta)$ is a right heavy--tailed distribution. Figure \ref{fig:1} shows the Beta probability density function with respect to the two shape parameters $\alpha$ and $\beta$.
\begin{figure}[h]
\centering\includegraphics[width = 0.45\textwidth]{4.jpg}
\caption{\textbf{Beta distribution $\mathcal{B}e(\alpha,\beta)$.} Two parameters, $\alpha$ and $\beta$, control the shape of the distribution. In particular, for $\alpha = 1$ and $\beta = 1$ the Beta distribution $\mathcal{B}e(\alpha,\beta)$ is equivalent to the Uniform distribution $\mathcal{U}(0,1)$. Conversely, if $\alpha = 1$ and $\beta \gtrsim 20$, the Beta distribution $\mathcal{B}e(\alpha,\beta)$ is a right heavy--tailed distribution.}
\label{fig:1}
\end{figure}
In our model, each post has a value drawn from a Beta distribution $v \sim \mathcal{B}e(1, \beta)$, with $\beta$ ranging between $1$ and $1,000,000$, indicating its attractiveness. We let the parameter $\beta$ assume those extreme values in order to obtain different distributions for posts' attractiveness.
Indeed, notice that when $\beta = 1$ the Beta distribution $\mathcal{B}e(1, \beta)$ is equivalent to a uniform distribution $\mathcal{U}(0,1)$, so that we have a collection of homogeneous--content posts --- i.e., each post has the same degree of attractiveness; whereas when $\beta \to \infty$ the Beta distribution $\mathcal{B}e(1, \beta)$ is equivalent to a right heavy--tailed distribution, so that we have a collection of heterogeneous--content posts --- i.e., there are few posts with a high level of attractiveness, while the vast majority of the posts is characterized by a low level of attractiveness. Moreover, each user is characterized by two parameters randomly drawn from power law distributions: her volume of activity, $a \sim p(x)$; and her fixed--preference about the posts, $b \sim p(x)$, where $p(x) = x^{-\gamma}$ with $\gamma = 1.5$.
Each user can not exceed her assigned volume of activity, $a$, and she likes a given post if and only if her normalized\footnote{Note that we performed a unity--based normalization in order to bring all values of $b \sim p(x) = x^{-1.5}$ in the range $[0,1]$, so that the fixed--preference of the user is comparable with the attractiveness of the posts.} fixed--preference, $b$, is smaller than the attractiveness, $v$, of that post. Note that in our model we do not take into account the users' network: since Facebook network is very dense --- indeed, the diameter of Facebook social network is $3.74$ \cite{Backstrom2012,WWW2015} --- the connections between users are not likely to influence posts' consumption dynamics.
We run simulations for $\beta$ ranging between $1$ and $1,000,000$, with $P = 10,000$ (posts) and $U = 20,000$ (users). Results are averaged over $100$ iterations.
Figure \ref{fig:7} shows the probability density function (PDF) of the users activity and the posts consumption patterns generated by a simulation of the model with $\beta = 1,000,000$ --- i.e., in the case of extremely heterogeneous--content posts. Observe that users' activity is heavy--tailed, and the distribution of posts' consumption is skewed. Such a result is consistent with empirical data shown in the previous section: if the content promoted by a page is heterogeneous, the heavy--tailed users' activity resolves in skewed posts consumption's patterns.
\begin{figure*}
\centering\includegraphics[width = 0.85\textwidth]{5.jpg}
\caption{\textbf{Users activity and post consumption patterns with extremely heterogenous--content posts.} Probability density function (PDF) of the users activity and the posts consumption patterns generated by a simulation of the model with $\beta = 1,000,000$. If the content promoted by a page is heterogeneous, the heavy--tailed users' activity resolves in skewed posts consumption's patterns.}
\label{fig:7}
\end{figure*}
Figure \ref{fig:8} shows the probability density function (PDF) of the users activity and the posts consumption patterns generated by a simulation of the model with $\beta = 1$ --- i.e., in the case of homogeneous--content posts. Notice that users' activity is heavy--tailed, whereas posts' consumption is approximately Gaussian. Such a result is consistent with empirical data shown in the previous section: if the content promoted by a page is always the same, the heavy--tailed users' activity resolves in approximately Gaussian posts consumption's patterns.
\begin{figure*}
\centering\includegraphics[width = 0.85\textwidth]{6.jpg}
\caption{\textbf{Users activity and post consumption patterns with homogeneous--content posts.} Probability density function (PDF) of the users activity and the posts consumption patterns generated by a simulation of the model with $\beta = 1$. If the content promoted by a page is always the same, the heavy--tailed users' activity resolves in approximately Gaussian posts consumption's patterns.}
\label{fig:8}
\end{figure*}
\section{Concluding Remarks}
Facebook is overflowed by different and heterogeneous contents, from the latest news up to satirical and funny stories. Each piece of that corpus reflects the heterogeneity of the underlying social background. Indeed, the World Wide Web caused a shift of paradigm in the production and consumption of information that increased the amount and heterogeneity of contents available to users. On online social networks such as Twitter and Facebook, people can express their attitudes, passions, and emotions by producing and consuming heterogeneous information.
In the Italian Facebook, we have found a fascinating case of contents' homogeneity: a page with more than $40K$ followers that every day posts the same picture of Toto Cutugno, a popular Italian singer. In this work, we use such a page as a benchmark to investigate and model the effect of contents’ heterogeneity on popularity. In particular, we use that page for a comparative analysis of information consumption patterns with respect to pages posting heterogeneous contents related to science and conspiracy.
We show that there are not remarkable differences in user activity patterns, whereas we find significant dissimilarities between post consumption patterns of the page promoting homogeneous contents and those of the pages producing heterogeneous contents. Finally, we derive a model of information consumption that accounts for the heterogeneity of contents. Hence, we show that the proposed model is able to reproduce the phenomenon observed
from empirical data.
\section{Acknowledgments}
Funding for this work was provided by EU FET project MULTIPLEX nr. 317532 and SIMPOL nr. 610704. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.
We want to thank Prof. Guido Caldarelli for precious insights and contribution on the data analysis.
Special thanks to Josif Stalin, Stefano Alpi, Michele Degani for giving access to the Facebook page of {\em La stessa foto di Toto Cutugno ogni giorno}.
\bibliographystyle{unsrt}
|
\section{Introduction}
The notion of large financial market was introduced by Kabanov and Kramkov in \cite{Kakra94} as a sequence of ordinary security market models. A suitable property for such kind of markets is the absence of asymptotic arbitrage opportunities. In the frictionless case, a standard assumption is that each small market is free of arbitrage. If, in addition, the small markets are complete, then the existence of asymptotic arbitrage opportunities is related to some contiguity properties of the sequence of equivalent martingale measures (see \cite{Kakra94}). These results are extended to the case of incomplete markets by Klein and Schachermayer in \cite{K:S:1996, K:Sch:1996} and by Kabanov and Kramkov in \cite{Kab:Kra:1998}. In the transaction costs case, i.e. when each small market is subject to proportional transaction costs, the standard assumption is that each small market is free of arbitrage under arbitrarily small transaction costs. In this context, characterizations of the existence of asymptotic arbitrage, similar to those in the frictionless case, can be found in \cite{Kl:Le:Pe}.
In this paper, we consider a non-standard large financial market, i.e. a sequence of market models which admit arbitrage for sufficiently small transaction costs, and we study its asymptotic arbitrage opportunities. Our large financial market is given by the sequence of binary markets approximating the fractional Black-Scholes model introduced by Sottinen in \cite{Sotti}. We refer to this large financial market as the large fractional binary market. Similarly, we call $N$-fractional binary market to the $N$-period binary market in the sequence. Sottinen proves in \cite{Sotti} that, for $N$ sufficiently large, the $N$-fractional binary market admits arbitrage. From the results in \cite{CKO}, we conclude that the smallest transaction cost, $\lambda_c^N$, needed to eliminate the arbitrage in the $N$-fractional binary market is strictly positive. Moreover, from \cite{CP}, we know that $\lambda_c^N$ converges to $1$, a result which contrasts with the fact that the fractional Black-Scholes model is free of arbitrage under arbitrarily small transaction costs. This is not a true contradiction, since the arbitrage strategies constructed in \cite{CP} provide profits, under big transaction costs, with probabilities vanishing in the limit. As explained in \cite{CP}, a more appropriate way to compare the arbitrage opportunities in the sequence of fractional binary markets with the arbitrage opportunities in the fractional Black-Scholes market is to study the problem for the former from the perspective of the large financial markets. A first step in this direction was done in \cite{CP}, where the authors study the existence of asymptotic arbitrage of first kind (AA1) and of second kind (AA2) under the restriction of using only 1-step self-financing strategies. In this respect, it has been shown the existence of $1$-step AA1 in the large fractional binary market when the transaction costs are such that $\lambda_N=o(1/N^H)$. If, instead, $\lambda_N\sqrt{N}$ converges to infinity, then no 1-step asymptotic arbitrage of any kind appears in the model. Moreover, one can also show that, when the Hurst parameter $H$ is chosen close enough to $1/2$, then even in the frictionless large market there is no 1-step AA2.
In the present work, using more general self-financing trading strategies, we aim to construct, for an appropriate sequence of transaction costs, a strong asymptotic arbitrage, i.e. the possibility of getting arbitrarily rich with probability arbitrarily close to one while taking a vanishing risk. This problem can be viewed as a continuation of the study of asymptotic arbitrage initiated in \cite{CP}, in the sense that our trading strategies are chosen beyond the $1$-step setting of \cite{CP}. Not only that, the existence of this form of asymptotic arbitrage is stronger than AA1 and AA2 and, moreover, is obtained for any Hurst parameter $H>1/2.$
First, in the case of frictionless markets, we construct a candidate sequence of self-financing strategies, and we show that the value process of the portfolio can be expressed as a sum of dependent random variables. Due to this dependency, special versions of the law of large numbers are needed in order to conclude on the asymptotic behaviour of the value process at maturity. In this respect, with the help of a law of large numbers for mixingales (see \cite{An}), we prove that our strategies provide a strictly positive profit with probability strictly close to one. In order to construct a strong asymptotic arbitrage, we modify the sequence of trading strategies, first to ensure that the admissibility condition is satisfied and then to obtain an arbitrarily big profit. Indeed, using a well chosen sequence of stopping times, we stop the self-financing strategies at the first time the admissibility condition fails to hold. The resulting sequence of trading strategies paves the way to a strong asymptotic arbitrage. When transaction costs are taken into account, we show, following a similar argument, the existence of a strong asymptotic arbitrage when the transaction costs are of order $o(\sqrt{\log{N}}/N^{(2H-1/2)\wedge(H+1/2)})$. In direct comparison with the results of \cite{CP}, one can observe that, even if, when using a sequence of $1$-step self financing trading strategies, the rate of convergence of the transaction costs leading to an AA1 is better, this won't allow us though to obtain an AA2.
We emphasize that the methods presented in this work are not restricted to the chosen large financial market. To the contrary, since, in discrete time setting, the value process can be written as a sum of random variables, we believe that these techniques may be applicable also for other examples of discrete large financial markets. This is indeed the case whenever we dispose of an appropriate law of large numbers theorem and of a maximal inequality for the value process, in a similar manner as seen in our results.
The paper is structured as follows. In Section \ref{S2} we introduce the framework of our results, starting with the definition of a fractional binary market. We end this part with a short presentation of the concept of strong asymptotic arbitrage. In Section \ref{S3} and Section \ref{S4} the main results are concentrated. In Section \ref{S3}, we introduce a sequence of self-financing strategies leading to a strictly positive profit with probability close to one. A strong asymptotic arbitrage is then constructed using the aforementioned stopping procedure. In Section \ref{S4}, we extend this construction to the case when transaction costs are considered in the model.
\section{Preliminaries}\label{S2}
\subsection{Fractional binary markets}\label{bmao}
In this section, we briefly recall the so-called fractional binary markets, which were defined by Sottinen in \cite{Sotti} as a sequence of discrete markets approximating the fractional Black-Scholes model.
First, we introduce the fractional Black-Scholes model. This continuous market takes the same form as the classical Black-Scholes model with the difference that the randomness of the risky asset is described by a fractional Brownian motion and not by a standard Brownian one. More precisely, the dynamics of the bond and of the stock are given by:
\begin{equation}\label{fbse}
dB_t=r(t)\,B_t\, dt\quad\textrm{and}\quad dS_t=(a(t)+\sigma\,dZ_t)\, S_t,
\end{equation}
where $\sigma>0$ is a constant representing the volatility and $Z$ is a fractional Brownian motion of Hurst parameter $H>1/2$. The functions $r$ and $a$ are deterministic and represent the interest rate and the drift of the stock. We assume in the sequel that the bond plays the role of the num\'eraire and, hence, $B_t=1$ for all times $t$ (i.e. $r=0$), and that $a$ is continuously differentiable.
Motivated by the construction of an easy example of arbitrage related to the fractional Black-Scholes model, Sottinen came up with the idea to express this special type of Black-Scholes model as a limiting process of a sequence of discrete markets with a binary structure. For this scope, he shows a Donsker-type theorem, in which the fractional Brownian motion is approximated by an inhomogeneous random walk. From this point on, he constructs a discrete model, called ``fractional binary market'', approximating \eqref{fbse}. Based on the results in \cite{CKP}, we provide here a simplified, but equivalent, presentation of these binary models.
Let $(\Omega,\Fs, P)$ be a finite probability space and consider a sequence of i.i.d. random variables $(\xi_i)_{i\geq 1}$ such that
$P(\xi_1=-1)=P(\xi_1=1)=1/2.$
We denote $(\Fs_{i})_{i\geq 0}$ the induced filtration, i.e. $\Fs_{i}:=\sigma(\xi_1,\dots,\xi_i)$, for $i\geq1$, and $\Fs_0:=\{\emptyset,\Omega\}$.
For each $N>1$, the $N$-fractional binary market is the discrete market in which the bond and stock are traded at the
times $\{0, \frac{1}{N},...,\frac{N-1}{N}, 1\}$ under the dynamics:
\begin{equation}\label{fbm}
B_n^{N}=1\quad\textrm{and}\quad S_n^{N}=\left(1+a_n^{N}+\frac{X_n}{N^H}\right)\, S_{n-1}^{N}.
\end{equation}
We assume that the value of $S^{N}$ at time $0$ is constant, i.e.~$S_0^{N}=s_0$. The drift $a^{N}$ approximates the continuous drift given in \eqref{fbse} via $a_n^{N}=\frac{1}{N}a(n/N)$ and the process $(X_n)_{n\geq 1}$ can be expressed as
\begin{equation}\label{scale}
X_n:=\sum\limits_{i=1}^{n-1}j_n(i)\,\xi_i+g_n\xi_n,
\end{equation}
where
$$j_n(i):=\sigma\, C_H \int\limits_{i-1}^{i}x^{\frac{1}{2}-H}\left(\int\limits_0^1 (v+n-1)^{H-\frac{1}{2}} (v+n-1-x)^{H-\frac{3}{2}}dv\right) dx,$$
and
$$g_n:=\sigma\, C_H\int\limits_{n-1}^{n}x^{\frac{1}{2}-H}(n-x)^{H-\frac{1}{2}}\left(\int\limits_0^1 (y(n-x)+x)^{H-\frac{1}{2}}y^{H-\frac{3}{2}}dy\right)dx.$$
From \eqref{scale}, we see that $X_n$ is the sum of a process depending only on the information until time $n-1$ and a process depending only on the present. More precisely, $X_n=\Ys_n+g_n\xi_n$, where
$$\Ys_n:=\sum_{i=1}^{n-1}j_n(i)\xi_i.$$
Therefore, given the history up to time $n-1$, which fixes the values of $\Ys_n$ and $S_{n-1}^{N}$, the price process can take only two possible values at the next step:
$$\left(1+a_n^{N}+\frac{\Ys_{n}-g_n}{N^H}\right)S_{n-1}^{N}\quad\textrm{or}\quad\left(1+a_n^{N}+\frac{\Ys_{n}+g_n}{N^H}\right)S_{n-1}^{N}.$$
This brings to light the binary structure of these markets.
\subsubsection{Some useful estimations}\label{est}
We recall some estimations obtained in or easily derived from \cite{CKP} for the quantities involved in the definition of the fractional binary markets, i.e., $a_n^{N}$, $j_n$ and $g_n$.
\begin{lemma}\label{eji}
There exist constants $C_1, C_2>0$ such that for all $i\geq 2$ we have
\begin{itemize}
\item For $1\leq\ell\leq i/4$:
$$j_i(\ell)\leq \frac{C_1}{i^{2-2H}\,\ell^{H-\frac{1}{2}}}.$$
\item For all $1\leq k\leq 3i/4$:
$$j_i(i-k)\leq \frac{C_2}{k^{\frac{3}{2}-H}}.$$
\end{itemize}
\end{lemma}
\begin{proof}
The proof of the first inequality follows using similar arguments to those used in the proof of \cite[Proposition 5.5]{CKP}. The second inequality uses analogous estimations to those obtained in the proof of \cite[Lemma 5.7]{CKP}.
\end{proof}
Next result corresponds to \cite[Lemma 3.5 and Theorem 5.10]{CKP}.
\begin{lemma}\label{eg1}\ For all $1< n\leq N$, we have
$$g\,\leq g_n\leq g\,\left(1+\frac{1}{n-1}\right)^{H-\frac{1}{2}}\leq g\,2^{H-\frac{1}{2}},$$
where $g:=\frac{\sigma c_H}{H+\frac12}$.
This implies that $\lim_{n\rightarrow \infty}g_n=g$. Moreover,
\begin{equation}\label{cys}
\Ys_n\xrightarrow[n\rightarrow\infty]{(d)}\Ys:= 2g\sum_{k=1}^{\infty}\rho(k)\xi_k,
\end{equation}
where $\rho$ denotes the autocovariance function of a fractional Brownian motion of Hurst parameter $h=\frac{H}2+\frac14\in(\frac12,\frac34)$, i.e.
$\rho(n):=\frac{1}{2}[(n+1)^{2h}+(n-1)^{2h}- 2n^{2h}]>0.$
\end{lemma}
For the drift term $a_n^{N}$, one can see using the definition and the continuity of the function $a$ that:
\begin{equation}\label{ea}
|a_n^{(N)}|\leq \frac{{||a||}_\infty}{N},\qquad n\in\{1,...,N\}.
\end{equation}
This indicates that the contribution of $a_n^{N}$ is significantly less than the contribution of the other parameters of the model. Since we are interested in asymptotic properties of the fractional binary markets, the problem can be simplified by studying the case without the drift. Therefore, we assume henceforth that $a_n^{N}=0$ for all $1\leq n\leq N$.
\subsection{\texorpdfstring{Strong asymptotic arbitrage under transaction costs}{}}
It is well known that the fractional Black-Scholes model without friction is not free of arbitrage. This fact is also reflected in the approximating sequence of fractional binary markets, since, as shown by Sottinen in \cite{Sotti}, the $N$-fractional binary market admits, for $N$ sufficiently large, arbitrage opportunities. However, a pathological situation occurs when one introduces transaction costs. On the one hand, the fractional Black-Scholes model is free of arbitrage under arbitrarily small transaction costs. On the other hand, one can choose transaction costs $\lambda_N$ converging to $1$ such that the $N$-fractional binary market, for $N$ large enough, admits arbitrage under transaction costs $\lambda_N$ (see \cite{CP}). Despite this, the corresponding arbitrage opportunities disappear in the limit, in the sense that, the explicit strategies behind this counterintuitive behaviour provide strictly positive profits with probabilities vanishing in the limit. In order to avoid this kind of situations, we look here to the whole sequence of fractional markets as a large financial market, the large fractional binary market, and we study its asymptotic arbitrage opportunities, as introduced by Kabanov and Kramkov in \cite{Kakra94} and \cite{Kab:Kra:1998}.
\begin{definition}[Large fractional binary market] The sequence of markets given by
$\{(\Omega,\Fs,(\Fs_n)_{n=0}^N, P,S^{N})\}_{N\geq 1}$, where $S^{N}$ is the price process defined in \eqref{fbm}, is called large fractional binary market.
\end{definition}
We assume that the $N$-fractional binary market is subject to $\lambda_N\geq 0$ transaction costs ($\lambda_N=0$ corresponds to the frictionless case). We assume, without loss of generality, that we pay $\lambda_N$ transaction costs only when we sell and not when we buy. This means that the bid and ask price of the stock $S^{N}$ are modelled by the processes ${((1-\lambda)S^{N}_n)}_{n=0}^N$ and ${(S^{N}_n)}_{n=0}^N$ respectively.
\begin{definition}[$\lambda_N$-self-financing strategy]\label{sfs}
Given $\lambda_N\in[0,1]$, a $\lambda_N$-self-financing strategy for the process $S^{N}$ is an adapted process $\phi^N={(\phi_n^{0,N},\phi_n^{1,N})}_{n=-1}^N$ satisfying, for all $n\in\{0,...,N\}$, the following condition:
\begin{equation}\label{selffin}
\phi_n^{0,N}-\phi_{n-1}^{0,N}\leq -{(\phi_n^{1,N}-\phi_{n-1}^{1,N} )}^+\,S_n^{N}\, +\, (1-\lambda_N)\,{(\phi_n^{1,N}-\phi_{n-1}^{1,N} )}^-\,S_n^{N}.
\end{equation}
Here $\phi^{0,N}$ denotes the number of units we hold in the bond and $\phi^{1,N}$ denotes the number of units in the stock. For such a $\lambda_N$-self-financing strategy, the liquidated value of the portfolio at each time $n$ is given by
$$V_n^{\la_N}(\phi^N):=\phi_n^{0,N}+(1-\la_N)(\phi_n^{1,N})^+S_n^{N}-(\phi_n^{1,N})^-S^{N}_n.$$
If $\la_N=0$, we simply write $V_n(\phi^N)$ instead of $V_n^0(\phi^N)$.
\end{definition}
\begin{remark}\label{csfs}
For the purposes of this work, we can restrict our attention to self-financing strategies satisfying \eqref{selffin} with equality and having also that $\phi_N^{1,N}=0$. In other words, we avoid throwing away money and, at maturity, we liquidate the position in stock. For these kind of self-financing strategies, the values of $\phi_n^{0,N}$, $n\in\{0,...,N\}$, can be expressed in terms of the values of $\lambda_N$, $\phi_{-1}^{0,N}$ and $(\phi_k^{1,N})_{k=-1}^n$ as follows:
\begin{equation}\label{selffineq}
\phi_n^{0,N}= \phi_{-1}^{0,N}-\sum_{k=0}^n{(\De_k\phi^{1,N})}^+\,S^{N}_k\, +\, (1-\lambda_N)\,\sum_{k=0}^n{(\De_k\phi^{1,N})}^-\,S^{N}_k.
\end{equation}
In the previous identity, we use the notation $\De_n h:=h_n-h_{n-1}$.
Equation \eqref{selffineq} gives us a way to construct self-financing strategies. More precisely, given $\lambda_N\geq 0$, a constant $\phi_{-1}^{0,N}$ and an adapted process $(\phi_k^{1,N})_{k=-1}^N$, we can use \eqref{selffineq} to define $(\phi_k^{0,N})_{k=0}^N$. The resulting adapted process ${(\phi_n^{0,N},\phi_n^{1,N})}_{n=-1}^N$ is by construction a $\lambda_N$-self-financing strategy, satisfying \eqref{selffin} with equality.
\end{remark}
In their work, Kabanov and Kramkov \cite{Kab:Kra:1998} distinguished between two kinds of asymptotic arbitrage, of the first kind and of the second kind. An asymptotic arbitrage of the first kind gives the possibility of getting arbitrarily rich with strictly positive probability by taking an arbitrarily small risk, whereas the second one is an opportunity of getting a strictly positive profit with probability arbitrarily close to $1$ by taking the risk of losing a uniformly bounded amount of money. The authors also considered a stronger version called ``strong asymptotic arbitrage'', which inherits the strong properties of the two mentioned kinds. More precisely, it can be seen as the possibility of getting arbitrarily rich with probability arbitrarily close to $1$ while taking a vanishing risk. We will work from now on with the latter concept.
We introduce now the definition of strong asymptotic arbitrage. For a detailed presentation on this topic, we refer the reader to \cite{Kab:Kra:1998} for frictionless markets and to \cite{Kl:Le:Pe} for markets with transaction costs.
\begin{definition}\label{SAAd}
There exists a strong asymptotic arbitrage (SAA) with transaction costs $\{\la_N\}_{N\geq 1}$ if there exists a subsequence of markets (again denoted by $N$) and self-financing trading strategies $\phi^N=(\phi^{0,N}, \phi^{1,N})$ with zero endowment for $S^{N}$ such that
\begin{enumerate}
\item{($c_N$-admissibility condition)} For all $i=0,\ldots,N$, $$V^{\la_N}_i(\phi^N)\geq-c_N,$$
\item $\lim_{N\to\infty}P^N(V_{N}^{\la_N}(\phi^N)\geq C_N)=1$
\end{enumerate}
where $c_N$ and $C_N$ are sequences of positive real numbers with $c_N\to0$ and $C_N\to\infty$.
\end{definition}
\begin{remark}
For self-financing strategies with zero endowment, and satisfying \eqref{selffin} with equality, the value process takes the following form:
\begin{align}\label{value}
V_n^{\lambda_N}(\phi^N)&=V_0^{\lambda_N}(\phi^N)+\sum\limits_{k=1}^{n}\phi_{k-1}^{1,N}\Delta_k S^{N}-\lambda_N\sum\limits_{k=1}^{n}\ind_{\{\Delta_k\phi^{1,N}\geq 0\}}\Delta_k\left[{(\phi^{1,N})}^+ S^{N}\right]\nonumber\\
&\qquad-\lambda_N\sum\limits_{k=1}^{n}\ind_{\{\Delta_k\phi^{1,N}< 0\}}\left\{\phi_{k-1}^{1,N}\Delta_k S^{N}+\Delta_k\left[{(\phi^{1,N})}^- S^{N}\right]\right\},
\end{align}
where
\begin{equation}\label{V0}
V^{\lambda_N}_0(\phi^N)=-\lambda_N|\phi^{1,N}_0|s_0.
\end{equation}
\end{remark}
\section{Strong asymptotic arbitrage without transaction costs}\label{S3}
As pointed out in the introduction, the large fractional binary market does not fulfil the standard conditions used in the theory of asymptotic arbitrage for large financial markets. For this reason, we use here a constructive approach to study the existence of strong asymptotic arbitrage.
In this section, we consider the frictionless case. Our first goal is to construct self-financing strategies providing strictly positive profits at maturity with probability converging to one. To do so, we choose, for each $N\geq 1$, a trading strategy $\phi^N:=(\phi^{0,N},\phi^{1,N})$ similar to the one provided in \cite{BeSoVa} for the continuous framework. We have seen in Remark \ref{csfs} that, it is enough to indicate the position in stock $\phi^{1,N}$, as the position in bond $\phi^{0,N}$ can be derived from \eqref{selffineq}, setting $\lambda_N=0$ and $\phi_{-1}^{0,N}:=0$. Moreover, $\phi^{1,N}$ is chosen to be given by
$$\phi_{-1}^{1,N}:=\phi_0^{1,N}:=0\ \ \text{and}\ \ \phi_k^{1,N}:=N^{H-1}\frac{X_k}{S_k^{N}},\quad k\in\{1,\ldots,N\}.$$
Using \eqref{value} and \eqref{V0} with $\lambda_N=0$, we deduce that the value process associated to $\phi^N$ is given by
$$V_n(\phi^N)=\frac1{N}\sum_{k=1}^nX_{k-1}X_k,\quad n\in\{0,...,N\}.$$
Note that the terms in the sum can be expressed as
\begin{equation}\label{theta}
X_{k-1}X_k=\theta_k^{(1)}+\theta_k^{(2)}+\theta_k^{(3)}+\theta_k^{(4)},
\end{equation}
where
\begin{equation*}
\theta_k^{(1)}:=g_{k-1}g_k\xi_{k-1}\xi_k,\quad
\theta_k^{(2)}:=g_{k}\xi_k\Ys_{k-1},\quad
\theta_k^{(3)}:=g_{k-1}\xi_{k-1}\Ys_k,\quad
\theta_k^{(4)}:=\Ys_{k-1}\Ys_k.
\end{equation*}
Defining, for $i=1,2,3,4$,
$$\Ss_n^{(i)}=\sum_{k=1}^n\theta_{k}^{(i)},\quad n=1,\ldots,N,$$
we see that
\begin{equation}\label{VN}
V_n(\phi^{N})=\frac{1}{N}\left(\Ss_n^{(1)}+\Ss_n^{(2)}+\Ss_n^{(3)}+\Ss_n^{(4)}\right).
\end{equation}
\subsection{The value process at maturity and the law of large numbers}\label{s31}
We aim to characterize the asymptotic behaviour of $V_N(\phi^{N})$ by studying the asymptotic properties of each term entering \eqref{VN}. We will see that the first term in \eqref{VN} is a sum of pairwise independent random variables and hence, an appropriate extension of the law of large numbers to this situation can be applied (see \cite{CTT}). The asymptotic behaviour of the second and third terms in \eqref{VN} will be deduced by studying their variances. Finally, for the last term, we use a different approach, since the random variables $(\theta_k^{(4)})_{k\geq 1}$ are pairwise correlated. The notion of mixingale and a law of large numbers for uniformly integrable $L^1$-mixingales will play a key role in studying the behaviour of this term. For the ease of the reader, we recall now the notion of mixingale.
\begin{definition}[$L^p$-Mixingale]\label{mix}
A sequence $\{X_k\}_{k\geq1}$ of random variables is an $L^p$-mixingale with respect to a given filtration $(\Fs_k)_{k\in\Zb}$, if there exist nonnegative constants $\{c_k\}_{k\geq1}$ and $\{\psi_m\}_{m\geq0}$ such that $\psi_m\to0$ as $m\to\infty$ and for all $k\geq1$ and $m\geq0$ the following hold:
\begin{itemize}
\item [(a)] ${\parallel E(X_k|\cF_{k-m}))\parallel}_p\leq c_k\psi_m$
\item [(b)] ${\parallel X_k-E(X_k|\cF_{k+m}))\parallel}_p\leq c_k\psi_{m+1}$.
\end{itemize}
\end{definition}
In order to study the last term in \eqref{VN}, we define:
$$\Ys_k^*:=\Ys_{k-1}\Ys_k-E[\Ys_{k-1}\Ys_k],$$ and we consider the filtration $\mathbb{F}^*:=(\Fs_{i}^*)_{i\in\Zb}$ given by $\Fs_i^*:=\Fs_{i-1}$, for $i\geq 2$, and $\Fs_i^*:=\{\emptyset,\Omega\}$, for $i\leq 1$.
\begin{proposition}\label{mixi}
The process $(\Ys_k^*:k\geq 1)$ is an $L^2$-bounded $L^2$-mixingale with respect to $\mathbb{F}^*$.
\end{proposition}
\begin{proof}
We first prove that the process $(\Ys_k^*:k\geq 1)$ is $L^2$-bounded. Note that
$$E\left[|\Ys_{k-1}\Ys_k|^2\right]\leq E\left[|\Ys_{k-1}|^4\right]+E\left[|\Ys_k|^4\right],$$
and using Khintchine's inequality (see \cite[(1)]{Khin}) for both terms on the right side, we obtain that
$$E\left[|\Ys_{k-1}\Ys_k|^2\right]\leq 3\left(E\left[|\Ys_{k-1}|^2\right]^2+E\left[|\Ys_k|^2\right]^2\right).$$
From \cite[Lemma 5.1]{CKP}, we conclude that $E[|\Ys_{k-1}\Ys_k|^2]$ is uniformly bounded, and therefore, $(\Ys_k^*:k\geq 1)$ is $L^2$-bounded.
Now, we proceed to show that $\Ys_k^*$ is an $L^2$-mixingale with respect to $\mathbb{F}^*$, i.e. we have to check the two conditions of Definition~\ref{mix}. Note that, since $\Ys_k^*$ is $\Fs_k^*$-measurable, condition (b) is automatically satisfied. Hence, it remains to prove condition (a) of Definition \ref{mix}, i.e.
\begin{equation}\label{mpa}
{\parallel E[\Ys_k^*|\Fs_{k-m}^*]\parallel}_2\leq c_k\psi_m,\quad k\geq 1, m\geq 0,
\end{equation}
for some nonnegative constants $c_k$ and $\psi_m$ such that $\psi_m\to0$ as $m\to\infty$.
Note that, for $k\leq m+1$, the left-hand side of \eqref{mpa} is equal to zero, and then, \eqref{mpa} holds for any choice of $c_k$ and $\psi_m$. The case $m=0$ can be easily treated using that $(\Ys_k^*:k\geq 1)$ is $L^2$-bounded. Now, we assume that $k-1>m\geq 1$, and we write $\Ys_{k-1}\Ys_k$ as follows:
\begin{align}\label{y}
\Ys_k\Ys_{k-1}&=\left(\sum_{l=1}^{k-1}j_k(l)\xi_l\right)\left(\sum_{l=1}^{k-2}j_{k-1}(l)\xi_l\right)\nonumber\\
&=\left(\underbrace{\sum_{l=1}^{k-m-1}j_k(l)\xi_l}_{P_{k-m}^{(1)}}+\underbrace{\sum_{l=k-m}^{k-1}j_k(l)\xi_l}_{F_{k-m}^{(1)}}\right)\left(\underbrace{\sum_{l=1}^{k-m-1}j_{k-1}(l)\xi_l}_{P_{k-m}^{(2)}}+\underbrace{\sum_{l=k-m}^{k-2}j_{k-1}(l)\xi_l}_{F_{k-m}^{(2)}}\right)\nonumber\\
&=P_{k-m}^{(1)}P_{k-m}^{(2)}+P_{k-m}^{(2)}F_{k-m}^{(1)}+P_{k-m}^{(1)}F_{k-m}^{(2)}+F_{k-m}^{(1)}F_{k-m}^{(2)},
\end{align}
Using that $P_{k-m}^{(i)}$ is independent of $F_{k-m}^{(j)}$, that $P_{k-m}^{(i)}$ is measurable with respect to $\Fs_{k-m}^*$ and that $F_{k-m}^{(i)}$ is independent of $\Fs_{k-m}^*$ for all $i,j\in\{1,2\}$, we deduce from \eqref{y} that
\begin{equation}\label{exp}
E[\Ys_k\Ys_{k-1}]=E\left[P_{k-m}^{(1)}P_{k-m}^{(2)}\right]+E\left[F_{k-m}^{(1)}F_{k-m}^{(2)}\right]
\end{equation}
and
\begin{equation}\label{condexp}
E\left[\Ys_k\Ys_{k-1}|\Fs_{k-m}^*\right]=P_{k-m}^{(1)}P_{k-m}^{(2)}+E\left[F_{k-m}^{(1)}F_{k-m}^{(2)}\right].
\end{equation}
From \eqref{exp} and \eqref{condexp}, we have that
\begin{align}\label{condexp2}
E\left[\Ys_k^*|\Fs_{k-m}^*\right]=P_{k-m}^{(1)}P_{k-m}^{(2)}-E\left[P_{k-m}^{(1)}P_{k-m}^{(2)}\right].
\end{align}
Now, we write
$$P_{k-m}^{(1)}P_{k-m}^{(2)}=\sum_{l\neq p}^{k-m-1}j_k(l)j_{k-1}(p)\xi_l\xi_p+\sum_{l=1}^{k-m-1}j_k(l)j_{k-1}(l),$$
which implies, using the independence of $\xi_l$ and $\xi_p$ for $l\neq p$, that
$$E\left[P_{k-m}^{(1)}P_{k-m}^{(2)}\right]=\sum_{l=1}^{k-m-1}j_k(l)j_{k-1}(l),$$
and hence
$$E\left[\Ys_k^*|\Fs_{k-m}^*\right]=P_{k-m}^{(1)}P_{k-m}^{(2)}-E\left[P_{k-m}^{(1)}P_{k-m}^{(2)}\right]=\sum_{l\neq p}^{k-m-1}j_k(l)j_{k-1}(p)\xi_l\xi_p=:P_{k-m}^*.$$
Note first that
\begin{equation}\label{pvar}
E\left[|P_{k-m}^*|^2\right]\leq 2\sum_{l\neq p}^{k-m-1}(j_k(l)\,j_{k-1}(p))^2\leq 2\sum_{l=1}^{k-m-1}(j_k(l))^2\sum_{l=1}^{k-m-1}(j_{k-1}(l))^2.
\end{equation}
Additionally, using Lemma~\ref{eji}, we see that
\begin{align*}
\sum_{l=1}^{k-m-1}(j_k(l))^2&\leq\sum_{l=1}^{\frac{k}4}(j_k(l))^2+\sum_{l=m+1}^{\frac{3k}4-1}(j_k(k-l))^2\leq\sum_{l=1}^{\frac{k}4}(j_k(l))^2+\sum_{l=1}^{\frac{3k}4}(j_k(k-l))^2\\
&\leq\frac{C_1}{k^{4-4H}}\sum_{l=1}^{\frac{k}4}\frac1{l^{2H-1}}+C_2\sum_{l=1}^{\frac{3k}4}\frac1{l^{3-2H}}\leq \frac{C^0}{k^{2-2H}}\leq \frac{C^0}{m^{2-2H}},
\end{align*}
where $C^0>0$ is a well chosen constant. In the previous estimation, the term $\sum_{l=m+1}^{\frac{3k}4-1}(j_k(k-l))^2$ has to be understood as equal to zero if $k-m-1\leq k/4$.
A similar argument shows that, there is a constant $C^*>0$ such that
\begin{equation*}
\sum_{l=1}^{k-m-1}(j_{k-1}(l))^2\leq \frac{C^*}{{m}^{2-2H}}.
\end{equation*}
Consequently, Equation \eqref{pvar} leads to
\begin{equation*}
E[|P_{k-m}^*|^2]\leq \frac{C}{m^{4-4H}},
\end{equation*}
where $C>0$ is an appropriate constant. We therefore obtain that, for an appropriate constant $c>0$, the following holds:
\begin{equation}
\sqrt{ E[|E[\Ys_k^*|\Fs_{k-m}^*]|^2]}=\sqrt{E[|P_{k-m}^*|^2]}\leq c\,\frac1{m^{2-2H}}.
\end{equation}
The result follows by choosing $c_k:=c$ and $\psi_m:=m^{2H-2}$.
\end{proof}
\begin{remark}
We have proved that $(\Ys_k^*,k\geq 1)$ is an $L^2$-bounded $L^2$-mixingale. In particular, $(\Ys_k^*,k\geq 1)$ is an uniformly integrable $L^1$-mixingale (we can use the same $c_\ell$ and $\psi_m$). Since in addition, $\sum_{k=1}^n c_k/n=c<\infty$, the conditions of the law of large numbers for mixingales given in \cite[Theorem 1]{An} are satisfied.
\end{remark}
The next result gives the asymptotic behaviour of $V_N(\phi^{N})$ by studying the convergence properties of each term appearing in \eqref{VN}.
\begin{theorem}[Law of large numbers]\label{conv}
The following statements hold:
\begin{enumerate}
\item $\frac1{N}\,\Ss_N^{(1)}\xrightarrow[N\to\infty]{a.s.}0$,\\
\item $\frac1{N}\,\Ss_N^{(2)}\xrightarrow[N\to\infty]{L^2(P)}0$,\\
\item $\frac1{N}\,\Ss_N^{(3)}\xrightarrow[N\to\infty]{L^2(P)}g^2(2^{H+\frac12}-2)>0$,
\item $\frac1{N}\,\Ss_N^{(4)}\xrightarrow[N\to\infty]{L^1(P)}4g^2\sum\limits_{k=2}^\infty\rho(k)\rho(k-1)>0$.
\end{enumerate}
In particular
$$V_N(\phi^{N})\xrightarrow[N\to\infty]{P}\vartheta>0,$$
where $\vartheta:=4g^2\sum_{k=2}^\infty\rho(k)\rho(k-1)+g^2(2^{H+\frac12}-2)$.
\end{theorem}
\begin{proof}[Proof of Theorem \ref{conv}]
(1) Note first that, for all $j\neq k$, we have
$$P(\xi_k\xi_{k-1}=x|\xi_j\xi_{j-1}=y)=\frac12,$$
independently of $y$, where $x,y\in\{-1,1\}$. Therefore, the random variables $(\theta_k^{(1)})_{k\geq 1}$ are pairwise independent. In addition, since $\Var[\xi_k\xi_{k-1}]=1$, we deduce from the estimates given in Lemma~\ref{eg1} that $\sum_{k=1}^N\frac1{k^2}\Var[\theta_k^{(1)}]<\infty$. Hence, the result follows as an application of the law of large numbers for pairwise independent random variables (see \cite[Theorem 1]{CTT}).
(2) Note that $\xi_k$ is independent of $\Ys_{k-1}$, and in particular $E[\xi_k\Ys_{k-1}]=0$. Consequently, the convergence in $L^2(P)$ of $\Ss^{(2)}_N/N$ to $0$ is equivalent to the convergence of the variance to $0$. In addition, for any $k<j$, one can also see that
$$E[\xi_k\Ys_{k-1}\xi_j\Ys_{j-1}]=E[\xi_k\Ys_{k-1}\Ys_{j-1}]E[\xi_j]=0.$$
It follows that
\begin{align*}
\Var\left[\frac1{N}\Ss^{(2)}_N\right]&=\frac1{N^2}\sum_{k=1}^N{g_k}^2\Var[\xi_k\Ys_{k-1}]=\frac1{N^2}\sum_{k=1}^N{g_k}^2E[\xi_k^2(\Ys_{k-1})^2]\\
&=\frac1{N^2}\sum_{k=1}^N {g_k}^2E[(\Ys_{k-1})^2]=\frac1{N^2}\sum_{k=1}^N {g_k}^2\Var[\Ys_{k-1}].
\end{align*}
By Lemma~\ref{eg1}, we have that $g_k\leq 2^{H-\frac12}g$ and by \cite[Lemma 5.1]{CKP}, we have that $$\sup_{k}\Var[\Ys_{k-1}]<\infty.$$ Consequently, we obtain
\begin{align*}
\Var\left[\frac1{N}\Ss^{(2)}_N\right]&\leq\frac{M}{N}\xrightarrow[N\to\infty]{}0
\end{align*}
for some constant $M>0$. This gives us the convergence of $\Ss_N^{(2)}$ to $0$ in $L^2(P)$.\\
(3) We write $\Ss_N^{(3)}$ as a sum of a random term and a deterministic one. We prove that the variance of the random term converges to $0$ and the deterministic term converges to $g^2(2^{H+\frac12}-2)>0$. Indeed, denoting $\tilde{\Ys}_{k-1}:=\sum_{l=1}^{k-2}j_k(l)\xi_l$, we have that
\begin{align*}
\frac1{N}\Ss_N^{(3)}&=\frac1{N}\sum_{k=1}^Ng_{k-1}\xi_{k-1}\Ys_k=\frac1{N}\sum_{k=1}^Ng_{k-1}\xi_{k-1}\tilde{\Ys}_{k-1}+\frac1{N}\sum_{k=1}^Ng_{k-1}j_k(k-1).
\end{align*}
From \eqref{eg1} and \cite[Eq. 5.3]{CKP}, we see that $g_{k-1}j_k(k-1)\xrightarrow[k\to\infty]{}g^2(2^{H+\frac12}-2)$. As a consequence, we deduce that
$$\frac1{N}\sum_{k=1}^Ng_{k-1}j_k(k-1)\xrightarrow[N\to\infty]{}g^2(2^{H+\frac12}-2).$$
For the random term, since $\tilde{\Ys}_{k-1}$ is independent of $\xi_{k-1}$, and using a similar argument like
in the previous part, one obtains that
$$\frac1{N}\sum_{k=1}^Ng_{k-1}\xi_{k-1}\tilde{\Ys}_{k-1}\xrightarrow[N\to\infty]{L^2(P)}0$$
and hence the desired result.\\
(4) Directly from Proposition~\ref{mixi} and the law of large numbers for uniformly integrable $L^1$-mixingales (see \cite[Theorem 1]{An}). Indeed, by the two mentioned results we have that
\begin{align}\label{a}
\frac1{N}\sum_{k=1}^N\left(\Ys_{k-1}\Ys_k-E[\Ys_{k-1}\Ys_k]\right)\xrightarrow[N\to\infty]{L^1(P)}0.
\end{align}
In addition, for $n\geq 4$, we have
$$E[\Ys_{n-1}\Ys_n]=\sum\limits_{i=1}^{\frac{n}{4}}j_n(i)j_{n-1}(i)+\sum\limits_{i=\frac{n}{4}+1}^{n-2}j_n(i)j_{n-1}(i).$$
Using the estimates given in Lemma \ref{eji}, we deduce that the first sum on the right-hand side converges to zero. For the second sum, following
the lines of the proof of \cite[Lemma 5.7]{CKP}, we obtain that
$$\sum\limits_{i=\frac{n}{4}+1}^{n-2}j_n(i)j_{n-1}(i)=\sum\limits_{k=2}^{\frac{3n}{4}-1}j_n(n-k)j_{n-1}(n-k)\xrightarrow[n\to\infty]{}4g^2\sum\limits_{k=2}^\infty\rho(k)\rho(k-1):=\Vs>0.$$
Consequently, we conclude that
$$E[\Ys_{n-1}\Ys_n]\xrightarrow[n\to\infty]{}\Vs,$$
and therefore
\begin{equation}\label{cyy}
\frac1{N}\sum_{k=1}^NE[\Ys_{k-1}\Ys_k]\xrightarrow[N\to\infty]{}\Vs.
\end{equation}
Thus, we have
\begin{align}\label{b}
E\left[\left|\frac1{N}\sum_{k=1}^N\Ys_{k-1}\Ys_k-\Vs\right|\right] &\leq E\left[\left|\frac1{N}\sum_{k=1}^N\left(\Ys_{k-1}\Ys_k-E[\Ys_{k-1}\Ys_k]\right)\right|\right]\nonumber\\
&\quad +\left|\frac1{N}\sum_{k=1}^N E[\Ys_{k-1}\Ys_k]-\Vs \right|.
\end{align}
By \eqref{a}, \eqref{cyy} and \eqref{b}, it follows immediately that
$$E\left[\left|\frac1{N}\sum_{k=1}^N\Ys_{k-1}\Ys_k-\Vs\right|\right] \xrightarrow[N\to\infty]{}0,$$
and hence
$$\frac1{N}\Ss_N^{(4)}\xrightarrow[N\to\infty]{L^1(P)}\Vs.$$
The proof of (4) is now complete.
\end{proof}
\begin{corollary}\label{AA1conv}
For all $\varepsilon>0$,
$$P(V_N(\phi^{N})>\vartheta(1-\varepsilon))\xrightarrow[N\to\infty]{}1.$$
\end{corollary}
\begin{proof}
The result follows using Theorem~\ref{conv} and the definition of the convergence in probability.
\end{proof}
\subsection{Admissibility condition through stopping procedure}\label{s32}
The sequence of self-financing strategies $(\phi^N)_{N\geq 1}$ constructed in the previous section gives the possibility to make a strictly positive profit with probability arbitrarily close to one. Now, we proceed to modify our strategies in such a way that the admissibility conditions are satisfied. More precisely, we will stop our self-financing strategies at the first time they fail the admissibility condition. To do so, we split the value process as in \eqref{VN}, and we study the stopping times corresponding to each part.
For each $i=1,2,3,4$ and any sequence of strictly positive numbers $(\varepsilon_N,N\geq 1)$, we define the stopping time
$$T_{\varepsilon_N}^{(N,i)}:=\inf\{k\in\{1,\ldots,N\}:\, \frac{1}{N}\,\Ss_k^{(i)}< -\varepsilon_N\},$$
with the convention that $\inf\emptyset =\infty$. Note that these stopping times have values on $\{1,\dots,N\}\cup\{\infty\}$.
We study the first three stopping times with the help of an extension of the Kolmogorov's maximal inequality. A different approach is used in the study of $T_{\varepsilon_N}^{(N,4)}$.
\subsubsection{Maximal inequalities for the first three terms}
We start with the following generalization of the Kolmogorov's maximal inequality, which fits our setting.
Let $c$ be a strictly positive constant and $W=(W_k)_{k=1}^N$ a sequence of centred random variables. We define the stopping time
$$T_c(W):=\inf\{k\in\{1,\ldots,N\}:\, |W_k|> c\}.$$
\begin{lemma}[Maximal inequality]\label{mi}
Assume that, for all $k\in\{1,\ldots,N\}$, we have
\begin{equation}\label{cmi}
E\left[(W_N-W_k)\,W_k\, 1_{\{T_c(W)=k\}}\right]=0.
\end{equation}
Then
$$P\left(\sup\limits_{1\leq i\leq N}|W_i|> c\right)=P(T_c(W)\leq N)\leq \frac{\Var(W_N)}{c^2}.$$
\end{lemma}
\begin{proof}
We may assume that $\Var(W_N)<\infty$. Note that
\begin{align*}
\Var(W_N)&=E[W_N^2]\geq \sum\limits_{k=1}^{N} E[(W_k +W_N-W_k)^2 1_{\{T_c(W)=k\}}]\\
&\geq \sum\limits_{k=1}^{N} \left(E[W_k^2 1_{\{T_c(W)=k\}}]+ 2E\left[(W_N-W_k)W_k 1_{\{T_c(W)=k\}}\right]\right)\\
&=\sum\limits_{k=1}^{N} E[W_k^2 1_{\{T_c(W)=k\}}]\geq c^2P(T_c(W)\leq N).
\end{align*}
The result follows.
\end{proof}
The next result will be obtained as an application of the previous lemma.
\begin{lemma}\label{mi123}
For each $i=1,2,3$, there is a constant $C^{(i)}>0$, such that
$$P\left(T_{\varepsilon_N}^{(N,i)}\leq N\right)\leq \frac{C^{(i)}}{N\,\varepsilon_N^2}.$$
\end{lemma}
\begin{proof}
(1) Define the stopping time $$\widetilde{T}_{\varepsilon_N}^{(N,1)}:=T_{\varepsilon_N}\left(\frac{1}{N}\Ss^{(1)}\right),$$
and note that $T_{\varepsilon_N}^{(N,1)}\geq\widetilde{T}_{\varepsilon_N}^{(N,1)}$. Therefore, it is enough to prove the result for $\widetilde{T}_{\varepsilon_N}^{(N,1)}$.
Since $g_k\leq 2g$ and the random variables $\{\xi_{k-1} \xi_k; k\geq 1\}$ are pairwise independent, we conclude that
$$\Var\left(\frac{1}{N}\Ss_N^{(1)}\right)=\frac{1}{N^2}\sum\limits_{\ell=1}^{N}g_{\ell-1}^2 g_\ell^2\leq \frac{(2g)^4}{N}.$$
Note that $\Ss_k^{(1)}1_{\{\widetilde{T}_{\varepsilon_N}^{(N,1)}=k\}}$ is $\sigma(\xi_1,...,\xi_k)$-measurable. In addition, we have that
$$\Ss_N^{(1)}-\Ss_k^{(1)}=\sum\limits_{\ell=k+2}^{N}g_{\ell-1} g_\ell \xi_{\ell-1} \xi_\ell+g_{k} g_{k+1} \xi_{k} \xi_{k+1}.$$
Moreover, $\sum_{\ell=k+2}^{N}g_{\ell-1} g_\ell\, \xi_{\ell-1} \xi_\ell$ is $\sigma(\xi_{k+1},...,\xi_N)$-measurable and
$$E\left[\xi_{k} \xi_{k+1}\,\Ss_k^{(1)}1_{\{\widetilde{T}_{\varepsilon_N}^{(N,1)}=k\}}\right]=E\left[\xi_{k+1}\right]E\left[ \xi_k\,\Ss_k^{(1)}1_{\{\widetilde{T}_{\varepsilon_N}^{(N,1)}=k\}}\right]=0.$$
Consequently, the condition \eqref{cmi} is satisfied and the result follows as an application of Lemma \ref{mi}.
(2) Define the stopping time $$\widetilde{T}_{\varepsilon_N}^{(N,2)}:=T_{\varepsilon_N}\left(\frac{1}{N}\Ss^{(2)}\right),$$
and note that $T_{\varepsilon_N}^{(N,2)}\geq\widetilde{T}_{\varepsilon_N}^{(N,2)}$. As before, we conclude that it is enough to prove the result for $\widetilde{T}_{\varepsilon_N}^{(N,2)}$.
Since, for $k>j$, $E[\xi_k\Ys_{k-1}\xi_j\Ys_{j-1}]=E[\xi_k]\,E[\Ys_{k-1}\xi_j\Ys_{j-1}]=0$, we have that
$$\Var\left(\frac{1}{N}\Ss_N^{(2)}\right)=\frac{1}{N^2}\sum\limits_{\ell=1}^{N}\sum\limits_{i=1}^{\ell-2} g_\ell^2 j_{\ell-1}^2(i)\leq \frac{(2g)^2}{N^2}\sum\limits_{\ell=1}^{N}\sum\limits_{i=1}^{\ell-2} j_{\ell-1}^2(i).$$
From \cite[Lemma 5.1]{CKP}, the quantities $\sum_{i=1}^{\ell-2} j_{\ell-1}^2(i)$ are uniformly bounded. We conclude that there is $C^{(2)}>0$ such that
$$\Var\left(\frac{1}{N}\Ss_N^{(2)}\right)\leq\frac{C^{(2)}}{N}.$$
Additionally, we have that
$$\Ss_N^{(2)}-\Ss_k^{(2)}=\sum\limits_{\ell=k+1}^{N} g_\ell\xi_\ell\Ys_{\ell-1}.$$
On the other hand, for all $\ell\in\{k+1,...,N\}$, we have that
$$E\left[\xi_{\ell} \Ys_{\ell-1}\,\Ss_k^{(2)}1_{\{\widetilde{T}_{\varepsilon_N}^{(N,2)}=k\}}\right]=E\left[\xi_{\ell}\right]E\left[ \Ys_{\ell-1}\,\Ss_k^{(2)}1_{\{\widetilde{T}_{\varepsilon_N}^{(N,2)}=k\}}\right]=0.$$
The condition \eqref{cmi} is verified and the result follows.
(3) Denote $\tilde{\Ys}_{k-1}:=\sum_{l=1}^{k-2}j_k(l)\xi_l$ and note that:
$$\theta_k^{(3)}=g_{k-1} j_k(k-1) + \underbrace{g_{k-1}\xi_{k-1}\tilde{\Ys}_{k-1}}_{=:\tilde{\theta}_k^{(3)}}>\tilde{\theta}_k^{(3)}.$$
As a consequence, for each $n\in\{1,\ldots ,N\}$, we have
$$\Ss_n^{(3)}>\sum\limits_{k=1}^n \tilde{\theta}_k^{(3)}=:\tilde{\Ss}_n^{(3)}.$$
Moreover, if we define
$$\widetilde{T}_{\varepsilon_N}^{(N,3)}:=T_{\varepsilon_N}\left(\frac{1}{N}\tilde{\Ss}^{(3)}\right),$$
it follows that $T_{\varepsilon_N}^{(N,3)}\geq \widetilde{T}_{\varepsilon_N}^{(N,3)}.$ Thus, it will be enough to prove the result for $\widetilde{T}_{\varepsilon_N}^{(N,3)}$.
Since, for $k>j$, $E[\xi_{k-1}\tilde{\Ys}_{k-1}\xi_{j-1}\tilde{\Ys}_{j-1}]=E[\xi_{k-1}]\,E[\tilde{\Ys}_{k-1}\xi_{j-1}\tilde{\Ys}_{j-1}]=0$, we get
$$\Var\left(\frac{1}{N}\tilde{\Ss}_N^{(3)}\right)=\frac{1}{N^2}\sum\limits_{\ell=1}^{N}\sum\limits_{i=1}^{\ell-2} g_{\ell-1}^2 j_{\ell}^2(i)\leq \frac{(2g)^2}{N^2}\sum\limits_{\ell=1}^{N}\sum\limits_{i=1}^{\ell-2} j_{\ell}^2(i).$$
We know from \cite[Lemma 5.1]{CKP} that the quantities $\sum_{i=1}^{\ell-2} j_{\ell}^2(i)$ are uniformly bounded. We conclude that there is $C^{(3)}>0$ such that
$$\Var\left(\frac{1}{N}\tilde{\Ss}_N^{(3)}\right)\leq\frac{C^{(3)}}{N}.$$
Additionally, we have
$$\tilde{\Ss}_N^{(3)}-\tilde{\Ss}_k^{(3)}=\sum\limits_{\ell=k+1}^{N} g_{\ell-1}\xi_{\ell-1}\tilde{\Ys}_{\ell-1}.$$
On the other hand, for all $\ell\in\{k+1,...,N\}$, we have that
$$E\left[\xi_{\ell-1} \tilde{\Ys}_{\ell-1}\,\tilde{\Ss}_k^{(3)}1_{\{\widetilde{T}_{\varepsilon_N}^{(N,3)}=k\}}\right]=E\left[\xi_{\ell-1}\right]E\left[ \tilde{\Ys}_{\ell-1}\,\tilde{\Ss}_k^{(3)}1_{\{\widetilde{T}_{\varepsilon_N}^{(N,3)}=k\}}\right]=0.$$
The condition in Lemma \ref{mi} is verified and the result follows.
\end{proof}
\subsubsection{Maximal inequality for the last term}
In the previous three cases, we consider the process $W^{(i)}=\Ss^{(i)}-E[\Ss^{(i)}]$, and we use either a pairwise independence argument or the orthogonality of some random variables to prove that condition \eqref{cmi} is satisfied. The desired results are obtained with the help of Lemma \ref{mi}. For the stopping time $T_{\varepsilon_N}^{(N,4)}$, we can not proceed in the same way, because the random variables $(\Ys_k^*)_{k\geq 1}$ are pairwise correlated. Nevertheless, the key ingredient is again a maximal inequality, this time for the process $(\Ys_k^*)_{k\geq 1}$.
Let's define the random variables
$$X_{i,k}:=E[\Ys_i^*|\Fs_{i-k}^*]-E[\Ys_i^*|\Fs_{i-k-1}^*],\quad i\in\Nb,\ k\in \Zb.$$
Note that $X_{i,k}=0$ if $k<0$ or $i\leq k+1$. As a consequence, we have that:
$$Y_{n,k}:=\sum\limits_{i=1}^n X_{i,k}=\sum\limits_{i=k+2}^n X_{i,k}.$$
We also denote
$$\Ss_n^*:=\sum\limits_{k=1}^n \Ys_k^*= \Ss_n^{(4)}-E[\Ss_n^{(4)}].$$
The following result provides the desired maximal inequality for $(\Ss_n^*)_{n\geq 1}$.
\begin{lemma}\label{lmilm}
For all $n\geq 1$, we have that
$$\Ss_n^*=\sum\limits_{k=-\infty}^\infty Y_{n,k}=\sum\limits_{k=0}^{n-2} Y_{n,k}\quad \textrm{a.s.},$$
and for any sequence of strictly positive numbers $(a_k:k\in\Zb)$, we have
\begin{equation}\label{milm}
E\left[\sup_{n\leq N}|\Ss_n^*|^2\right]\leq 4\left(\sum\limits_{k=0}^{N-2} a_k\right)\left(\sum\limits_{k=0}^{N-2} a_k^{-1}\Var(Y_{N,k})\right).
\end{equation}
\end{lemma}
\begin{proof}
From Proposition \ref{mixi}, we deduce that $(\Ys^*_n)_{n\geq 1}$ is a sequence of square integrable random variables with zero mean. It is also straightforward to see that
$$E[\Ys_n^*|\Fs^*_{-\infty}]=\Ys_n^*-E[\Ys_n^*|\Fs^*_{\infty}]=0\quad \textrm{a.s.},$$
where $\Fs^*_{-\infty}:=\{\emptyset,\Omega\}$ and $\Fs^*_{\infty}:=\sigma(\xi_i:i\geq 1)$. Therefore, the first statement follows as a direct application of \cite[Lemma 1.5]{ML} and the fact that $Y_{n,k}=0$ for $k<0$ and $k>n-2$. For the remaining part, we need to slightly modify the arguments of \cite[Lemma 1.5]{ML}.
First note that, for $k\geq 1$, $(Y_{n,k})_{n\geq 1}$ is a square integrable $(\Fs^*_{n-k})_{n\geq 1}$-martingale. On the other hand, using Cauchy-Schwartz
$$(\Ss_n^*)^2=\left(\sum\limits_{k=0}^{n-2} \sqrt{a_k}\,\frac{Y_{n,k}}{\sqrt{a_k}}\right)^2\leq \left(\sum\limits_{k=0}^{n-2}a_k\right)\left(\sum\limits_{k=0}^{n-2} \frac{Y_{n,k}^2}{a_k}\right)\leq \left(\sum\limits_{k=0}^{N-2}a_k\right)\left(\sum\limits_{k=0}^{N-2} \frac{Y_{n,k}^2}{a_k}\right).$$
Now, we take $\sup_{n\leq N}$ in both extremes of this inequality, then we take expectations, and we apply Doob's inequality to bound the right-hand side. The result follows.
\end{proof}
In order to obtain an explicit upper bound for the left-hand side in \eqref{milm}, we start by studying the variance of $Y_{N,k}$.
\begin{lemma}\label{rxni}
For all $0\leq k< i-1$ we have that
$$X_{i,k}=\xi_{i-k-1}\sum\limits_{\ell=1}^{i-k-2}j_i(\ell,i-k-1)\xi_\ell,$$
where $j_i(\ell,p):=j_i(\ell)j_{i-1}(p)+j_i(p)j_{i-1}(\ell).$ In particular, for each $k\geq 0$, the random variables $(X_{i,k}: i> k+1)$ are centred and pairwise uncorrelated. As a consequence, we have that
$$\Var(Y_{N,k})=\sum\limits_{i=k+2}^N\sum\limits_{\ell=1}^{i-k-2}j_i(\ell,i-k-1)^2.$$
\end{lemma}
\begin{proof}
It is straightforward from the expression for $E[\Ys_i^*|\Fs_{i-k}^*]$ obtained in the proof of Proposition \ref{mixi}.
\end{proof}
Next result gives an explicit upper bound for the left-hand side in \eqref{milm}.
\begin{lemma}\label{mub}
There is a constant $C^*>0$ such that
$$E\left[\sup_{n\leq N}|\Ss_n^*|^2\right]\leq C^* \ln(N)N^{4H-2}.$$
\end{lemma}
\begin{proof}
We note first that
$$j_i(\ell,p)^2\leq 2\left( j_i(\ell)^2j_{i-1}(p)^2+j_i(p)^2j_{i-1}(\ell)^2\right).$$
Using Lemma \ref{rxni}, we see that
$$\frac{\Var(Y_{N,k})}{2}\leq \underbrace{\sum\limits_{i=k+2}^N\sum\limits_{\ell=1}^{i-k-2}j_i(i-k-1)^2j_{i-1}(\ell)^2}_{:=V_{N,k}}+\underbrace{\sum\limits_{i=k+2}^N\sum\limits_{\ell=1}^{i-k-2}j_i(\ell)^2j_{i-1}(i-k-1)^2}_{:=W_{N,k}}.$$
Now, we write $V_{N,k}=V_{N,k}^1+V_{N,k}^2+V_{N,k}^3$, where
\begin{align*}
V_{N,k}^1&:=\sum\limits_{i=k+2}^{\frac{4(k+1)}{3}\wedge N}j_i(i-k-1)^2\sum\limits_{\ell=1}^{i-k-2}j_{i-1}(\ell)^2,\\
V_{N,k}^2&:=\sum\limits_{i=\frac{4(k+1)}{3}\wedge N+1}^{N}j_i(i-k-1)^2\sum\limits_{\ell=1}^{\frac{i-1}{4}}j_{i-1}(\ell)^2,\\
V_{N,k}^3&:=\sum\limits_{i=\frac{4(k+1)}{3}\wedge N+1}^{N}j_i(i-k-1)^2\sum\limits_{\ell=\frac{i+3}{4}}^{i-k-2}j_{i-1}(\ell)^2.
\end{align*}
For $V_{N,k}^1$, we use Lemma \ref{eji} to obtain, that
\begin{align*}
V_{N,k}^1&\leq C_1^4\sum\limits_{i=k+2}^{\frac{4(k+1)}{3}}\frac{1}{(i-1)^{8-8H}(i-k-1)^{2H-1}}\sum\limits_{\ell=1}^{i-k-2}\frac{1}{\ell^{2H-1}}\\
&\leq \frac{{C_1^4}}{(2-2H)}\sum\limits_{i=k+2}^{\frac{4(k+1)}{3}}\frac{(i-k-2)^{2-2H}}{(i-1)^{8-8H}(i-k-1)^{2H-1}}\\
&\leq \frac{{C_1^4}}{(2-2H)(k+1)^{6-6H}}\sum\limits_{i=k+2}^{\frac{4(k+1)}{3}}\frac{1}{(i-k-1)^{2H-1}}\leq \frac{\widehat{C}_1}{(k+1)^{4-4H}},
\end{align*}
where $\widehat{C}_1>0$ is an appropriate constant. For the other terms, we assume that $\frac{4(k+1)}{3}\leq N$, otherwise they are trivially equal to zero. Thus, for $V_{N,k}^2$, we have
\begin{align*}
V_{N,k}^2&\leq \frac{(C_1\, C_2)^2}{(k+1)^{3-2H}}\sum\limits_{i=\frac{4(k+1)}{3}+1}^{N}\frac{1}{(i-1)^{4-4H}}\sum\limits_{\ell=1}^{\frac{i-1}{4}}\frac{1}{\ell^{2H-1}}\\
&\leq \frac{(C_1\, C_2)^2}{(2-2H)(k+1)^{3-2H}}\sum\limits_{i=2}^{N}\frac{1}{(i-1)^{2-2H}}\leq \frac{\widehat{C}_2 N^{2H-1}}{(k+1)^{3-2H}},\\
\end{align*}
where $\widehat{C}_2>0$ is a well chosen constant. Similarly, for the last term we have
\begin{align*}
V_{N,k}^3&\leq \frac{C_2^2}{(k+1)^{3-2H}}\sum\limits_{i=\frac{4(k+1)}{3}+1}^{N}\sum\limits_{\ell=k+1}^{\frac{3(i-1)}{4}}j_{i-1}(i-1-\ell)^2\leq \frac{\widehat{C}_3 N}{(k+1)^{5-4H}},
\end{align*}
where $\widehat{C}_3>0$ is a well chosen constant. We conclude that, there exists $C_0>0$ such that
$$V_{N,k}\leq\frac{C_0 N}{(k+1)^{5-4H}}.$$
for all $k\leq N-2$. An upper bound of the same order for $W_{N,k}$ can be obtained using similar arguments. Therefore, there is $C^*>0$, such that
$$\Var(Y_{N,k})\leq\frac{C^* N}{(k+1)^{5-4H}}.$$
The result follows by plugging this upper bound in \eqref{milm} with $a_k:=(k+1)^{-1}$.
\end{proof}
As a consequence of the previous results, we obtain the following analogue of Lemma \ref{mi123} for $T_{\varepsilon_N}^{(N,4)}$.
\begin{corollary}\label{T4}
There is a constant $C^{(4)}>0$ such that
$$P\left(T_{\varepsilon_N}^{(N,4)}\leq N\right)\leq \frac{C^{(4)}\ln(N)}{N^{4-4H}\varepsilon_N^2}.$$
\end{corollary}
\begin{proof}
First, note that,
$$E\left[\Ss_n^{(4)}\right]=\sum_{k=3}^{n}\sum_{i=1}^{k-2}j_{k}(i)j_{k-1}(i)\geq 0.$$
Therefore, we have that
$$\Ss_n^{(4)}=\Ss_n^*+E[\Ss_n^{(4)}]\geq \Ss_n^*.$$
Consequently, if we define the stopping time $T_{\varepsilon_N}^{(N,*)}$ as follows
$$T_{\varepsilon_N}^{(N,*)}:=\inf\left\{k\in\{1,\ldots,N\}:\, \frac{1}{N} |\Ss_k^{*}|> \varepsilon_N\right\},$$
then $T_{\varepsilon_N}^{(N,4)}\geq T_{\varepsilon_N}^{(N,*)}$. In particular, we deduce that
$$P\left(T_{\varepsilon_N}^{(N,4)}\leq N\right)\leq P\left(T_{\varepsilon_N}^{(N,*)}\leq N\right)=P\left(\sup_{n\leq N}\frac1{N}|\Ss_n^*|> \varepsilon_N\right).$$
The result follows as an application of the Tchebychev inequality and Lemma \ref{mub}.
\end{proof}
\subsection{A strong asymptotic arbitrage}
In this section, using the results of section \ref{s32}, we modify the sequence $(\phi^N)_{N\geq 1}$ constructed in section \ref{s31}, in order to construct an explicit strong asymptotic arbitrage. A first modification will lead to a sequence of self-financing strategies $(\hat{\phi}^N)_{N\geq 1}$ providing a strictly positive profit with probability arbitrarily close to one and verifying the admissibility conditions. Finally, after a second modification, we will obtain a new sequence of self-financing strategies $(\psi^N)_{N\geq 1}$ leading to the desired strong asymptotic arbitrage.
The sequence $(\hat{\phi}^N)_{N\geq 1}$ is defined as follows. The position in stock is given by
$$\hat{\phi}_k^{1,N}:=1_{\{k<T_N\}}\phi_k^{1,N},\quad k\in\{-1,0,...,N\},$$ where
\begin{equation}\label{stoptime}
T_N:=T_{\varepsilon_N}^{(N,1)}\wedge T_{\varepsilon_N}^{(N,2)}\wedge \left(T_{\varepsilon_N}^{(N,3)}-1\right)\wedge \left(T_{\varepsilon_N}^{(N,4)}-1\right),
\end{equation}
and the position in bond is derived from \eqref{selffineq} setting $\lambda_N=0$ and $\hat{\phi}_{-1}^{0,N}=0$. Note that, since the random variables $\Ss_n^{(3)}$ and $\Ss_n^{(4)}$ are $\Fs_{n-1}$-measurable, $T_{\varepsilon_N}^{(N,3)}-1$ and $T_{\varepsilon_N}^{(N,4)}-1$ are stopping times with respect to $(\Fs_n)_{n=0}^N$. Clearly, $T_{\varepsilon_N}^{(N,1)}$ and $T_{\varepsilon_N}^{(N,2)}$ are also stopping times with respect to $(\Fs_n)_{n=0}^N$, and consequently, $T_N$ as well.
By construction, the corresponding value process is given by
$$V_n(\hat{\phi}^N)=\frac{1}{N}\sum\limits_{i=1}^{4}\Ss_{n\wedge T_N}^{(i)}.$$
In particular, we have
\begin{equation}\label{fac2}
V_n(\hat{\phi}^N)=V_{n\wedge T_N}(\phi^N)\geq -4\,\varepsilon_N + \frac{1}{N}\left(\theta_{T_N}^{(1)}+\theta_{T_N}^{(2)}\right)1_{\{n\geq T_N\}}.
\end{equation}
Next lemma provides a uniform control for the second term on the right-hand side of \eqref{fac2}.
\begin{lemma}\label{th12}
For $i=1,2,3$, there exists a constant $C_{i}^\theta>0$ such that
$$\sup\limits_{1\leq n\leq N}|\theta_{n}^{(i)}|\leq C_{i}^\theta\,N^{H-\frac{1}{2}}.$$
\end{lemma}
\begin{proof}
It follows from the definition of the random variables $\theta_{n}^{(i)}$ and Lemma \ref{eji}.
\end{proof}
Now, motivated by our previous results, we choose
$$\varepsilon_N:=\frac{\ln(N)}{N^{\frac{1}{2}\wedge(2-2H)}}\quad\textrm{ and }\quad\hat{c}_N:=4\varepsilon_N+\frac{C_{1,2}}{N^{\frac{3}{2}-H}},$$
where $C_{1,2}=C_1^\theta+C_2^\theta$.
Finally, for each $N\geq 1$, we define $\psi^N=(\psi^{0,N},\psi^{1,N})$ as follows. The position in stock, $\psi^{1,N}$, is given by:
\begin{equation}\label{str}
\psi_k^{1,N}:=\frac{1}{\sqrt{\hat{c}_N}}\hat{\phi}_k^{1,N},\quad k\in\{-1,0,...,N\},
\end{equation}
and the position in bond, $\psi^{0,N}$, is constructed as before, through the self-financing conditions \eqref{selffineq}, setting $\lambda_N=0$ and $\psi_{-1}^{0,N}=0$.
\begin{theorem}
The sequence of self-financing strategies $(\psi^N)_{ N\geq 1}$ provides a strong asymptotic arbitrage in the large fractional binary market.
\end{theorem}
\begin{proof}
In order to have a strong asymptotic arbitrage, we need to show that the two conditions of Definition~\ref{SAAd} are satisfied.
More precisely, we prove that these two conditions are verified for
$$c_N:=\sqrt{\hat{c}_N}\xrightarrow[N\rightarrow\infty]{} 0\quad\textrm{and}\quad C_N:=\frac{\vartheta}{2\sqrt{\hat{c}_N}}\xrightarrow[N\rightarrow\infty]{} \infty.$$
Note that, from Lemma \ref{th12} and equation \eqref{fac2}, the self-financing strategy $\hat{\phi}^{N}$ is $\hat{c}_N$-admissible. Since, in addition
$$V_k(\psi^{N})=\frac{1}{c_N}V_k(\hat{\phi}^{N}),\quad k\in\{0,...,N\},$$
we deduce that ${\psi}^{N}$ is $c_N$-admissible. Regarding the second condition, we use the convergence behaviour of $V_N(\phi^{N})$ given in Corollary~\ref{AA1conv}. First, note that $$\{T_{\varepsilon_N}^{(N,i)}-1\leq N\}=\{T_{\varepsilon_N}^{(N,i)}\leq N\},$$
and then, from the choice of $\varepsilon_N$ and Lemma \ref{mi123} and Corollary \ref{T4}, we obtain that
\begin{equation}\label{tntz}
P(T_N\leq N)\leq\sum\limits_{i=1}^4 P\left(T_{\varepsilon_N}^{(N,i)}\leq N\right)\xrightarrow[N\rightarrow\infty]{} 0.
\end{equation}
On the other side, over the set $\{N<T_N\}$, we have
$$V_N(\psi^{N})=\frac{1}{c_N}V_N(\phi^{N}).$$
In particular, we get
\begin{align*}
P\left(V_N(\phi^{N})>\vartheta/2\right)&=P\left(\{V_N(\phi^{N})>\vartheta/2\}\cap\{T_N\leq N\}\right)\\
&\qquad+P\left(\{V_N(\phi^{N})>\vartheta/2\}\cap\{T_N> N\}\right)\\
&\leq P\left(T_N\leq N\right)+P\left(V_N(\psi^{N})>C_N\right).
\end{align*}
Letting now $N\to\infty$ and applying the results of Corollary~\ref{AA1conv} and \eqref{tntz}, we get that
$$\lim_{N\to\infty}P\left(V_N(\psi^{N})>C_N\right)=1.$$
The desired result is then proven.
\end{proof}
\section{Strong asymptotic arbitrage with transaction costs}\label{S4}
In this section, we let each $N$-fractional binary market be subject to $\la_N$ transaction costs, and we show that there exists a strong asymptotic arbitrage if the sequence of transaction costs $(\la_N)_{N\geq1}$ converges to zero fast enough. The corresponding sequence of self-financing strategies $(\psi^N(\lambda_N))_{N\geq 1}$ is constructed as follows. The position in stock is given by $\psi^{1,N}$, as in \eqref{str} of the frictionless case. The position in bond, $\psi^{0,N}(\lambda_N)$, is constructed from $\psi^{1,N}$ through the $\lambda_N$-self-financing conditions \eqref{selffineq}, setting $\psi_{-1}^{0,N}(\lambda_N):=0$.
\begin{theorem}
The self-financing strategies $(\psi^N(\lambda_N))_{N\geq 1}$, where
$$\lambda_N=o\left(\frac{\sqrt{\ln{N}}}{N^{(2H-\frac14)\wedge(H+\frac12)}}\right),$$
provide a strong asymptotic arbitrage in the large fractional binary markets with $(\la_N)_{N\geq 1}$ transaction costs.
\end{theorem}
\begin{proof}
In order to show the first condition of Definition~\ref{SAAd}, we have to make sure that the admissibility condition in the presence of transaction costs is fulfilled. Since $\psi_k^{1,N}=\frac{1}{c_N}\hat{\phi}_k^{1,N}$, we have that
\begin{equation}\label{vpr}
V_n^{\la_N}(\psi^{N}(\lambda_N))=\frac{1}{c_N}\,V_n^{\la_N}(\hat{\phi}^{N}(\lambda_N)),
\end{equation}
where $\hat{\phi}^N(\lambda_N)=(\hat{\phi}^{0,N}(\lambda_N),\hat{\phi}^{1,N})$ and $\hat{\phi}^{0,N}(\lambda_N)$ is determined from $\hat{\phi}^{1,N}$ by means of the $\lambda_N$-self-financing conditions \eqref{selffineq}. Additionally, from \eqref{value} we deduce that
\begin{equation}\label{abc}
V_n^{\la_N}(\hat{\phi}^{N}(\lambda_N))=V_0^{\la_N}(\hat{\phi}^{N}(\lambda_N))+V_n(\hat{\phi}^{N})-\lambda_N\left(\Vs_n^1+\Vs_n^2+\Vs_n^3\right),
\end{equation}
where
\begin{align*}
\Vs_n^1&:=\sum\limits_{k=1}^{n}\ind_{\{\Delta_k\hat{\phi}^{1,N}\geq 0\}}\,\Delta_k\left[{(\hat{\phi}^{1,N})}^+ S^{N}\right],\\
\Vs_n^2&:=\sum\limits_{k=1}^{n}\ind_{\{\Delta_k\hat{\phi}^{1,N}< 0\}}\,\Delta_k\left[{(\hat{\phi}^{1,N})}^- S^{N}\right],\\
\Vs_n^3&:=\sum\limits_{k=1}^{n}\ind_{\{\Delta_k\hat{\phi}^{1,N}< 0\}}\,\hat{\phi}_{k-1}^{1,N}\Delta_k S^{N}.
\end{align*}
Using \eqref{V0} and that $\phi_0^{1,N}=0$, we see that $V_0^{\la_N}(\hat{\phi}^{N}(\lambda_N))=0$. The second term in \eqref{abc} is exactly the value process with $0$ transaction costs for the trading strategy $\hat{\phi}^{N}$ and then, from the results of the previous section we have
$$V_n(\hat{\phi}^{N})\geq - \hat{c}_N.$$
For the third term, we proceed as follows. Using that $|\hat{\phi}_k^{1,N}|\leq |\phi_k^{1,N}|$, we obtain that
\begin{align}\label{aa}
|\Vs_n^1|&\leq \sum\limits_{k=1}^{n}|\phi^{1,N}_k|S_k^N+\sum\limits_{k=1}^{n}|\phi^{1,N}_{k-1}|S_{k-1}^N\nonumber\\
&\leq \frac1{N^{1-H}}\left(\sum\limits_{k=1}^{n}|X_k|+\sum\limits_{k=2}^{n}|X_{k-1}|\right).
\end{align}
For the latter sums, we use the estimates in Lemma~\ref{eg1} and Lemma~\ref{eji} to obtain
\begin{align}\label{Xk}
\sum\limits_{k=1}^{n}|X_k|&\leq\sum\limits_{k=1}^{n}\sum_{l=1}^{k-1}j_k(l)+\sum\limits_{k=1}^{n}g_k\nonumber\\
&\leq \sum_{k=1}^{n}\left(\frac{C_1}{k^{2-2H}}\sum_{l=1}^{\frac{k}4}\frac1{l^{H-\frac12}}+C_2\sum_{l=1}^{\frac{3k}4}\frac1{l^{\frac32-H}}\right)+2^{H-\frac12}g\,n\nonumber\\
&\leq \tilde{C}\sum_{k=1}^n k^{H-\frac12}+2^{H-\frac12}g\,
n\nonumber\\
&\leq\hat{C}_1\, n^{H+\frac12},
\end{align}
where $\tilde{C}$ and $\hat{C}_1$ are appropriate strictly positive constants. Similarly, we have
$$\sum\limits_{k=2}^{n}|X_{k-1}|\leq \hat{C}_1\,n^{H+\frac12}.$$
Hence, we deduce that
\begin{align}\label{Ca}
|\Vs_n^1|\leq\hat{C}_1\frac{n^{H+\frac12}}{N^{1-H}}\leq \hat{C}_1 N^{2H-\frac12}.
\end{align}
For the term $\Vs_n^2$ in \eqref{abc}, we proceed in a similar way, and we obtain, for some constant $\hat{C}_2>0,$ that
\begin{align}\label{Cb}
|\Vs_n^2|\leq \hat{C}_2 N^{2H-\frac12}.
\end{align}
It is left to find an estimate for $\Vs_n^3$. Using \eqref{theta} we write
\begin{align}\label{C1234}
|\Vs_n^3|&\leq \sum\limits_{k=1}^{n}|\phi_{k-1}^{1,N}\Delta_k S^N|=\frac1{N}\sum\limits_{k=1}^{n}|X_{k-1}X_k|\nonumber\\
&\leq\frac1{N}\sum_{k=1}^n\left|\theta_k^{(1)}+\theta_k^{(2)}+\theta_k^{(3)}\right|+\frac1{N}\sum_{k=1}^n\left|\theta_k^{(4)}\right|.
\end{align}
From Lemma \ref{th12}, we have, for $C_{1,2,3}=C_1^\theta+C_2^\theta+C_3^\theta>0$, that
$$\sup\limits_{1\leq k\leq N}\left|\theta_k^{(1)}+\theta_k^{(2)}+\theta_k^{(3)}\right|\leq C_{1,2,3}\,N^{H-\frac{1}{2}}.$$
We conclude that
$$\frac{1}{N}\,\sum_{k=1}^n\left|\theta_k^{(1)}+\theta_k^{(2)}+\theta_k^{(3)}\right|\leq C_{1,2,3} N^{H-\frac{1}{2}}$$
For the last term in \eqref{C1234}, we first notice that, using Lemma \ref{eji} and performing a similar calculation like in \eqref{Xk}, one gets
that $\sum_{\ell=1}^{k-1}j_k(\ell)\leq \bar{C} k^{H-\frac12}$, for some constant $\bar{C}>0$. Using this and the definition of $\theta_k^{(4)}$, we obtain
\begin{align*}
\frac1{N}\sum_{k=1}^n\left|\theta_k^{(4)}\right|\leq \frac{\bar{C}^2}{N}\sum_{k=1}^n k^{2H-1}\leq \bar{C}^2 N^{2H-1}.
\end{align*}
Hence, we derived that
\begin{align}\label{Cc}
|\Vs_n^3|\leq \hat{C}_3 N^{2H-1},
\end{align}
for an appropriate constant $\hat{C}_3>0$.
From \eqref{abc} we deduce that:
\begin{align}\label{vphi}
V_n^{\la_N}(\hat{\phi}^{N}(\lambda_N))\geq V_n(\hat{\phi}^{N}) -c_* \lambda_N\,N^{2H-\frac12},
\end{align}
for some constant $c_*>0$.
We return to the self-financing trading strategy $\psi^N$. Thanks to \eqref{vpr}, we get
\begin{equation}\label{vapsi}
V_n^{\la_N}(\psi^{N}(\lambda_N))\geq V_n(\psi^{N}) -c_*\frac{\lambda_N}{c_N} \,N^{2H-\frac12}.
\end{equation}
Since $\psi^N$ is $c_N$-admissible, we deduce that
\begin{equation*}
V_n^{\la_N}(\psi^{N}(\lambda_N))\geq-c_N -c_*\frac{\lambda_N}{c_N} \,N^{2H-\frac12}=:-c_N(\lambda_N),
\end{equation*}
or equivalently, that $\psi^N(\lambda_N)$ is $c_N(\lambda_N)$-admissible. Note that, it is enough to choose
$$\la_N=o\left(\frac{c_N}{N^{2H-\frac12}}\right)=o\left(\frac{\sqrt{\ln{N}}}{N^{(2H-\frac14)\wedge(H+\frac12})}\right),$$
to have $c_N(\lambda_N)\xrightarrow[N\rightarrow\infty]{}0$.
The second condition of Definition~\ref{SAAd} follows immediately. Indeed, defining
$$C_N(\lambda_N):=C_N-c_*\frac{\lambda_N}{c_N} \,N^{2H-\frac12}\xrightarrow[N\rightarrow\infty]{}\infty,$$
and using \eqref{vapsi}, we obtain that
\begin{align*}
P\left(V_N^{\la_N}(\psi^{N}(\lambda_N))\geq C_N(\lambda_N)\right)&\geq P\left(V_N(\psi^{N})\geq C_N\right).\\
\end{align*}
The second condition follows from the properties of $(\psi^N)_{N\geq 1}$, and the desired result is proven.
\end{proof}
\bibliographystyle{plain}
|
\section{Introduction}
\label{intro}
The last five years have seen the emergence of a new class of
ultra-bright stellar explosions: superluminous supernovae
\citep[SLSNe; for a review see][]{Gal-Yam12}, some 50 times brighter
than classical supernova (SN) types. Data on these rare and extreme
events are still sparse, with only $\sim$50 SLSN detections reported
in the literature. These events have generally been poorly studied
with incomplete imaging and spectroscopy, leaving many aspects of
their observational characteristics, and their physical nature,
unknown.
Yet these SLSNe could play a key role in many diverse areas
of astrophysics: tracing the evolution of massive stars, driving feedback in
low mass galaxies at high redshift, and providing potential line-of-sight probes
of interstellar medium to their high-redshift hosts \citep{Berger12}.
SLSNe have also been detected to $z\sim4$ \citep{Cooke12}, far beyond
the reach of the current best cosmological probe, type Ia supernovae
(SNe Ia). If SLSNe can be standardised \citep[e.g.][]{inserra14}, as is done with SNe~Ia
\citep{tripp98,1993ApJ...413L.105P,Riess96}, then a new era of SN cosmology would be
possible, with the potential to accurately map the expansion rate of
the universe far into the epoch of deceleration \citep{delubac14}.
SLSNe have been divided into three possible types: SLSN-II, SLSN-I and
SLSN-R \citep[see Fig. 1 of][]{Gal-Yam12}. SLSNe-II show signs of interactions with CSM via narrow hydrogen lines
\citep[e.g.,][]{2007ApJ...659L..13O,2007ApJ...666.1116S}, and thus may simply represent
the bright end of a continuum of Type IIn SNe (although this is not well established).
SLSN-I are
spectroscopically classified as hydrogen free \citep{Quimby11}, and
are possibly related to Type Ic SNe at late times
\citep{Pastorello10}, but normal methods of powering such SNe (e.g.,
the radioactive decay of $^{56}$Ni or energy from the gravitational
collapse of a massive star) do not appear to be able to simultaneously reproduce
their extreme brightness, slowly-rising light curves, and decay rates
\citep{Inserra13}. Many alternative models have been proposed: the
injections of energy into a SN ejecta via the spin down of a young magnetar
\citep{Kasen10,Woosley10,Inserra13}; interaction of the SN ejecta with
a massive (3-5\,M$_{\odot}$) C/O-rich circumstellar material
\citep[CSM; e.g., ][]{2010arXiv1009.4353B}; or collisions between high-velocity shells
generated from a pulsational pair-instability event \citep{Woosley07}. These explanations
are still actively debated in the literature.
SLSNe-R are rare and characterised by possessing extremely long, slow-declining
light curves ($>200$\,days in the rest frame). SLSNe-R originally
appeared consistent with the death of $\gtrsim$100\,M$_{\odot}$ stars
via the pair instability mechanism \citep{Gal-Yam09}. However, new
observations -- and the lack of significant spectral differences between
SLSN-I and SLSN-R -- have challenged the notion that these classes are
truly distinct \citep{Nicholl13} and maybe better described together as Type Ic SLSNe \citep{Inserra13}.
In summary, the origin of the power source for SLSNe (of all types) remains unclear,
with the possibility that further sources of energy may be viable. A
more detailed understanding requires an increase in both the quantity
and quality of the data obtained on these events. However, finding
more SLSNe (of any type) with existing transient searches is
challenging, especially in the local universe where the rates are only
$\simeq10^{-4}$ that of the core-collapse SN rate
\citep{Quimby13,2014arXiv1402.1631M} thus making it hard to assemble
large samples of events over reasonable timescales, even from large surveys. For example,
\citet{2014arXiv1402.1631M} detected only 10 SLSN candidates over
$0.3\le z\le1.4$ within the first year of the PanSTARRS1 Medium Deep
Survey.
Searches for SLSNe at higher redshifts may be more profitable as the rates
of SLSNe appear to rise by a factor of $\sim$10-15 at $z>1.5$
\citep{Cooke12}, perhaps tracking the increased cosmic star-formation
rate and/or decreasing cosmic metallicity: SLSNe are preferentially
found in faint, low-metallicity galaxies \citep{Neill11,Chen13,lunnan14}.
Moreover, the peak of a SLSN's energy output is located in the UV
region of the electromagnetic spectrum, which is redshifted to optical
wavelengths at high redshift.
In this paper, we outline our first search for SLSNe in the Dark
Energy Survey (DES). In Section~\ref{data}, we discuss details of DES
and our preliminary search, while in Section~\ref{sec:des13s2cmm}, we
describe the first SLSN detected in DES and provide details of its
properties, including our spectroscopic confirmation. In
Section~\ref{sec:results}, we compare our SLSN to other events in the
literature and explore the possible power sources for our event. We
conclude in Section~\ref{sec:conclusions}. Throughout this paper, we
assume a flat $\Lambda$-dominated cosmology with $\Omega_{M}=0.28$ and
$H_0=70$\,km\,s$^{-1}$\,Mpc$^{-1}$, consistent with recent
cosmological measurements \citep[e.g.][]{anderson14, betoule14}.
\section{The Dark Energy Survey}
\label{data}
The Dark Energy Survey \citep[DES;][]{2005IJMPA..20.3121F} is a new
imaging survey of the southern sky focused on obtaining accurate
constraints on the equation-of-state of dark energy. Observations are
carried out using the Dark Energy Camera
\citep[DECam;][]{2012SPIE.8446E..11F,2012PhPro..37.1332D} on the
4-metre Blanco Telescope at the Cerro Tololo Inter-American
Observatory (CTIO) in Chile. DECam was successfully installed and
commissioned in 2012, and possesses a 3\,deg$^2$ field-of-view with 62
fully depleted, red-sensitive CCDs.
\subsection{DES Supernova Survey}
\label{sec:des-supernova-survey}
DES conducted science verification (SV) observations from November
2012 to February 2013, and then began full science operations in
mid-August 2013, with year one running until February 2014.
DES is performing two surveys in parallel: a wide-field, multi-colour
($grizY$) survey of 5000\,deg$^2$ for the study of clusters of
galaxies, weak lensing and large scale structure; and a deeper,
cadenced multi-colour ($griz$) search for SNe. In this paper we focus
on this DES SN Survey, as outlined in \citet{Bernstein12}, which
surveys 30\,deg$^2$ over 10 DECam fields (two `deep', eight `shallow')
located in the XMM-LSS, ELAIS-S, CDFS and `Stripe82' regions.
As of November 2014, the DES SN Survey has already discovered over a thousand SN Ia candidates, with more than 70 spectroscopic classifications to date.
This rolling search for SNe Ia is also ideal for discovering other
transient phenomena including SLSNe. Using the light curve of
SNLS-06D4eu \citep[a SLSN-I; ][]{Howell13}, we calculate that DES
could detect such an event to $z=2.5$ given the limiting magnitudes
for the DES deep fields. For example, assuming a volumetric rate of
$32\,\rmn{Gpc}^{-3}\,\rmn{yr}^{-1}$ from \citet{Quimby13} and
correcting for time dilation, we estimate DES should find
approximately 3.5 SLSN-I events per DES season (five months) in the
deep fields and many more in the shallow fields (although at a lower
redshift). This basic prediction is uncertain because: i) We have not
corrected for `edge effects' that will reduce the number of SLSNe with
significant light-curve coverage; ii) The volumetric rates themselves
may evolve with redshift \citep[e.g.,][]{Cooke12} and have large
uncertainties \citep{2014arXiv1402.1631M}; and iii) We have not
included any corrections for survey completeness.
\subsection{Selecting SLSNe candidates}
\label{sec:select-slsne-cand}
We began our search for SLSNe using the data products from the DES SN
Survey. All imaging data were de-trended and co-added using a standard
photometric reduction pipeline at the National Center for
Supercomputing Applications (NCSA) using the DES Data Management
system \citep[DESDM;][]{2012SPIE.8451E..0DM,2012ApJ...757...83D},
producing approximately 30 `search images' per field (in all filters)
over the duration of the five-month DES season. We perform difference
imaging on each of these search images, using deeper template images
for each field created from the co-addition of several epochs of data
obtained during the SV period in late 2012. Before differencing, the
search and template images were convolved to the same point spread
function (PSF).
Objects were selected from the difference images using
\textsc{SExtractor} \citep{1996A&AS..117..393B} v2.18.10, and previously
unknown transient candidates were identified and examined using
visual inspection (or `scanning') by DES team members. Over the course
of the first year of DES approximately $25,000$ new transient
candidates were selected in this manner, including many data artefacts
in addition to legitimate SN-like objects. A machine learning algorithm (Goldstein et al., in preparation)
was used to improve our efficiency of selecting real transients. Transient candidates were
photometrically classified using the Photometric SN IDentification
software \citep[PSNID;][]{Sako08,Sako11} to determine the likely SN
type based on fits to a series of SN templates. These SN candidates
were stored in a database and prioritized for their spectroscopic
follow-up.
During the first season of DES, PSNID only included templates for the
normal SN types Ib/c, II and Ia, so a fully automated classification
of SLSN candidates was not possible. Therefore, for this first season
of DES, we searched for candidate SLSNe using the following criteria:
i) At least one month of multi-colour data, i.e., typically five to six
detections ($S/N>3.5$) in each of $griz$; ii) A low PSNID fit probability to any of the
standard SN sub-classes; iii) Located greater than one pixel from the
centroid of the host galaxy (to eliminate AGN); and iv) Peak observed brightness no fainter than one magnitude below that of its host galaxy. These broad criteria will be
satisfied by most SLSNe, while also helping to eliminate many of the
possible contaminating sources.
Using this methodology, we selected 10 candidate SLSNe over the course
of the first season. We secured a useable spectrum for
one of our SLSN candidates, DES13S2cmm, which is discussed in
detail in Section~\ref{sec:des13s2cmm}, while none of the other candidates
were spectroscopically confirmed. We are attempting to obtain spectroscopic redshifts from the host galaxies of these other SLSN candidates, though this will be challenging in the several cases where no host is detected in our template images.
\section{DES13S2cmm}
\label{sec:des13s2cmm}
\begin{figure*}
\includegraphics[width=0.99\textwidth,clip]{DES13S2cmm_TRIPLET_edited.eps}
\caption{Colour images (\textit{riz}) of the field surrounding
DES13S2cmm. \textit{Left:} Deep template image created from
the co-addition of several epochs of data obtained during
the DES Science Verification period in late 2012; the likely
host galaxy is indicated by the arrow. \textit{Centre:} The
search image taken close to maximum light (28-Sept-2013)
with the position of DES13S2cmm indicated. \textit{Right:}
The difference of the two previous images, clearly
identifying DES13S2cmm at center. This image demonstrates
the quality of difference images used for discovery and
monitoring during the DES observing season.}
\label{fig:triplet}
\end{figure*}
In Fig. \ref{fig:triplet} we show DES \textit{riz} imaging data at the
position of DES13S2cmm: pre-explosion, post-explosion, and their
difference image. DES13S2cmm is located at
$\rmn{RA}(2000)=02^{\rmn{h}} 42^{\rmn{m}} 32\fs82$,
$\rmn{Dec.}~(2000)=-01\degr 21\arcmin 30\farcs 1$, and was first
detected on 2013 August 27 in the DES SN `S2' field (a `shallow'
field) located in the Stripe 82 region. DES transient names are formatted
following the convention \hbox{DESYYFFaaaa}, where YY are the last two digits of the year in
which the observing season began (13), FF is the two character field name (S2), and the final
characters (all letters, maximum of 4) provide a running candidate
identification, unique within an observing season, as is traditional
in SN astronomy (cmm).
\subsection{Light Curve}
\label{sec:light-curve}
\begin{figure*}
\includegraphics[width=0.9\textwidth,clip]{S2cmm_LC_photometric.eps}
\caption{The observed AB magnitude light curve of DES13S2cmm
in the four DES SN search filters ($griz$) as a function of
the observed phase. `S' denotes the epoch when the ESO VLT
spectrum was obtained. The error-bars represent 1-$\sigma$
uncertainties, and the arrow symbols are 3-$\sigma$ upper
limits. The phase relative to peak was defined using the
observed-frame $r$-band. We have artificially offset (in
magnitudes) the $r$, $i$ and $z$ band data for clarity.}
\label{fig:light_curve}
\end{figure*}
In Fig.~\ref{fig:light_curve} we present the multi-colour light curve
for DES13S2cmm constructed from the first year of DES observations. We
have not included upper limits from the earlier, pre-explosion epochs
available from the DES SV period in 2012-2013, as these data were
taken hundreds of days prior to the data shown in
Fig.~\ref{fig:light_curve} and are already used to create the
reference template image for this field used in the difference imaging
detection of the SNe.
We use the photometric pipeline from
\citet{Sullivan11} \citep[also used in][]{Maguire12,Ofek13} to reduce
the DES photometric data for DES13S2cmm. This pipeline is based on
difference imaging, subtracting a deep, good seeing, pre-explosion
reference image from each frame the SN is present in. The photometry
is then measured from the differenced image using a PSF-fitting method, with the PSF determined
from nearby field stars in the unsubtracted image. This average PSF is
then fit at the position of the SN event, weighting each pixel
according to Poisson statistics, yielding a flux, and flux error, for
the SN.
We calibrate the flux measurements of DES13S2cmm to a set of tertiary
standard stars produced by the DES collaboration (Wyatt et al., in preparation), intended to be in
the AB photometric system \citep{Oke83}.
The photometry shown in Fig.~\ref{fig:light_curve} has also been
checked against the standard DES detection photometry, used to find
and classify all candidates for follow-up, and found to be in good
agreement. The light-curve data are presented in
Table~\ref{tab:photometry_table}.
We correct for Galactic extinction using the maps of \citet{SF11}, which estimate ${\rm E(B-V)}=0.028$ at the location of DES13S2cmm,
leading to extinction values of 0.123, 0.070, 0.051 and 0.039 magnitudes in the DES $griz$ filters respectively. We do not include these
corrections in the values reported in Table~\ref{tab:photometry_table}, but do correct for extinction in all figures and analysis herein,
including Figure~\ref{fig:light_curve}.
\begin{table*}
\begin{center}
\caption{Light curve data for DES13S2cmm used in
Fig.~\ref{fig:light_curve}. We provide the observed calendar date,
modified Julian Day (MJD), observed phase relative to peak MJD
(defined relative to peak flux in the \textit{r}-band), and fluxes (not corrected for Galactic extinction)
with 1-$\sigma$ uncertainties in the DES \textit{griz} passbands. These fluxes
can be converted to AB magnitudes using
$\rmn{mag}_{\rmn{AB}}=-2.5\log_{10}\left(f\right)+31$. }
\begin{tabular}{ c | c | c || r @{$\pm$}l | r @{$\pm$} l | r @{$\pm$} l | r @{$\pm$} l |}
\hline \hline
\multicolumn{1}{|c|}{Date} &\multicolumn{1}{|c|}{MJD} & \multicolumn{1}{|c|}{Phase} & \multicolumn{2}{|c|}{$f_g$} & \multicolumn{2}{|c|}{$f_r$} & \multicolumn{2}{|c|}{$f_i$} & \multicolumn{2}{|c|}{$f_z$} \\
&&(days)&\multicolumn{2}{|c|}{}&\multicolumn{2}{|c|}{}&\multicolumn{2}{|c|}{}&\multicolumn{2}{|c|}{}\\
\hline
30-AUG-2013 & 56534.3 & -25 & 1138 & 77 & 1705 & 94 & 1730 & 181 & 1913 & 231 \\
3-SEP-2013 & 56538.3 & -21 & 1112 & 103 & 2063 & 129 & 2234 & 186 & 1845 & 288 \\
8-SEP-2013 & 56543.3 & -16 & 1369 & 104 & 2348 & 148 & 2285 & 198 & 2285 & 277 \\
12-SEP-2013 & 56547.2 & -12 & 1373 & 130 & 2329 & 167 & 2433 & 207 & 2376 & 284 \\
15-SEP-2013 & 56550.2 & -9 & & & 2691 & 413 & 2818 & 380 & 2193 & 389 \\
24-SEP-2013 & 56559.2 & 0 & 2022 & 388 & 3727 & 363 & 3183 & 346 & 3548 & 295 \\
28-SEP-2013 & 56563.2 & 4 & 1668 & 97 & 3688 & 133 & 3400 & 187 & 3488 & 241 \\
2-OCT-2013 & 56567.2 & 8 & 1805 & 141 & 3666 & 166 & 3742 & 209 & 3169 & 261 \\
10-OCT-2013 & 56575.2 & 16 & 1370 & 108 & 3285 & 137 & 3636 & 192 & 3182 & 230 \\
14-OCT-2013 & 56579.1 & 20 & 1260 & 243 & 3182 & 223 & 3482 & 259 & 2663 & 295 \\
25-OCT-2013 & 56590.3 & 31 & 886 & 346 & 1782 & 319 & 2924 & 277 & 3163 & 352 \\
29-OCT-2013 & 56594.1 & 35 & 667 & 146 & 1782 & 187 & 2398 & 267 & 2987 & 366 \\
6-NOV-2013 & 56602.1 & 43 & 344 & 103 & 1698 & 132 & 2326 & 185 & 2433 & 257 \\
10-NOV-2013 & 56606.1 & 47 & 803 & 426 & 1338 & 418 & 2633 & 510 & 1817 & 634 \\
13-NOV-2013 & 56609.1 & 50 & -312 & 276 & 970 & 252 & 2672 & 281 & 2726 & 426 \\
20-NOV-2013 & 56616.1 & 57 & 347 & 195 & 1855 & 239 & 1848 & 307 & 1739 & 331 \\
1-DEC-2013 & 56627.0 & 68 & 184 & 107 & 1306 & 129 & 2032 & 176 & 2322 & 216 \\
9-DEC-2013 & 56635.0 & 76 & 259 & 208 & 1322 & 178 & 1766 & 225 & 2318 & 304 \\
19-DEC-2013 & 56645.1 & 86 & 302 & 226 & 1305 & 211 & 1697 & 230 & 1808 & 248 \\
23-DEC-2013 & 56649.1 & 90 & 261 & 136 & 956 & 157 & 1667 & 248 & 2038 & 276 \\
27-DEC-2013 & 56653.1 & 94 & 41 & 156 & 1295 & 174 & 1613 & 201 & 1930 & 248 \\
3-JAN-2014 & 56660.0 & 101 & 252 & 119 & 1000 & 132 & 1736 & 196 & 1993 & 261 \\
10-JAN-2014 & 56667.1 & 108 & 310 & 343 & 1144 & 364 & 1609 & 408 & 1586 & 387 \\
18-JAN-2014 & 56675.0 & 116 & -75 & 236 & 717 & 220 & 1290 & 261 & 1863 & 297 \\
25-JAN-2014 & 56682.0 & 123 & 306 & 158 & 992 & 188 & 1382 & 215 & 1254 & 270 \\
30-JAN-2014 & 56687.0 & 128 & 345 & 188 & 523 & 179 & 1518 & 217 & 1404 & 261 \\
6-FEB-2014 & 56694.0 & 135 & -346 & 359 & 1459 & 376 & 1602 & 465 & 2443 & 461 \\
\hline
\end{tabular}
\label{tab:photometry_table}
\end{center}
\end{table*}
\subsection{Spectroscopy}
\label{sec:spectrum}
On 2013 November 12, we requested Director's Discretionary Time
at the European Southern Observatory (ESO) Very Large Telescope (VLT)
to observe DES13S2cmm, and were awarded target-of-opportunity time within days.
The closeness of the object to the moon and instrument scheduling
issues conspired to delay observations until 2013 November 21, at
which point DES13S2cmm was approximately 30 days after peak brightness
(rest frame). A spectrum was obtained with the FOcal Reducer and low
dispersion Spectrograph \citep[FORS2][]{Appenzeller98} using the
GRIS\_300I+11 grism, the OG590 order blocker, and a 1\arcsec\ slit,
with an exposure time of 3600s (3$\times$1200s). This configuration
provided an effective wavelength coverage of $5950$\,\AA\ to
$9400$\,\AA.
The reduction of the FORS2 data followed standard procedures using the
Image Reduction and Analysis Facility\footnote{The Image Reduction and
Analysis Facility (\textsc{iraf}) is distributed by the National
Optical Astronomy Observatories, which are operated by the
Association of Universities for Research in Astronomy, Inc., under
cooperative agreement with the National Science Foundation.}
environment (v2.16), using the pipeline described in \citet{Ellis08}. This
pipeline includes an optimal two-dimensional (2D) sky subtraction
technique as outlined in \citet{Kelson03}, subtracting a 2D sky frame
constructed from a sub-pixel sampling of the background spectrum and a
knowledge of the wavelength distortions determined from 2D arc
comparison frames. The extracted 1D spectra were then scaled to the
same flux level, the host-galaxy emission lines interpolated over, and the
spectra combined using a weighted mean and the uncertainties in
the extracted spectra.
\begin{figure*}
\includegraphics[width=0.95\textwidth,clip]{2Dspectrum_v2.eps}
\caption{The reduced VLT FORS2 two-dimensional spectrum for DES13S2cmm. The key nebular emission lines used to derive the redshift of the host galaxy, and therefore supernova, are labelled.}
\label{2Dspectrum}
\end{figure*}
In Fig \ref{2Dspectrum} we present our reduced 2D FORS2 spectrum, where wavelength increases to the right along the horizontal axis and the vertical axis is distance along the slit. We highlight the obvious narrow nebular emission lines in the host galaxy of DES13S2cmm, specifically [O\,\textsc{ii}]~$\lambda3727$\,\AA, H$\beta$ $\lambda4861$\,\AA, and [O\,\textsc{iii}]~$\lambda4959,5007$\,\AA. We assume these lines originate from the underlying host galaxy as their spatial extent along the vertical axis is greater than the observed trace of the main supernova spectrum. This observation does not exclude the possibility that such emission lines come from the supernova, but it is difficult to address this issue with our low resolution, and low signal-to-noise, data. We note the observed emission lines are not significantly broadened or asymmetrical, as witnessed in type IIn supernovae.
We see no evidence of any further emission lines, from the SN and/or host galaxy. Therefore, it is unlikely DES13S2cmm is a type II supernova (normal or superluminous).
Using these nebular emission lines, we determine a redshift of $z=0.663\pm0.001$ for the host galaxy of
DES13S2cmm, where the uncertainty is derived from the
wavelength dispersion between individual emission lines.
This measurement is consistent with the redshift of
the SN based on identifications of the broad absorption features seen
in the spectrum. As the redshift derived from the host-galaxy spectrum is significantly
more precise, henceforth we adopt this value as the redshift to DES13S2cmm.
The spectral classification for DES13S2cmm is based on the
\textsc{superfit} program \citep{Howell05}, which compares
observed spectra of SNe to a library of template SN spectra (via
$\chi^2$ minimisation) while accounting for the possibility of host-galaxy
contamination in the data. We added as template spectra of SLSNe-I
(PTF09atu, PTF09cnd, PTF10nmn) and SLSN-R (SN\,2007bi) to
\textsc{superfit} to facilitate our classification.
We found that the highest-ranked fits to
DES13S2cmm were these SLSNe templates, preferred over templates of normal SN types (e.g., SN Ia). The SLSNe-I templates were slightly better fits to DES13S2cmm than SN\,2007bi (SLSN-R), although this may be due to not having a template for SN\,2007bi at a similar phase to our spectrum for DES13S2cmm.
\begin{figure*}
\includegraphics[width=0.9\textwidth,clip]{specmatch_v4.eps}
\caption{The VLT FORS2 spectrum of DES13S2cmm (second from
top) compared to other SLSNe of similar phase in their light
curves. From top to bottom, we also show SN\,2011kf
\citep{Inserra13}, PTF10hgi \citep{Quimby11}, SN\,2011ke
\citep{Inserra13} and SN\,2007bi\citep[classified as Type R in][]{Gal-Yam09}; as discussed in
Section \ref{sec:spectrum}, \textsc{superfit} identifies these SLSNe as
some of the best-fitting spectral templates to DES13S2cmm. In each
case, the light grey shows the raw data and the solid line
is a Savitsky-Golay smoothed version of the spectrum. The
spectra are labelled with the SN name, SLSN type (except
DES13S2cmm), redshift, and the phase when the spectrum
was taken. We show the spectrum of the `super-Chandrasekhar'
SN Ia SN\,2007if as discussed in \citet{scalzo10}. Note that the
spectrum for 2007bi, while at a later phase than the other data, is the
earliest spectral observation of this SN.}
\label{fig:spectra}
\end{figure*}
We show our spectrum of DES13S2cmm in Fig.~\ref{fig:spectra}, alongside
spectra from other well-studied SLSNe in the literature at a similar
phase in their light curves \citep{Quimby11,Inserra13,Gal-Yam09}.
There is a deficit of spectra for SLSNe in the literature at these
late phases for a robust comparison. However, the broad resemblance
between the spectra of PTF10hgi, SN\,2011kf, SN\,2011ke and
DES13S2cmm, all taken at approximately $+30$ days past peak, suggests
DES13S2cmm is a similar object to these other SLSNe-I. The spectrum of
DES13S2cmm is also similar to the spectrum of SN\,2007bi (a SLSN-R in \citealp{Gal-Yam09}), although the differences between Type I and Type R SLSNe remains unclear, and either classification would be interesting given the sparseness of the
data. We also show in Fig.~\ref{fig:spectra} the spectrum of
SN\,2007if, a `super-Chandrasekhar' SN Ia \citep{scalzo10}, an
over-luminous ($M_V = -20.4$) and likely thermonuclear SN located in a
faint host galaxy. This comparison clearly shows that DES13S2cmm is
unlikely to be a super-Chandrasekhar SN Ia.
\subsection{Host-galaxy properties}
\label{sec:host-galaxy}
The host galaxy of DES13S2cmm is detected in our reference images that
contain no SN light. Using \textsc{sextractor} in dual-image mode and the $i$-band image as the detection image, we
measure host-galaxy magnitudes of $g=24.24\pm0.13$,
$r=23.65\pm0.10$, $i=23.31\pm0.10$ and $z=23.26\pm0.17$ (AB \textsc{mag\_auto}
magnitudes). The large photometric errors are due to the relatively
shallow depth, and poor seeing, of the template image from the DES SV data -- we cannot use the DES first-year data for
the host photometry as all epochs of this data contain some light from
DES13S2cmm.
The estimated photometric redshift of the host galaxy at the time of discovery was
$z_{\rmn{photo}}=0.86\pm0.13$ (68 percentile error). This photometric
redshift was derived by the DESDM neural network photo-z module code, with uncertainties estimated
from the nearest neighbour method, using data taken during SV \citep{sanchez14}.
Photometric redshifts (when available) are
used when selecting SLSN candidates for spectroscopic follow-up, and
for DES13S2cmm this redshift implied the event was brighter than
$M^{\rmn{peak}}_U =-21$ \citep[typical for SLSNe; ][]{Gal-Yam12}.
Since the start of the first DES observing season the photometric catalog derived from SV data has improved, and presently the DESDM photo-z for the host of DES13S2cmm has improved to $z_{\rmn{photo}}=0.71\pm0.06$, which is in better agreement with our measured spectroscopic redshift ($z=0.663\pm0.001$).
Using the DES host-galaxy photometry and the spectroscopic redshift from VLT, we estimate that the stellar mass of the host galaxy is
$\log(M/M_{\odot})=9.3\pm0.3$ using stellar population models from
PEGASE.2 \citep{1997A&A...326..950F}, and
$\log(M/M_{\odot})=9.0\pm0.3$ using the stellar population templates
from \citet{2005MNRAS.362..799M} (both corrected for Milky Way extinction and fixing the redshift of the templates to $z=0.663$). Such a low stellar-mass host galaxy
is consistent with the findings for other SLSNe \citep{Neill11,Chen13,lunnan14}.
We estimate the host-galaxy metallicity from our VLT spectrum using the double-valued metallicity indicator $R_{23}$, defined as $([\textrm{O}\textsc{II}]~\lambda 3727 + [\textrm{O}\textsc{III}]~\lambda\lambda 4959, 5007)/H\beta$. Fluxes and uncertainties are derived individually for each emission line from fits of a gaussian plus a first-order polynomial, with the assumption that there is no contamination of the galaxy emission lines from the SN. We compute metallicities using the calibrations of \citet{KK04} and \citet{M91}, and use the formulae derived in \citet{KE08} to convert these metallicities into the calibration of \citet{KD02} -- a step that allows for direct comparison between different methods. We note that since $H\alpha$ and $\textrm{N}\textsc{II}~\lambda6584$ are redshifted to wavelengths beyond our spectral coverage, we cannot determine the branch of the $R_{23}$ function that the metallicity lies on. However, the derived value for $R_{23}$ is close to its theoretical maximum, which minimizes the difference in metallicity estimates between the two branches.
Assuming the lower (upper) branch, we find a metallicity ($12+\log$[O/H]) of 8.30 (8.38) from \citet{M91} and 8.30 (8.42) from \citet{KK04}, which is consistent with the median metallicity of $8.35$ found by \citet{lunnan14} for a sample of 31 SLSNe host galaxies. The total uncertainty is dominated by the calibration ($\sim0.15$ dex), while the statistical uncertainty and the scatter induced by the conversion to \citet{KD02} are comparatively negligible. We thus find a host-galaxy abundance ratio that is distinctly sub-solar \citep[8.69;][]{Asplund09}, though not as extremely low as has been seen for other SLSNe, such as SN2010gx ($12+\log$[O/H]=7.5;~\citealp{Chen13}).
\section{Results}
\label{sec:results}
\subsection{Bolometric light-curve of DES13S2cmm}
\label{peakM}
\begin{figure*}
\includegraphics[width=0.9\textwidth,clip]{All_bolometric.eps}
\caption{The bolometric light curve of DES13S2cmm (red
circles) compared to other SLSNe in the literature (see text for explanation). These bolometric
light curves were constructed by fitting a single black-body
spectrum to the multicolour data available per epoch
(requiring at least three photometric measurements per
epoch), and then by integrating the black-body spectral
energy distribution to produce the bolometric luminosity per
rest-frame epoch (described in detail in
Section~\ref{Bol_lc}). We show statistical errors on the blackbody fits to each epoch of these light curves, but in most cases this error is
smaller than the plotting symbol. We stress that these errors do not account for systematic uncertainties
involved in the bolometric calculation (e.g., assuming a blackbody
for the spectral energy distribution of SLSNe, especially for the UV part of the rest-frame spectral energy distribution). The phase relative to peak (y-axis) for the literature SLSNe is taken directly from the
published individual studies and not recalculated herein. The dashed lines shown in both
panels represent the expected $^{56}$Co decay rate at late times. In the lower panel, we show the bolometric light curve expected for a typical Type Ia SNe (solid line) to again illustrate the extraordinary brightness, and extended duration, of these SLSNe}
\label{fig:Bol_lc}
\end{figure*}
We show in Fig.~\ref{fig:Bol_lc} the bolometric light curve of
DES13S2cmm. This single light curve was constructed by fitting a
single black-body spectrum to the DES multi-colour data (requiring
measurements in a minimum of three photometric bandpasses for a fit),
on each epoch in the light curve (Table 1). In detail, we redshift
black-body spectra to the observer frame and integrate these spectra
through the response functions of the DES filters. We then determine
the best-fitting parameters for a black body to the
extinction-corrected photometry and integrate this spectrum in the
rest frame to obtain the bolometric luminosity per epoch. This
procedure benefits from the homogeneity of the DES data as we possess
approximately equally-spaced epochs across the whole light curve --
typically in four DES passbands -- due to our rolling search. The
errors on the bolometric luminosities were calculated from the fitting
uncertainties on the black-body spectra.
We visually checked our best-fit blackbody curves for each epoch of the DES13S2cmm light curve against the observed $griz$ photometry and found them to be reasonable within the errors. The largest discrepancies were seen in the g-band where some epochs were below the fitted curves, possibly due to UV absorption in the underlying spectral energy distributions relative to a blackbody curve. Inserra et al. (2013) compared the predicted UV flux (from a blackbody curve) for two SLSNe with SWIFT UV observations, and found the fluxes were consistent within the error. Therefore we make no additional correction
to the derived bolometric light curve, but note that quantifiable systematic errors
in the bolometric luminosity require spectral templates, which in turn requires
more spectral data of SLSNe than is currently publicly available.
Using the bolometric light curve of DES13S2cmm, we can determine the
peak luminosity of the event and compare it to the definition of a
SLSNe in \citet{Gal-Yam12}. This was achieved by integrating the
best-fitting black-body spectra through a standard $U$-band filter
response function (on the Vega system) to obtain the rest-frame absolute magnitude in
$U$-band ($M_U$). We find the 2013 September 24 epoch ($t=0$ relative
phase in Table 1) gives the largest (peak) absolute $U$-band magnitude
of $M^{peak}_U=-21.05^{+0.10}_{-0.09}$, which is consistent with the
SLSNe threshold of $M^{peak}_U<-21$ in \citet{Gal-Yam12}.
We check this result using a model-independent estimate based on the observed $r$-band peak magnitude.
At the redshift of DES13S2cmm, the observer-frame $r$ band maps into the rest frame as a synthetic filter
with an effective wavelength of $\approx3800$\AA, slightly redward of the standard $U$-band. Defining a synthetic filter
as such means the k-correction is independent of the SN spectral energy distribution, and since we are in the AB system
the correction is simply $2.5\log(1+z)$. We apply this k-correction to the observed $r$-band peak magnitude, as well as
corrections for Galactic extinction and distance modulus, to yield an estimated peak absolute magnitude
of $-20.47$ (AB), or $-20.74$ (Vega) in our synthetic filter (where we have computed the AB to Vega conversion for our filter).
While this is fainter than our blackbody-based estimate, Vega magnitudes are sensitive to the exact
shape of the filter in this region due to the jump in Vega from 3700-3900\AA; thus even the small difference in effective
wavelength between $U$ band and our synthetic filter can result in significant offsets. The absolute magnitude in our
synthetic filter from the best-fitting blackbody method yields $M_{\rm synthetic}=-20.59$, proving that our
model-independent estimate is consistent with our results from the blackbody fitting.
We note that the peak magnitude (Vega) for DES13S2cmm classifies it
as a SLSN, provided the magnitude is intrinsic to the SN and not
enhanced by other effects. Visual inspection of the images in
Fig.~\ref{fig:triplet} (including the surrounding areas) shows little
evidence for strong gravitational lensing which
could affect the observed brightness of this SN
(we note that the host galaxy would have to be lensed as well). We cannot
conclusively rule out the possibility of strong lensing, as in the
case of PS1-10afx \citep{Quimby14}, but as discussed above the
spectrum and extended light curve of DES13S2cmm are inconsistent
with a normal SN, while the probability of such a strong lensing event
in our DES data is less than the probability of discovering a SLSN,
based on the lensing statistics given in \citet{Quimby14}, and the
lower redshift of DES13S2cmm and smaller search area of DES. Moreover, the observed $r-i$ peak colour of DES13S2cmm is consistent with the expected colours of other unlensed SNe (including SLSNe) shown in Figure 4 of \citet{Quimby14}, i.e., DES13S2cmm is located below the thick black line in this plot.
Therefore, we do not consider gravitational lensing further in this
paper, but higher-resolution imaging data should be obtained for
DES13S2cmm, after the SN event has vanished, to provide better
constraints on possible strong lensing as discussed in
\citet{Quimby14}.
\subsection{Comparison of bolometric light-curves}
\label{Bol_lc}
For comparison with DES13S2cmm, we also show in Fig.~\ref{fig:Bol_lc}
the bolometric light curves for fourteen SLSNe-I (or Ic, as discussed in \citealp{Inserra13}) in the literature: PS1-10ky. PS1-10awh \citep{Chomiuk11} and PS1-10bzj \citep{Lunnan13}; SN\,2010gx \citep{Pastorello10}; PTF\,10hgi, SN\,2011ke, PTF\,11rks, SN\,2011kf and SN\,2012il \citep{Inserra13}; SNLS\,06D4eu and SNLS\,07D2bv \citep{Howell13}; and LSQ\,12dlf, SSS\,120810 and SN\,2013dg \citep{Nicholl14}. We exclude from this figure, and our analysis, other SLSNe-I that do not possess data for at least three passbands for each epoch (defined as being taken within a 24 hour window) over a majority of their light curves. This criterion, so defined to ensure well-constrained bolometric light-curve fits, excludes namely PTF\,09atu, PTF\,09cnd, and PTF\,09cwl \citep{Quimby11}; SCP\,06F6 \citep{Barbary09}; and SN\,2006oz \citep{Leloudas12}.
The full bolometric light curves for these SLSNe-I were calculated using the same methodology as described above for DES13S2cmm (i.e. using the published photometric data and requiring three or more passbands per epoch). As a check, our bolometric light curves for PTF\,10hgi, SN\,2011ke, PTF\,11rks, SN\,2011kf and
SN\,2012il are similar to those published in \citet{Inserra13}, with
any offsets due to differences in the integration of
the blackbody curve; we integrate over all wavelengths, while
\citet{Inserra13} integrate over the wavelength range of their observed
optical passbands (we are able to reproduce their luminosities if we
follow their procedure). Our bolometric light-curves for PS1-10ky. PS1-10awh, and PS1-10bzj are also similar to those published in the literature.
In Fig.~\ref{fig:Bol_lc}, the phase relative to peak for DES13S2cmm was defined as
the peak of the observer-frame r-band, which occurred on MJD 56563.2, while for other
SLSNe-I we simply used the phase relative to peak as reported by the individual
literature studies. In the case of SNLS\,06D4eu and SNLS\,07D2bv, \citet{Howell13} reported a phase based on their calculation of the bolometric light curves, which we use herein, but note there may be greater uncertainty in the definition of the time of peak for these light curves.
The top panel of Fig.~\ref{fig:Bol_lc} shows that DES13S2cmm is one of the faintest SLSNe-I around peak compared to the other SLSNe-I in the literature, and possesses one of the slowest declining tails (beyond approximately $+30$ rest-frame) of the SLSNe studied herein. In the bottom panel of Fig.~\ref{fig:Bol_lc}, we compare DES13S2cmm to possible Type R SLSNe (to be consistent with the classification scheme of \citealp{Gal-Yam12}) and find the late-time light curve of DES13S2cmm is similar to both PS1-11ap \citep{2014arXiv1402.1631M} and SN2007bi \citep{Gal-Yam09}, highlighting the possible overlap between these two SLSNe types, as discussed in Section~\ref{sec:spectrum} and \citet{Inserra13}.
Fig.~\ref{fig:Bol_lc} also demonstrates that the fifteen SLSNe-I light curves (top panel) have similar luminosities at approximately $+25$ days (rest frame) past peak, close to the inflection point in the light curves when their extended
tails begin to appear. We calculate the dispersion, as a function of relative phase, for the SLSNe-I light curves shown in Fig.~\ref{fig:Bol_lc}, linearly interpolating between large gaps in each light curve to provide an evenly sampled set of data. We find a minimum dispersion of $\sigma({\rm log_{10} L_{bol}})=0.11$ -- equivalent to $0.29$ magnitudes -- around $+30$ days past peak (rest frame), derived from the nine SLSNe in Fig.~\ref{fig:Bol_lc} constraining this phase. This dispersion in the light-curves can be decreased to $\sigma({\rm log_{10} L_{bol}})=0.083$, or $0.20$ magnitudes, at approximately $+25$ days past peak (rest frame) if we exclude the four highest redshift SLSNe-I above $z\simeq1$ (SNLS\,06D4eu, SNLS\,07D2bv, PS1-10ky. PS1-10awh). The bolometric luminosities of these SLSNe-I are more difficult to estimate because of the uncertainties associated with modelling the UV part of the spectrum by assuming a simply blackbody curve. There could also be evolution in the population with redshift.
As commented in \citet{Quimby11}, such light curve characteristics
could point towards a possible `standardisation' of SLSNe-I, as with
SNe Ia, leading to their use as high-redshift `standard candles' for
cosmological studies \citep{king14}.
Recently, \citet{inserra14} has proposed a `standardisation' of the peak magnitude of 16 SLSNe using a
stretch-luminosity correction similar to the $\Delta m_{15}$
relationship of \citet{1993ApJ...413L.105P} for SNe Ia. They find that SLSNe that are brighter at peak have slower declining light curve over the first $30$ days after peak (rest frame) and, correcting for this correlation, the
dispersion of the peak magnitude of SLSNe can be reduced to $\sigma\simeq 0.2$. This is similar in size to the dispersion between $+25$ to $+30$ days past peak seen in Fig.~\ref{fig:Bol_lc} and is only a factor of two worse than the dispersion obtained for standardised SNe Ia.
Both Fig.~\ref{fig:Bol_lc} and \citet{inserra14} suggest some standardisation of SLSNe light curves is possible but further observations and analyses are required to find the best approach. For example, the light curve of DES13S2cmm does not appear to follow the correlation found by \citet{inserra14} as it has a faint peak magnitude, relative to other SLSNe, but is also slowly declining towards $+30$ days after peak. A direct comparison to \citet{inserra14} is difficult as our study uses different sample selection criteria (only 9 of their 16 SLSNe are included here because of our requirement to have three passbands per epoch), while \citet{inserra14} estimate a common synthetic peak magnitudes for all their SLSNe using their own time-evolving spectral template. A more detailed comparison will be possible with further DES SLSNe.
\subsection{Power source of DES13S2cmm}
\label{power}
The details of the power source of SLSN-I remain unclear and are much
debated (see Section~\ref{intro}). Here we investigate the two popular explanations in the literature for SLSNe-I: radioactive decay
of $^{56}$Ni, and energy deposition from a Magnetar.
\subsubsection{Radioactive $^{56}$Ni model}
\label{power_Ni}
As can been seen in Fig.~\ref{fig:Bol_lc}, the decline rate of
DES13S2cmm is approximately that expected from the decay of $^{56}$Co
(dashed line). This is suggestive of the SN energy source being the
production and subsequent decay of large quantities of $^{56}$Ni
produced in the explosion. To test this theory we fit the bolometric
light curve of DES13S2cmm to a model of $^{56}$Ni, following the
prescription of \citet{Arnett82}, which is the approximation of an
homologously expanding ejecta \citep{Chatzopoulos09,Inserra13} where
the luminosity at any given time (in erg s$^{-1}$) is given by
\begin{align}
L(t) &= M_{\textrm{Ni}} \, e^{ -(t / \tau_m)^2 } \biggl[ \epsilon_{\textrm{Ni}} \int_{0}^{t} \frac{2u}{\tau_m^2}\,e^{(u/\tau_m)^2}\,e^{-u/\tau_{\textrm{Ni}}} du \nonumber \\
& + \epsilon_{Co} \int_{0}^{t} \frac{2u}{\tau_m^2}\,e^{(u/\tau_m)^2}\biggl(e^{-u/\tau_{\textrm{Ni}}}-e^{-u/\tau_{\textrm{Co}}}\biggr) du \biggr] \delta_{\gamma},
\end{align}
where $u$ is the (time) integration variable. The energy production rate and decay rate for $^{56}$Ni are
$\epsilon_{\textrm{Ni}} =3.9\times10^{10}$ erg\,s$^{-1}$\,g$^{-1}$ and
$\tau_{\textrm{Ni}}=8.8$ days, while for $^{56}$Co these values are
$\epsilon_{\textrm{Co}} =6.8\times10^{9}$ erg\,s$^{-1}$\,g$^{-1}$ and
$\tau_{\textrm{Co}}=111.3$ days. The amount of $^{56}$Ni produced in
the explosion is $M_{\textrm{Ni}}$. Parameterizing the rise-time of the light curve,
$\tau_m$ (from Eqns 18 and 22 of \citealp{Arnett82}) is the geometric mean of the diffusion
and expansion timescales, and is given by
\begin{equation}
\tau_m = 1.05\, \Bigl( \frac{\kappa}{\beta c} \Bigr)^{1/2} \Bigl( \frac{M^3_{\textrm{ej}}}{E} \Bigr)^{1/4}
\label{eq:tau_m}
\end{equation}
Here $E$ is the explosion energy, $M_{\textrm{ej}}$ is the total amount of ejected mass, $\kappa$ is the optical opacity (assumed to be 0.1 cm$^{2}$ g$^{-1}$ and constant throughout, as in \citet{Inserra13}), and $\beta$ is a constant with value $\approx13.7$ \citep{Arnett82}.
We use $\delta_{\gamma}$ to denote the gamma-ray deposition function: the efficiency with which gamma-rays are trapped within the SN ejecta. For this function we follow \citet{Arnett82}, which is also used by \citet{Inserra13} and uses the deposition function defined in \citet{Colgate80},
\begin{equation}
\delta_{\gamma} = G[1+2G(1-G)(1-0.75G)],
\end{equation}
where $G\equiv\tau_{\gamma}/(\tau_{\gamma}+1.6)$, with the `optical depth' for gamma-rays approximately given by
\begin{equation}
\tau_{\gamma} \approx \Bigl( \frac{0.1}{\kappa} \Bigr) \Bigl( \frac{\tau_{m}^2}{4\tau_{\textrm{Ni}}^2} \Bigr) \Bigl( \frac{5.53 \times 10^{10}}{v_{ej}(0.1+t/\tau_{\textrm{Ni}})^2} \Bigr).
\end{equation}
Here $v_{ej} = \sqrt{10E_k/3M_{\textrm{ej}}}$ and is in units of cm s$^{-1}$. We note that the deposition function used in \citet{Chatzopoulos09} has a functional form that is similar to that used here, but in their approximation $\tau_{\gamma}$ only deviates from $\approx1$ at much later epochs, resulting in a light-curve decay rate mirroring $^{56}$Co and largely insensitive to changes in the explosion energy or ejecta mass.
We thus have four parameters in our model: the explosion epoch
($t_0$), the energy of the explosion ($E$), the total ejected mass
(M$_{\textrm{ej}}$), and the amount of ejected mass which is $^{56}$Ni
($M_{\textrm{Ni}}$).
We define the fraction of the ejecta mass that is in $^{56}$Ni as
$f_{\textrm{Ni}} = M_{\textrm{Ni}}/M_{\textrm{ej}}$.
We fit our model to the data with a variety of upper limits on $f_{\textrm{Ni}}$, varying
from 0.3 to 1.0.
However, we follow \citet{Inserra13} in assuming a physically motivated upper limit on $f_{\textrm{Ni}}$ of 0.5, which they base on the prevalence of intermediate-mass elements in their spectra of lower-$z$ SLSNe, and the calculations of $^{56}$Ni production in core-collapse SNe by \citet{Umeda2008}.
We report the best-fitting parameters and goodness
of fit for these models in Table~\ref{fitparams} as well as show the range of $f_{\textrm{Ni}}$ models in Figure 6 to illustrate the scale of uncertainties in the theoretical modelling.
\begin{table}
\begin{center}
\caption{Best-fitting parameters for a variety of $^{56}$Ni model
fits to the bolometric light curve for DES13S2cmm.
$f_{\textrm{Ni,max}}$ is the maximum fraction of $^{56}$Ni allowed
in the ejecta for each fit; for DES13S2cmm the best-fitting model
always maximizes this value. Ejecta and $^{56}$Ni masses are
given in units of solar masses. There are 22 degrees of freedom
for each fit.}
\begin{tabular}{ c | c | c | c | c | c }
\hline \hline
$f_{\textrm{Ni,max}}$ & $E(10^{51}$\,erg) & $M_{\textrm{ej}}$ & $M_{\textrm{Ni}}$ & $t_0$ (MJD) & $\chi^2$ \\
\hline
0.30 & 31.88 & 14.77 & 4.43 & 56508.8 & 81.5 \\
0.40 & 14.83 & 10.59 & 4.23 & 56510.4 & 76.0 \\
0.50 & 8.22 & 8.21 & 4.10 & 56511.5 & 71.5 \\
0.60 & 5.08 & 6.68 & 4.01 & 56512.4 & 67.9 \\
0.70 & 3.38 & 5.62 & 3.93 & 56513.1 & 65.0 \\
0.80 & 2.38 & 4.84 & 3.88 & 56513.7 & 62.7 \\
0.90 & 1.74 & 4.25 & 3.83 & 56514.2 & 60.9 \\
1.00 & 1.32 & 3.79 & 3.79 & 56514.6 & 59.4 \\
\hline
\end{tabular}
\label{fitparams}
\end{center}
\end{table}
\begin{figure}
\includegraphics[width=1.0\columnwidth,clip]{S2cmm_Bol_LC_v5_3.eps}
\caption{\textit{Upper panel:} Bolometric light curve of DES13S2cmm, with the best-fit magnetar model (green solid line) and three $^{56}$Ni models with different $f_{\rmn{Ni,max}}$ (blue lines) overplotted. The magnetar model is a better match to the data than the best-fit $^{56}$Ni, and has a significantly better $\chi^2$ than the models with a more realistic $f_{\rmn{Ni,max}}$ (see text). However both models have difficulty reproducing the peak luminosity, the post-peak decline, and the late-time flattening. \textit{Middle and Bottom panels:} Temperature and radius evolution of the best-fitting blackbody to each photometric epoch which is used to construct the bolometric light curve.
\label{fig:Models}}
\end{figure}
We find that the $^{56}$Ni model is not a good fit to the bolometric
light curve, as the best-fitting model has a $\chi^2/\rmn{dof}=2.7$
($\chi^2$ of 59 for 22 degrees of freedom), and is physically
unrealistic ($f_{\textrm{Ni}}=1$). As can be seen in
Table~\ref{fitparams}, the model prediction for $M_\rmn{Ni}$ is
relatively robust ($3.8-4.4 M_{\odot}$), as this parameter is
primarily constrained by the peak luminosity of the light curve.
However, to match the late time ($t>30$ days) flattening of the light
curve a large amount of ejecta mass or low explosion energy (see Eqn.~\ref{eq:tau_m}) is required to increase the diffusion time-scale, which simultaneously makes the fit to the post-peak decline poor while
predicting a much longer rise time than is seen. We also note that these parameters are poorly constrained individually as $E$ and $M_{\textrm{ej}}$ are correlated (Equation~\ref{eq:tau_m}), and this degeneracy cannot be broken without the presence of high-quality spectral data \citep{Mazzali2013}.
We do not provide uncertainties on the best fit parameters for each $f_{\textrm{Ni}}$ model in Table~\ref{fitparams}. This is because the formal uncertainties on parameters (such as $M_{\textrm{Ni}}$) are smaller for each fit than they are between fits with different constraints on $f_{\textrm{Ni}}$. This is due to the simplicity of the $^{56}$Ni model; we do not consider herein Ni mixing, non-standard density profiles, nor other possible variations in this basic model that could yield different light-curve shapes. For example, it has been noted that a two-component model that is denser in the inner component would produce a quicker flattening post-peak \citep{Maeda2003}.
\subsubsection{Magnetar model}
\label{power_mag}
As an alternative to the radioactive decay of $^{56}$Ni, we also fit
our derived bolometric light curve for DES13S2cmm to a magnetar model,
which proposes that SLSNe are powered by the rapid spinning-down of a
neutron star with an extreme magnetic field \citep{Woosley10,Kasen10}.
In this model the input energy is defined by the initial spin
period ($P_{ms}$, in milliseconds) and magnetic field strength
($B_{14}$, in units of $10^{14}$ Gauss) of the magnetar, and the
rise-time parameter ($\tau_m$, Eqn.~\ref{eq:tau_m}) reflects the
combined effect of the explosion energy ($E$), the ejected mass
($M_{\textrm{ej}}$), and the opacity ($\kappa$). We use the
semi-analytical model outlined in \citet{Inserra13}. The luminosity of the
magnetar model at time $t$ is given by (in erg s$^{-1}$)
\begin{align}
L(t) &= 4.9\times 10^{46}\,e^{ -(t / \tau_m)^2 }\delta \int_{0}^{t} \frac{2u}{\tau_m^2}\,e^{(u/\tau_m)^2}\,\frac{B_{14}^{2}\,P_{ms}^{-4}}{(1+u/\tau_{p})^2} du,
\end{align}
where $u$ is the (time) integration variable, $\tau_{p}$ is the spin-down timescale,
$\tau_{p}=4.7\,B_{14}^{-2}\,P_{ms}^{2}$ days, and $\delta$ is the
deposition function for the magnetar radiation; here we assume full
trapping (i.e., $\delta=1$) following \citet{Inserra13}. We do not include any additional
contribution from $^{56}$Ni decay in this model.
In Fig.~\ref{fig:Models} we include the best-fitting magnetar model to
the bolometric light curve of DES13S2cmm. The best-fitting parameters
of this model are a magnetic field of $1.43\times10^{14}$~Gauss, an
initial spin period of $5.28$~ms, a diffusion timescale of
$22.9$~days, and an explosion date of MJD 56511.0. This model has a
$\chi^2/\rmn{dof}=2.0$ ($\chi^2$ of 44 for 22 degrees of freedom) and is
therefore a significantly better fit to our data than the $^{56}$Ni
model. However, it is still clear that the model does not fit
\textit{well}: it is unable to reproduce the factor of two drop in
luminosity post-peak within 30 days of explosion (although it does a
better job than $^{56}$Ni). Intriguingly the late-time epochs seem to
fit the magnetar model well. Our time coverage of DES13S2cmm allows
us to determine whether either model is able to reproduce detailed
features of the light curve, but discerning between the two models
would be improved by an even longer time-scale over which the two
models cannot mimic one another; DES alone will unlikely provide
longer light curve coverage as its observing season is typically not much more than 5 months a year.
\section{Conclusions}
\label{sec:conclusions}
In this paper we have presented DES13S2cmm, the first
spectroscopically-confirmed superluminous supernova (SLSN) from the
Dark Energy Survey. Using spectroscopic data obtained from the
ESO Very Large Telescope, we measured a redshift of $z=0.663\pm0.001$,
and assigned a classification of SLSN-I. However, we cannot exclude
the possibility that DES13S2cmm is a type R (Fig.~\ref{fig:spectra} and \ref{fig:Bol_lc})
if this is in fact a distinct SLSN type.
Using this redshift and correcting for Milky Way extinction, we find
the rest-frame $U$-band absolute magnitude of DES13S2cmm at peak to be
$M^{peak}_U=-21.05^{+0.10}_{-0.09}$, consistent with the SLSNe
definition of \citet{Gal-Yam12}. Like other SLSNe \citep{lunnan14}, DES13S2cmm is
located in a faint low stellar-mass host galaxy
($\log(\rmn{M}/\rmn{M}_{\odot})=8.9\pm0.3$) with sub-solar metallicity.
In Fig.~\ref{fig:Bol_lc}, we compare the bolometric light curve of
DES13S12cmm to fourteen similarly well-observed SLSNe-I light curves in literature, and see that DES13S2cmm has one of the slowest declining tails (beyond $+30$ rest frame past peak) of all SLSNe-I studied herein, as well as likely being the faintest at peak. We further find that the dispersion between the bolometric light curves of all SLSNe-I shown in Fig.~\ref{fig:Bol_lc} has a minimum of $0.29$ magnitudes around $+30$ days past peak (rest frame). This reduces to $0.20$ magnitudes around $+25$ days if we remove the four SLSNe-I above $z\simeq1$ from the measurement because of the increased uncertainty in their rest-frame UV luminosities. This observation raises the tantalising possibility of
`standardising' these SLSNe like SNe Ia \citep{Quimby11,inserra14} and starting a new era of supernova cosmology \citep{king14}. Further study is required to confirm if SLSNe can be standardised and investigate further the k-corrections of SLSNe, which would be needed to use SLSNe as robust distance indicators.
In Fig.~\ref{fig:Models}, we fit the bolometric light curve of
DES13S2cmm with two possible models for the power source of these
extreme events; the radioactive decay of $^{56}$Ni and a
magnetar \citep[e.g. see ][]{Inserra13}.
We find that the $^{56}$Ni model is not a good fit to the bolometric
light curve, as the best-fitting model has a $\chi^2/$DoF$=2.7$ with an unrealistic fraction of $^{56}$Ni to ejected mass ($f_{\textrm{Ni}}=1$). The model does provide a
relatively robust prediction for the overall $^{56}$Ni mass ($3.8-4.4 M_{\odot}$), as this is primarily
constrained by the peak luminosity of the light curve, but is unable to reproduce the late time ($t>30$ days) evolution of the light curve. The Magnetar model provides a smaller $\chi^2$ (for the same degrees of freedom) than $^{56}$Ni, but is likewise not a good fit to the data as
it is unable to fully reproduce the drop in luminosity within 30 days of explosion.
In the future, DES should find many more SLSN-I. Such data should allow us to further test
these two models, especially if we can measure the (rest frame)
light curves to $>100$ days (see Fig.~\ref{fig:Models} and Inserra et
al. 2013). Therefore, we have begun a new project entitled `SUrvey with Decam for Superluminous
Supernovae' (SUDSS), which will enable extended monitoring of some
DES SN fields (in addition to several non-DES fields) to over 8 months, greatly enhancing both
our ability to detect many more SLSNe and measure their long light curves.
\section*{Acknowledgments}
The authors wish to thank Kate Maguire for her assistance with the ESO
VLT Directors Discretionary Time proposal. We also thank Cosimo Inserra and Stephen Smartt for helpful discussions regarding the classification, standardisation and k--corrections of superluminous supernovae. AP acknowledges the
financial support of SEPnet (www.sepnet.ac.uk) and the Faculty of
Technology of the University of Portsmouth. Likewise, CD and RN thank
the support of the Faculty of Technology of the University of
Portsmouth during this research, and MS acknowledges support from the
Royal Society and EU/FP7-ERC grant no [615929]. Part of TE's research was
carried out at JPL/Caltech, under a contract with NASA
Based on observations made with ESO telescopes at the La Silla Paranal Observatory under DDT program ID 292.D-5013
We are grateful for the extraordinary contributions of our CTIO
colleagues and the DES Camera, Commissioning and Science Verification
teams in achieving the excellent instrument and telescope conditions
that have made this work possible. The success of this project also
relies critically on the expertise and dedication of the DES Data
Management organization. Funding for the DES Projects has been
provided by the U.S. Department of Energy, the U.S. National Science
Foundation, the Ministry of Science and Education of Spain, the
Science and Technology Facilities Council of the United Kingdom, the
Higher Education Funding Council for England, the National Center for
Supercomputing Applications at the University of Illinois at
Urbana-Champaign, the Kavli Institute of Cosmological Physics at the
University of Chicago, Financiadora de Estudos e Projetos, Funda{\c
c}{\~a}o Carlos Chagas Filho de Amparo {\`a} Pesquisa do Estado do
Rio de Janeiro, Conselho Nacional de Desenvolvimento Cient{\'i}fico e
Tecnol{\'o}gico and the Minist{\'e}rio da Ci{\^e}ncia e Tecnologia,
the Deutsche Forschungsgemeinschaft and the Collaborating Institutions
in the Dark Energy Survey.
The Collaborating Institutions are Argonne National Laboratory, the
University of California at Santa Cruz, the University of Cambridge,
Centro de Investigaciones Energeticas, Medioambientales y
Tecnologicas-Madrid, the University of Chicago, University College
London, the DES-Brazil Consortium, the Eidgen{\"o}ssische Technische
Hochschule (ETH) Z{\"u}rich, Fermi National Accelerator Laboratory,
the University of Edinburgh, the University of Illinois at
Urbana-Champaign, the Institut de Ciencies de l'Espai (IEEC/CSIC), the
Institut de Fisica d'Altes Energies, Lawrence Berkeley National
Laboratory, the Ludwig-Maximilians Universit{\"a}t and the associated
Excellence Cluster Universe, the University of Michigan, the National
Optical Astronomy Observatory, the University of Nottingham, The Ohio
State University, the University of Pennsylvania, the University of
Portsmouth, SLAC National Accelerator Laboratory, Stanford University,
the University of Sussex, and Texas A\&M University.
This paper has gone through internal review by the DES collaboration. Please contact the author(s) to request access to research materials discussed in this paper.
\bibliographystyle{mn2e}
|
\section{Introduction}
\subsection{} \label{sec:N} Let $k$ be $p$-adic field with ring of
integers $\foo_k$, prime ideal $\fpp_k = (\varpi)$ and residue field~$\fff$ of odd characteristic~$p$ where $\varpi$ is a fixed
uniformizer. Let $\bar{k}$ be its algebraic closure and let $\kur$ be the maximal unramified extension of $k$. Let $\val \colon \bar{k} \to \bQ \cup \set{\infty}$ denote the
valuation map where $\val(\varpi) = 1$.
We will denote an algebraic variety by a boldface letter, say $\bfH$. For an algebraic extension $E$ of $k$, we will denote its
$E$-points by $\bfH(E)$, and its $k$ points by the corresponding normal letter
$H$.
If $\bfG$ is an algebraic group, then we let $\fgg$ be its Lie algebra
or the $k$ points of the Lie algebra, depending on the context. If
$G = \bfG(k)$ acts on a set $X$, then ${}^g x$ will denote $g\cdot x$
for $g\in G$ and $x\in X$. In order to simplify the situation, we
always assume that $p$ is sufficiently large compared to the rank
of $G$. For an reductive algebraic group, we will fix a maximally $k$-split
torus $\bfS$, a maximally $\kur$-split torus $\bfT$ which is defined
over $k$ and contains $\bfS$. Since $\bfG(\kur)$ is always
quasi-split, we set $\bfY := \Cent_\bfG(\bfT)$ to be the fixed Cartan
subgroup of $\bfG$
\subsection{}
We recall the classification of irreducible type~I reductive dual pairs. Let
$D$ be a division algebra over $k$ with a fixed involution $\tau$, which is
either \begin{inparaenum}[(i)]
\item $k$,
\item a
quadratic field extension of $k$, or
\item the quaternion algebra over $k$.
\end{inparaenum}
We continue to use $\val$ to denote the unique extension of the
valuation $\val$ from $k$ to $D$. Let $\foo_D$ be the ring
of integers of $D$. Let $V$ be a right $D$-module with an
$\epsilon$-Hermitian sesquilinear form $\inn{}{}_V$ and
$G = \bfG(k) = \rU(V, \inn{}{}_V)$ be the unitary group preserving
$\inn{}{}_V$. Similar notation applies to $G'$ with $\epsilon' = -\epsilon$.
The $k$-vector space $W = V \otimes_D V'$
has a natural symplectic form and $(G,G')$ is an irreducible type~I
reductive dual pair in $\Sp:=\Sp(W)$.
We will let $\widetilde{E}$ be the inverse image of a subgroup $E$ in the
metaplectic $\bC^\times$-cover $\Mp$ of $\Sp$.
We fix a non-trivial additive character $\psi \colon k \rightarrow \bC^\times$ with conductor $\fpp_k$
and consider local theta correspondence $\theta$ arising from the
oscillator representation of $\Mp$ with respect to the
character $\psi$.
For general information on theta correspondences, see \cites{Ho,Wa,GT}.
In this paper, we will investigate the theta correspondence between
epipelagic supercuspidal representations of $\tG$ and $\tG'$. In many
situations, the roles of $G$ and $G'$ are interchangeable. In such
cases, we only discuss $G$ and extend all objects and
notation to $G'$ implicitly by adding `primes'.
\subsection{} \label{sec:SD}
We briefly review some facts about epipelagic supercuspidal
representations. See \Cref{sec:Epi} and \cite{RY} for more
details. Let $\cB(\bfG,k)$ be the (extended) building of $G$. In the
setting of type~I dual pairs, $G$ and $G'$ have compact
centers. Therefore $\Cent(G)\subseteq G_{x}$ and $G_x = G_{[x]}$ for
any $x\in \cB(\bfG,k)$ where $[x]$ is the image of $x$ in the reduced
building. Following Reeder-Yu~\cite{RY}, we construct a tamely
ramified irreducible supercuspidal representation $\pi_\Sigma$ of
$\tG$ in \Cref{sec:CEpi} from the data $\Sigma = (x,\lambda,\chi)$
where \crefname{SD}{SD}{SD} \crefformat{SD}{(SD#2#1#3)}
\begin{enumerate}[(SD1)]
\item \label[SD]{SD1} $x\in \cB(\bfG,k)$ is an epipelagic point of order $m$
where $p \nmid m$ (c.f. \Cref{def:epipoint}),
\item \label[SD]{SD2}$\lambda$ is a stable vector in
$\fgg_{x,-\fracmm}/\fgg_{x,-\fracmm^+}$ and
\item \label[SD]{SD3} $\chi$ is a character of the stabilizer
$\sfS_\lambda$ of $\lambda$ in $G_x/G_{x,0^+}$.
\end{enumerate}
We call $\Sigma = (x,\lambda,\chi)$ an {\it epipelagic data} of order
$m$ for $G$. The supercuspidal representation~$\pi_\Sigma$ has depth
$\frac{1}{m}$ and it is called an {\it epipelagic supercuspidal
representation} of $\tG$. Since the cover $\tG$ depends on the dual
pair $(G,G')$, the notation $\pi_\Sigma$ only makes sense relative to
the dual pair $(G,G')$.
\subsection{} \label{S14}
Let $\Sigma = (x, \lambda, \chi)$ and
$\Sigma' = (x', \lambda', \chi')$ be epipelagic data of $G$ and $G'$
of order $m$ and $m'$ respectively. Suppose
$\theta(\pi_\Sigma) = \pi'_{\Sigma'}$. By \cite{Pan,Pan02}, $\piSigma$ and
$\piSigmap$ have the same depth $\frac{1}{m} = \frac{1}{m'}$, i.e.
$m = m'$. It turns out that the data $(x,\lambda,\chi)$ and
$(x',\lambda',\chi')$ are related by a geometric picture which we now
briefly explain.
\def\sfMm{{\sfM_{-\fracdmm}}}
\def\sfMmp{{\sfM'_{-\fracdmm}}}
The points $x$ and $x'$ correspond to
self dual $\foo_D$-lattice functions $\sL$ and $\sL'$ in $V$ and $V'$
respectively (c.f. \cite{BT4,BS,Le,GY}). The tensor product $\sB = \sL \otimes \sL'$ is a self dual
$\foo_k$-lattice
function in $W$. The quotient $\sfX = \sB_{-\frac{1}{2m}}/ \sB_{-\frac{1}{2m}^+}$ is an $\fff$-vector
space. In \eqref{eq:MM.s} we define the moment maps $\sfMm$ and $\sfMmp$:
\[
\xymatrix@C=3em{
\fgg_{x,-\fracmm} & \ar[l]_<>(.5){\sfMm} \sfX \ar[r]^<>(.5){\sfMmp} &
\fgg'_{x',-\fracmm}.
}
\]
Our first result is a refinement of special cases of \cite{Pan}.
\begin{prop}[\Cref{prop:M.w}]
Suppose $\theta(\piSigma) = \piSigmap$. Then
\begin{equation} \label[cnd]{cnd:M}\tag{M}
\parbox{0.8\textwidth}{
there exists a $\bar{w} \in \sfX$ such that $\lambda = \sfMm(\bw)$ and
$\lambda' = - \sfMmp(\bw)$ }.
\end{equation}
\end{prop}
\Cref{cnd:M} imposes severe restriction to
the ranks of $G$ and $G'$:
\begin{prop}[\Cref{prop:RSP}] \label{prop:I.RSP} Suppose
$\theta(\pi_\Sigma) =
\pi'_{\Sigma'}$.
Then $(\bfG,\bfG')$ or $(\bfG',\bfG)$ is one of the following types:
\begin{inparaenum}[(i)]
\item $(\rD_n,\rC_n)$,
\item $(\rC_n,\rD_{n+1})$,
\item $(\rC_n,\rB_n)$,
\item $(\rA_n,\rA_n)$ or
\item $(\rA_n,\rA_{n+1})$.
\end{inparaenum}
\end{prop}
Let $(x,\lambda)$ and $(x',\lambda')$ be parts of data for $G$ and $G'$
of order $m$ respectively satisfying (SD1) and (SD2). It turns
out that Condition~\eqref{cnd:M}, in all but one exceptional case
(see \Cref{case:E} in \Cref{sec:thetaII}), is a sufficient
condition for the epipelagic supercuspidal representation $\piSigma$
to lift to an epipelagic supercuspidal representation of $\tG'$.
For the ease of explaining in this introduction, we will omit the exceptional case.
Using \cref{cnd:M}, we will construct a group homomorphism
$\alpha \colon \sfS_{\lambda'}' \rightarrow \sfS_\lambda$ in
\Cref{lem:Xll}. We can now state a part of the main \Cref{thm:main0}.
\delete{\begin{thm} \label{thm:intro}
Let $(G,G')$ be an irreducible type~I reductive dual pair such that
$(\bfG,\bfG')$ has the form (i)-(v) in
\Cref{prop:I.RSP}. Let $(x,\lambda)$ and $(x',\lambda')$ be data of
$G$ and $G'$ respectively of order $m$ satisfying \cref{SD1},
\cref{SD2} and \cref{cnd:M}. Suppose we are not in the
exceptional Case~\eqref{case:E}. Then for every character
$\chi$ of $\sfSl$,
\[
\theta( \piSigma) = \piSigmap
\]
where $\Sigma = (x,\lambda,\chi)$,
$\Sigma' = (x',\lambda',\chi^*\circ \alpha)$ and $\chi^*$ is the
contragredient of $\chi$. In particular the theta lift is nonzero.
\end{thm}
We state a converse of the above theorem.
\begin{cor} \label{cor:Converse}
Suppose $\theta(\piSigma) = \piSigmap$ where $\Sigma = (x,\lambda, \chi)$ and $\Sigma' = (x', \lambda',\chi')$ are epipelagic data, and $(G,G')$ and $\Sigma$ do not belong to the exceptional case (E). Then
\begin{enumerate}[(i)]
\item Condition (M) is satisfied so that $\alpha \colon \sfS_{\lambda'}' \rightarrow \sfS_\lambda$ is well-defined and
\item $\chi' = \chi^* \circ \alpha$.
\end{enumerate}
\end{cor}}
\begin{thm} \label{thm:intro}
Let $(G,G')$ be an irreducible type~I reductive dual pair such that
$(\bfG,\bfG')$ has the form (i)-(v) in
\Cref{prop:I.RSP}. Let $(x,\lambda)$ and $(x',\lambda')$ be data of
$G$ and $G'$ respectively of order $m$ satisfying \cref{SD1},
\cref{SD2}. We assume that we are not in the
exceptional \Cref{case:E}
\begin{enumerate}[(i)]
\item
Suppose \cref{cnd:M} is satisfied. Then for every character
$\chi$ of $\sfSl$,
\[
\theta( \piSigma) = \piSigmap
\]
where $\Sigma = (x,\lambda,\chi)$,
$\Sigma' = (x',\lambda',\chi^*\circ \alpha)$ and $\chi^*$ is the
contragredient of $\chi$. In particular the theta lift is nonzero.
\item Conversely, suppose $\theta(\piSigma) = \piSigmap$ where $\Sigma = (x,\lambda, \chi)$ and $\Sigma' = (x', \lambda',\chi')$ are epipelagic data. Then
\begin{enumerate}[(a)]
\item Condition (M) is satisfied so that $\alpha \colon \sfS_{\lambda'}' \rightarrow \sfS_\lambda$ is well-defined and
\item $\chi' = \chi^* \circ \alpha$.
\end{enumerate}
\end{enumerate}
\end{thm}
\Cref{thm:main0} also contains a result for the exceptional
Case~\eqref{case:E} where $\pi_\Sigma$ lifts only for half of the
characters of~$\sfSl$. This should be compared with \cite{Moen93}.
\subsection*{Acknowledgment}
We would like to thank Wee Teck Gan and Jiu-Kang Yu for their valuable
comments. Hung Yean Loke is supported by a MOE-NUS AcRF Tier 1 grant
R-146-000-208-112. Jia-Jun Ma is partially supported by ISF Grant 1138/10 during his postdoctoral Fellowship at Ben Gurion University and HKRGC Grant CUHK 405213 during his postdoctoral fellowship in IMS of CUHK. Gordan Savin is supported by an NSF grant DMS-1359774.
\section{Epipelagic representations} \label{sec:Epi}
In this section we review Reeder-Yu's construction of epipelagic supercuspidal
representations (c.f. \cite{RY}).
\subsection{} Let $k$ be a $p$-adic field as in \Cref{sec:N}.
Let $\kur$ be the
maximal unramified extension of $k$ with residue field $\bfff$.
We let $\Fr$ denote the Frobenius element such that
$\Gal(\kur/k) = \Gal(\bfff/\fff) = \langle \Fr \rangle$.
\subsection{Epipelagic points}\label{sec:EpiP} Let $\bfG$ be an algebraic
group defined over $k$. Let $E$ be a tamely ramified extension of $k$.
For $x\in \cB(\bfG,k) \subseteq \cB(\bfG,E)$, we set up some notation which will be used
in the rest of the paper.
Let
\begin{itemize}
\item
$\bfG(E)_{x,r}$ ($r\geq 0$) and $\fgg(E)_{x,r}$ be the Moy-Prasad filtrations corresponding to $x$.
\item $\bfG(E)_x = \Stab_{\bfG(E)}(x)$ and $\bfG(E)_{[x]} =
\Stab_{\bfG(E)}([x])$ where $[x]$ is the image of $x$ in the reduced
building;
\item $\sfG_x(E) = \bfG(E)_x/\bfG(E)_{x,0^+}$;
\item
$\sfG_{x,r}(E) = \bfG(E)_{x,r:r^+} := \bfG(E)_{x,r}/\bfG(E)_{x,r^+}$, and
\item $\sfg_{x,r}(E) = \fgg(E)_{x,r:r^+} := \fgg(E)_{x,r}/\fgg(E)_{x,r^+}$.
\end{itemize}
In order to abbreviate the notations of objects corresponding to
$G = \bfG(k)$, we let
\begin{itemize}
\item $G_{x,r} := \bfG(k)_{x,r} = \bfG(\kur)_{x,r} \cap \bfG(k)$,
\item $\fgg_{x,r} := \fgg(k)_{x,r} = \fgg(\kur)_{x,r} \cap \fgg(k)$,
\item {$G_{x,r:r'} := \bfG(k)_{x,r}/\bfG(k)_{x,r'}$ and $\fgg_{x,r:r'} := \fgg(k)_{x,r}/\fgg(k)_{x,r'}$ for $r < r'$.}
\end{itemize}
In order to abbreviate the notations of objects corresponding to
$\bfG(\kur)$, we let
\begin{itemize}
\item $\sfG_x := \sfG_x(\kur)$, $\sfG_{x,r} := \sfG_{x,r}(\kur)$ and
$\sfg_{x,r} := \sfg_{x,r}(\kur)$.
\end{itemize}
The quotient space $\sfg_{x,r}$ is an $\bfff$-vector space and we
denote its dual space $\Hom_{\bfff}(\sfg_{x,r}, \bfff)$ by
$\csfg_{x,r}$. We have assumed in the introduction that $p$ is large
compared to the rank of $G$. Then by \cite[Prop. 4.1]{AR},
$\csfg_{x,r}$ could be identified with $\sfg_{x,-r}$ via an invariant
bilinear form on $\fgg$. Since we are only treating classical groups
and $p \neq 2$, we will use a trace form defined later in
\Cref{def:Parings} in this paper.
The group $\sfG_x$ acts on $\csfg_{x,r}$.
A vector $\lambda \in \csfg_{x,r}$ is called a \emph{stable vector} if the $\sfG_x$-orbit of $\lambda$ is Zariski closed in $\csfg_{x,r}$ and the stabilizer of $\lambda$ in $\sfG_x$ modulo $\Cent(\bfG(\kur))_{0:0^+}$ is a finite group.
\medskip
Let {$\Psi_{\kur}$} be the set of affine $\bfT(\kur)$-roots. Suppose
$x \in \cA(\bfS, k) = \cA(\bfT,\kur)^\Fr$, i.e. $x$ is in the apartment
defined by $\bfS$. Let $r(x)$ be the smallest positive value in
$\set{ \psi(x) | \psi \in \Psi_{\kur} }$. Then
$\bfG(\kur)_{x,0^+} = \bfG(\kur)_{x,r(x)}$.
\begin{definition} \label{def:epipoint} Let $m$ be an integer where
$p \nmid m$. A point $x$ in the apartment
$\cA(\bfS,k)$ is called an {\it epipelagic point} of order $m$ if
$r(x) = \frac{1}{m}$ and $\csfg_{x,\fracmm}$ contains a stable vector.
\end{definition}
By \cite{Gross}, $m$ is an even integer except when $\bfG(\kur)$ is split
and of type $\rA_{m-1}$. In particular~$m \geq~2$.
\def\rx{{r(x)}}
\subsection{Epipelagic supercuspidal representations} \label{sec:Epirep}
Let $x$ be an epipelagic point
of order $m$ so that $r(x) = \fracmm$. We fix an isomorphism of
abelian groups $c \colon G_{x,\rx:\rx^+} \rightarrow \fgg_{x,\rx:\rx^+}$.
For the classical group, we choose the isomorphism to be the one
induced by the Cayley transform $c(g) = 2(g-1)(g+1)^{-1}$ so that $g-1 \equiv
c(g) \pmod{\fgl(V)_{x,r(x)^+}}$.
\def\Jx{{G_{x,\fracmm}}}
\def\Jxp{{G_{x,\fracmm^+}}}
Let $\lambda \in \csfg_{x,\fracmm}^{\Fr}$ be an $\fff$-rational
functional on $\sfg_{x,\fracmm}$. Then we get a character
\[
\psi_\lambda := \psi \circ \lambda \colon
G_{x,\fracmm:\fracmm^+ } = \sfg_{x,\fracmm}(k) \cong \sfg_{x,\fracmm}^\Fr
\xrightarrow{\ \lambda\ } \fff \xrightarrow{\ \psi\ }
\bC^\times.
\]
The inflation of $\psi_{\lambda}$ to $\Jx$ will also be denoted by
$\psi_{\lambda}$.
Let
\begin{equation} \label{eq1}
H_{x,\lambda} :=\Stab_{G_{[x]}}(\lambda), \ \ \sfA_{x,\lambda} := H_{x,\lambda}/\Jx.
\end{equation}
Then $H_{x,\lambda}$ is the stabilizer of $\psi_\lambda$ in~$G_{[x]}$.
\medskip
We will assume that $\lambda$ is $\sfG_x$-stable in
$\csfg_{x,\frac{1}{m}}$. In all the cases that we will consider in
this paper, $\sfA_{x,\lambda}$ is a finite abelian group.
Note that the order of $\sfS_{x,\lambda}$ is prime to $p$ and $\Jx$ is pro-$p$.
\begin{prop} The group $\sfA_{x,\lambda}$ splits in $H_{x,\lambda}$, i.e.
$H_{x,\lambda} = \sfA_{x,\lambda} \ltimes \Jx$.
\end{prop}
\begin{proof} We shall show that $\sfA_{x,\lambda}$ splits in $H_{x,\lambda}/G_{x,\frac{i}{m}}$ for all $i \geq 1$. For $i=1$ there is nothing to prove.
Assume that we have constructed a splitting $s_i \colon \sfA_{x,\lambda} \rightarrow H_{x,\lambda}/G_{x,\frac{i}{m}}$. The obstruction to lift this splitting to
$H_{x,\lambda}/G_{x,\frac{i+1}{m}}$ lies in $H^2(\sfS_{x,\lambda},G_{x,\frac{i}{m}:\frac{i+1}{m}})$. Since the order of $\sfS_{x,\lambda}$ is prime to $p$ and
$G_{x,\frac{i}{m}:\frac{i+1}{m}}$ is an elementary $p$-group, this cohomology vanishes. Hence $s_i$ can be lifted to $s_{i+1}$ and the proposition follows
by passing to a limit.
\end{proof}
We extend
$\psi_\lambda$ to a character of $H_{x,\lambda}$ by setting
$\psi_\lambda$ to be trivial on $\sfS_{x,\lambda}$. By
$H^1(\sfS_{x,\lambda},\Jx) = 0$, we know that all splittings are
conjugate up to $\Jx$-conjugation. Hence the extension $\psi_\lambda$
is unique and therefore canonical.
Let $\chi$ be a character of $\sfA_{x,\lambda}$ and let
$\pi_x(\lambda, \chi) := \ind_{H_{x,\lambda}}^G \psi_{\lambda} \otimes
\chi$. By \cite[Prop.~5.2]{RY}, $\pi_x(\lambda,
\chi)$ is an irreducible supercuspidal representation of $G$.
We will call $(x, \lambda, \chi)$ an {\it epipelagic data} of
order $m$ and we call $\pi_x(\lambda, \chi)$ an (irreducible) {\it
epipelagic supercuspidal representation} attached to the data. It
contains a minimal $K$-type represented by a coset
$\lambda = [\Gamma] = \Gamma + \fgg_{x,-\fracmm^+}$ in
$\fgg_{x,-\fracmm:-\fracmm^+}$ where $K = G_{x,\fracmm}$.
\begin{prop} \label{prop:MKT}
Suppose $G$ is a group appearing in a type I reductive dual pair.
\begin{asparaenum}[(i)]
\item
All unrefined minimal $K$-types of
$\pi_x(\lambda,\chi)$ are $G$-conjugate to
$\Gamma + \fgg_{x,-\fracmm^+}$.
\item
If $\pi_x(\lambda, \chi)$ and $\pi_x(\lambda',\chi')$ are isomorphic
$G$-modules, then $\lambda$ and $\lambda'$ are in
the same $G_x$-orbit. In addition if $\lambda = \lambda'$ then
$\chi = \chi'$.
\end{asparaenum}
\end{prop}
\begin{proof}
(i) Let $\Gamma + \fgg_{x,-\frac{1}{m}^+}$ represent the unrefined
minimal $K$-type of $\pi$ as above. We will see in \Cref{lem:reg}
later that $\Gamma \in \fgg_{x,-\fracmm}$ is a good element and
$\bfH =\Cent_\bfG(\Gamma)$ is a torus. Let $H = \bfH(k)$ and
let $\fhh = \Cent_{\fgg}(\Gamma)$ be its Lie algebra. Since
$(x,\Gamma,\chi)$ is a tamely ramified supercuspidal data,
$H/\Cent(G)$ is $k$-anisotropic (or see for example,
\cite[Proposition~14.5]{Kim}). We are considering type I classical
dual pairs so $\Cent(G)$ is anisotropic. Therefore $H$ is
$k$-anisotropic. Hence $\cB(\bfH,k) = \set{x}$.
Suppose $\Gamma_y + \fgg_{y,-\fracmm^+}$ is another unrefined
minimal $K$-type of $\pi$ for some $y \in \cB(\bfG,k)$.
Now we show that we can move $y$ to $x$. Since $\Gamma
+ \fgg_{x,-\frac{1}{m}^+}$ and $\Gamma_y + \fgg_{y,-\frac{1}{m}^+}$
are minimal $K$-types of $\pi$, they are associates \cite{MP}, i.e. there
exists a $g \in G$ such that
\begin{equation} \label{eq3}
{}^g\left( \Gamma_y + \fgg_{y,-\fracmm^+} \right) \cap \left( \Gamma
+\fgg_{x,-\fracmm^+} \right) = \left({}^g\Gamma_{y}
+\fgg_{gy,-\fracmm^+} \right) \cap \left( \Gamma
+\fgg_{x,-\fracmm^+} \right)
\end{equation}
is nonempty.
Now we apply the argument in the proof of \cite[Corollary~2.4.8]{KM}. By
\eqref{eq3}, there are $X''\in \fgg_{x,-\fracmm^+}$ and
$Y''\in \fgg_{gy,-\fracmm^+}$ such that ${}^g\Gamma_{y}+ Y'' = \Gamma + X'' $.
By \cite[Corollary 2.3.5]{KM}, there exists $h\in G_{x,0^+}$ such that
${}^h(\Gamma+X'') = \Gamma+ X'\in (\Gamma+\fgg_{x,-\fracmm^+})\cap \fhh \cap
\fgg_{hgy,-\fracmm}$
where $X' \in \fhh_{x,-\frac{1}{m}^+}$. By \cite[Lemma~2.4.7]{KM}
$hgy \in \cB(\bfH,k) = \set{x}$, i.e. $hgy = x$. We consider the isomorphism
$\Ad((hg)^{-1}) \colon \fgg_{y,-\fracmm}/\fgg_{y,-\fracmm^+} \iso
\fgg_{x,-\fracmm}/\fgg_{x,-\fracmm^+}$.
It is now clear that the coset $[{}^{hg} \Gamma_y] = [\Gamma] = \lambda$. This
proves (i).
(ii) The last assertion of (ii) is \cite[Lemma 2.2]{RY}. Now we
prove the first assertion. Note that
$\lambda = \Gamma + \fgg_{x,-\frac{1}{m}^+}$ and
$\lambda' = \Gamma' + \fgg_{x,-\frac{1}{m}^+}$ represent unrefined
minimal $K$-types of $\pi_x(\lambda, \chi)$ and
$\pi_x(\lambda', \chi')$ respectively. Since
$\pi_x(\lambda, \chi) \cong \pi_x(\lambda', \chi')$, the two minimal
$K$-types are associates. By the proof in (i) where $y = x$ and
$\Gamma_y = \Gamma'$, we conclude that there exists $g \in G$ such
that $gx = x$ and $g \lambda' = \lambda$. In particular $g \in G_x$.
Hence $\lambda$ and $\lambda'$ are in the same $G_x$-orbit in
$\sfg_{x,-\fracmm}(k)$.
\end{proof}
\section{Classical reductive dual pairs and local theta
correspondence}
\subsection{Classical groups} \label{SS21} In this section, we will
define the classical groups which appear in the irreducible dual
pairs.
Let $D$ be a division algebra over $k$ with an involution $\tau$ in
one of the following cases:
\begin{enumerate}[(i)]
\item \label{D.k}$D = k$, $\tau$ is the identity map on $k$ and $\varpi_D = \varpi$.
\item \label{D.qd}$D$ is a quadratic extension of $k$, $\tau$ is the nontrivial
Galois element in $\Gal(D/k)$, $\varpi_D$ is a uniformizer of $D$ so
that $\varpi_D = \varpi$ if $D/k$ is unramified or
$\tau(\varpi_D) = -\varpi_D$ if it is ramified.
\item \label{D.qt}$D$ is the quaternion algebra over $k$, $\tau$ is the
usual involution on $D$, $\varpi_D$ is a uniformizer of $D$ such
that $\varpi_D^2 = \varpi$.
\end{enumerate}
Let $\foo_D$ denote the ring of integers of $D$,
$\fpp_D = \varpi_D\foo_D$ denote its maximal prime ideal and
$\fff_D = \foo_D/\fpp_D$ denote its residue field. We set
$\val_D := \val(\varpi_D)$.
\def\rhalf{{\half \val_D}}
\def\iotag{{\iota_{\sfggE}}}
Let $V$ be a right $D$-vector space. Let $\End_D(V)$ denote the space
of $D$-linear endomorphisms of $V$ which acts on the left. For
$\epsilon = \pm 1$, let $\inn{}{}_V \colon V\times V\rightarrow D$ be
an $\epsilon$-Hermitian sesquilinear form, i.e.
\[
\inn{v_1}{v_2}_V = \epsilon\inn{v_2}{v_1}_V^\tau \quad\text{and}\quad
\inn{v_1a_1}{v_2a_2}_V = a_1^\tau\inn{v_1}{v_2}_V a_2
\]
for all $v_1,v_2 \in V$ and $a_1,a_2 \in D$. The $\epsilon$-Hermitian
form induces a conjugation $*\colon \End_D(V)\rightarrow \End_D(V)$
such that $\inn{gv_1}{v_2}_V = \inn{v_1}{g^* v_2}_V$ for all
$v_1, v_2 \in V$ and $g \in \End_D(V)$. Then
\begin{align*}
G & = \rU(V) = \rU(V, \inn{}{}_V) := \set{ g \in \End_D(V) | gg^* = \Id }
\quad \text{and} \\
\fgg & = \fuu(V) = \fuu(V, \inn{}{}_V)
:= \set{ X \in \End_D(V) | X + X^* = 0}
\end{align*}
are a classical group and its Lie algebra.
\subsection{Irreducible reductive dual pairs of
type~I} \label{sec:dualpair} Let $V$ be a right $D$-vector space
equipped with an $\epsilon$-Hermitian sesquilinear form $\inn{}{}_V$
and let $V'$ be a right $D$-vector space equipped with an
$\epsilon'$-Hermitian sesquilinear form $\inn{}{}_{V'}$ where $\epsilon' = - \epsilon$. Let $G$ and
$G'$ be the classical groups defined by $(V,\inn{}{}_V)$ and
$(V',\inn{}{}_{V'})$ respectively.
We view $V'$ as a left $D$-module by $a v = v a^\tau$ for all
$a \in D$ and $v \in V'$. Let $W = V \otimes_D V'$. It is a
symplectic $k$-vector space with symplectic form $\inn{}{}$ given by
\begin{equation} \label{equ4}
\inn{v_1 \otimes v_1'}{v_2 \otimes v_2'} =
\tr_{D/k}(\inn{v_1}{v_2}_V \inn{v_1'}{v_2'}_{V'}^{\tau}).
\end{equation}
Then $G$ and $G'$ commute with each other in the
symplectic group $\Sp(W)$. We call $(G,G')$ an \emph{irreducible reductive
dual pair of type~I}.
\subsection{Lattice model} \label{sec:LM} We recall that
$\psi \colon k \rightarrow \bC^\times$ is a non-trivial additive character
with conductor $\fpp_k$. Let $A$ be a self dual lattice in $W$, i.e. $A = \set{w \in W | \inn{w}{w'} \in \fpp_k,
\forall w' \in A }$. The lattice model with respect to~$A$ of the
oscillator representation $\omega$ with respect to the character $\psi$ is
defined by
\[
\sS(A) = \Set{f\colon W \to \bC|\begin{array}{c} f(a+w) =
\psi(\frac{1}{2}\inn{w}{a})f(w)\ \forall a\in A\\
\text{$f$ locally constant, compactly supported}
\end{array}}.
\]
Let $\Mp(W)$ be
the metaplectic $\bC^\times$-covering of $\Sp(W)$ which acts on the
oscillator representation naturally by its definition. The lattice
model with respect to $A$ gives a section
$\omegaA \colon \Sp(W) \hookrightarrow \Mp(W)$ of the natural projection
$\Mp(W) \twoheadrightarrow \Sp(W)$ (c.f. \cite{MVW,Wa}). Let
$\SpA := \Stab_{\Sp(W)}(A) = \Set{g\in\Sp(W)|gA \subseteq A}$. We
only describe $\omegaA(g)$ for $g \in \SpA$:
\begin{equation} \label{eq:LA}
(\omegaA(g)f)(w) = f(g^{-1}w) \quad
\forall g \in \SpA, f \in \sS(A) \text{ and } w \in W.
\end{equation}
The splitting $\omegaA$ does not depend on the choice of the self-dual
lattice $A$. More precisely, we have the following proposition which follows immediately from \Cref{lem:SS}.
\begin{prop} \label{prop:liftSpA}
There is a section
\[
\xymatrix{
\omega_0 \colon {\displaystyle \bigcup_{A \text{ is self-dual}}} \Sp_A
\ar@{^(->}[r]& \Mp(W)
}
\]
such that $\omega_0|_{\SpA} = \omega_A$ for every self dual lattice
$A$.
\end{prop}
\subsection{Epipelagic supercuspidal representations of covering
groups} \label{sec:CEpi} Let $\Sigma = (x,\lambda,\chi)$ be an
epipelagic datum of order $m$. We retain the notation for the subgroup
$H_{x,\lambda} = \sfA_{x,\lambda} \ltimes \Jx$ and its character
$\psi_\lambda \otimes \chi$ in \Cref{sec:Epirep}. Since $G$ is
a member of a type I dual pair, we recall that $G_x = G_{[x]}$ in \Cref{sec:SD}
and $H_{x,\lambda}$ is a subgroup of $G_x$. We will show in
\Cref{sec:sfX} later that $G_x$ stabilizes a self-dual lattice $A$ in
$W$. Then \Cref{prop:liftSpA} gives a splitting
\[\xymatrix{
\omega_0|_{G_x}\colon G_x \ar@{^(->}[r]& \tG_x
}
\]
of $\tG_x \rightarrow G_x$. We will identify $H_{x,\lambda}$ and $G_x$
as subgroups of $\tG_x$ via $\omega_0$. Let
$\rid_{\bC^\times} \colon \bC^\times \rightarrow \bC^\times$ be the identity map.
Under this splitting, $\tG_x = G_x \times \bC^\times$ with $\bC^\times$ acting
on the oscillator representation $\sS$ via $\rid_{\bC^\times}$. Now
\[
\pi_\Sigma :=\ind_{H_{x,\lambda} \times \bC^\times}^{\tG} ((\psi_{\lambda} \otimes
\chi) \boxtimes \rid_{\bC^\times})
\]
is an irreducible supercuspidal representation of $\tG$ which is also
denoted by $\pi_\Sigma^{\tG}$ or $\pi_x^{\tG}(\lambda, \chi)$. We
will also call $\pi_\Sigma$ an epipelagic supercuspidal representation
attached to the epipelagic data $\Sigma$.
By \Cref{sec:wplus} the splitting of $G_{x,0^+}$ is canonically
defined for any $x\in \cB(\bfG,k)$. In particular, it still makes
sense to talk about positive depth minimal $K$-types. In addition
\Cref{prop:MKT} holds if we replace $\pi_x(\lambda, \chi)$ with
$\pi_\Sigma^{\tG}$ without any modification.
\section{Bruhat-Tits Buildings and Moy-Prasad filtrations of classical
groups}\label{sec:BT}
In this section we recall some known facts about the Bruhat-Tits
buildings of classical groups. Our references are \cite{BT3,BT4,BS,Le}.
\subsection{Lattice functions}
A (right) $\foo_D$-lattice $L$ in a right $D$-vector space $V$ is a
right $\foo_D$-submodule such that $L\otimes_{\foo_D} D = V$.
\begin{definition} \label{def:DLattic}
\begin{asparaenum}[1.]
\item Let $\Latt_V$ be the set of $\foo_D$-lattice valued functions
$s\mapsto \sL_s$ on $\bR$ such that
\begin{inparaenum}[(i)]
\item $\sL_s \supseteq \sL_t$ if $s<t$,
\item $\sL_{s+\val_D} = \sL_s\varpi_D$ and
\item $\sL_s = \bigcap_{t<s} \sL_t$.
\end{inparaenum}
\item We set $\sL_{r^+} := \bigcup_{t>r} \sL_t$ and $\Jump(\sL) = \set{ r \in \bR | \sL_{r} \supsetneq \sL_{r^+}
}$.
\item Given any lattice function $\sL$, we define
\begin{align*}
\fgl(V)_{\sL,r} :=& \Set{X\in \fgl(V)| X \sL_s \subseteq \sL_{s+r}, \forall s
\in \bR } & & \forall r \in \bR, \\
\GL(V)_{\sL,r} := & \Set{g\in \GL(V)| (g-1) \sL_s \subseteq \sL_{s+r},
\forall s \in \bR} & & \forall
r>0, \\
\GL(V)_{\sL} :=& \Set{g\in \GL(V)|g\sL_s\subseteq \sL_s}.
\end{align*}
\item For $r < s$, we denote $\sL_{r:s} = \sL_r/\sL_s$.
\end{asparaenum}
\end{definition}
\begin{definition} \label{def:DNorm}
\begin{asparaenum}[1.]
\item A $D$-norm of $V$ is a function $l \colon V \rightarrow \bR
\cup \set{ \infty }$ such that for all $x, y \in V$ and $d \in D$, (i)
$l(xd) = l(x) + \val(d)$, (ii) $l(x+y) \geq \min(l(x), l(y))$ and (iii)~$l(x) = \infty$ if and only if $x = 0$.
\item The norm $l$ is called {\it splittable} if there is a $D$-basis
$\set{ e_i | i \in I }$ of $V$ such that
$l(\sum_{i \in I} e_i d_i) = \inf_{i \in I}(l(e_i) + \nu(d_i))$. Let
$\cS \cN(V)$ denote the splittable norms on $V$. In this paper, all
norms refer to splittable $D$-norms.
\end{asparaenum}
\end{definition}
There is a natural bijection between $\cS \cN(V)$ and $\Latt_V$ given by
$l \mapsto ( \sL_r = l^{-1}([r,+\infty)))$. Then $\Jump(\sL)$ is the
image of $l$.
The following theorem is well known and follows directly from the definition of
Moy-Prasad filtration \cite{MP}.
\begin{thm} \label{thm:BT3.MP} The (extended) building $\cB(\GL(V))$
could be identified with $\Latt_V$ as $\GL(V)$-sets. This
identification is unique up to translation
(c.f. \cite[Theorem~2.11]{BT3}). Suppose $x\in \cB(\GL(V))$
corresponds to the lattice function $\sL \in \Latt_V$. Then
\begin{asparaenum}[(a)]
\item
$\fgl(V)_{x,r} = \fgl(V)_{\sL,r}$ for $r\in \bR$,
\item $\GL(V)_{x,r} = \GL(V)_{\sL,r}$ for $r>0$ and
\item $\GL(V)_\sL = \GL(V)_x$.
\end{asparaenum}
\end{thm}
For the rest of this paper, we will freely interchange the notion of
points in the building of $\GL(V)$, $D$-norms and lattice functions.
\subsection{Tensor products} \label{S42} Suppose $l$ and $l'$ are two
norms on $D$-modules $V$ and $V'$. Then there is an induced norm on
$W :=V \otimes_D V'$ such that
$(l\otimes l') (v\otimes v') = l(v)+l'(v')$ (c.f.~\cite[\S~1.11]{BT3}). Let $\sL$ and $\sL'$ be the corresponding lattice
functions. We denote by $\sL\otimes \sL'$ the corresponding
$\foo_k$-lattice function on $V\otimes_D V'$ where
\[
(\sL\otimes \sL')_t = \sum_{r+r'=t} \sL_r\otimes_{\foo_D} \sL'_{r'}.
\]
It is easy to see that
\begin{equation} \label{eqJLL}
\Jump(\sL\otimes \sL') = \Jump(\sL)+\Jump(\sL').
\end{equation}
The norms $l$ and $l'$ also induce a natural norm $\Hom(l,l')$ on
$\Hom_D(V,V')$ whose corresponding lattice function is
\begin{equation} \label{eqhomr}
(\Hom(\sL,\sL'))_r := \set{w\in \Hom_D(V,V')|w(\sL_s)\subseteq \sL_{s+r}' \,
\forall s\in \bR}.
\end{equation}
In particular, every norm $l$ on $V$ defines a dual norm
$l^*:=\Hom(l,\val)$ on $V^*:= \Hom_D(V,D)$.
Under the isomorphism $\Hom_D(V,V') \cong V' \otimes_D V^*$, the
norms $\Hom(l,l')$ and $l' \otimes l^*$ coincide. If $V = V'$, then
the Moy-Prasad lattice function $r\mapsto \fgl(V)_r$ defined in
\Cref{def:DLattic} is the tensor product lattice function
$\sL\otimes \sL^*$ on $\fgl(V) = \End_D(V)$.
\subsection{Self-dual lattice functions} \label{S43} Let $V$ be a
space with a non-degenerate sesquilinear form $\inn{}{}_V$.
\begin{definition} \label{D421}
\begin{asparaenum}[1.]
\item For a lattice $L$ in V, we set
\[
L^\sharp := \set{v \in V | \inn{v}{v'}_V \in \fpp_D,
\forall v'\in L }.
\]
A lattice $L$ is called {\it self-dual} if $L = L^\sharp$. A lattice $L$ is called {\it good} if
$L^\sharp \fpp_D \subseteq L \subseteq L^\sharp$.
\item For a lattice function $\sL$ we define its dual lattice function
$\sL^\sharp$ by $(\sL^{\sharp})_s = (\sL_{(-s)^+})^\sharp$. If $l$
is the norm corresponding to $\sL$, then we denote the norm
corresponding to $\sL^\sharp$ by~$l^\sharp$. If we identify $V$ with
$V^*$ using the form $\inn{}{}_V$, then the norm $l^*$ on $V^*$
translates to the norm $l^\sharp$ on $V$.
\item A lattice function $\sL$ is called {\it self-dual} if and only
if $\sL = \sL^\sharp$. In terms of norm, it is equivalent to
$l^\sharp = l$ (c.f. \cite[Prop.~3.3]{BS}) and we say that $l$ is
self-dual. Let $\sLatt$ be the set of self-dual lattice
functions. Clearly $\sLatt$ is the $\sharp$-fixed point set of
$\Latt_V$.
\item When $\sL$ is self-dual, we define
$\fgg_{\sL,r} := \fgg\cap \fgl(V)_{\sL,r}$,
$G_{\sL,r} := G\cap \GL(V)_{\sL,r}$, $G_{\sL} := G \cap \GL(V)_{\sL}$,
$\sfG_\sL := G_{\sL}/G_{\sL,0^+}$ and
$\fgg_{\sL,r:s} := \fgg_{\sL,r}/\fgg_{\sL,s}$.
\end{asparaenum}
\end{definition}
\noindent \remark If we identify $V^*\otimes_D V'^*$ as $(V\otimes_D V')^*$, then by a
calculation on a splitting basis, we have
$(l\otimes l')^* = l^* \otimes l'^*$
(c.f. \cite[(18),(21), Sect. 1.12]{BT3}). In particular, suppose that $V$
and $V'$ are formed spaces, and $\sL$ and $\sL'$ are self-dual lattice
functions. It is easy to see that
$(l_\sL\otimes l_{\sL'})^\sharp = l_\sL \otimes l_{\sL'}$, i.e. it is
self-dual. Hence $\sL\otimes \sL'$ is a self-dual lattice on
$V \otimes_D V'$.
\subsection{}
We recall that $k$ is a $p$-adic field with $p\neq 2$. For a
classical group $G$ defined over~$k$, $\cB(\bfG,k)$ could be
identified canonically with the set of splittable self-dual norms on
$V$ (c.f. \cite{BT4,GY}). The following theorem is the culmination of
\cite{BT4}, \cite{BS}, \cite{Le} and \cite{GY}.
\begin{thm}\label{thm:SB2}
\begin{enumerate}[(i)]
\item There is a natural $G$-equivariant bijection between $\cB(\bfG,k)$ and
$\sLatt$.
\item Suppose $x\in \cB(\bfG,k)$ corresponds to $\sL \in \sLatt$. Then
\begin{enumerate}[(a)]
\item $\fgg_{\sL,r} = \fgg_{x,r}$ for $r\in \bR$,
\item $G_{\sL,r} = G_{x,r}$ for $r>0$ and
\item $G_{\sL} = G_x$.
\end{enumerate}
\end{enumerate}
\end{thm}
\subsection{}
Let $r \in \Jump(\sL)$ so that
$\sfL_r := \sL_{r:r^+} = \sL_r/ \sL_{r^+}$ which is nonzero. The
sesquilinear form $\inn{}{}_V$ induces a nonzero pairing
$\sL_{r}\times \sL_{-r} \rightarrow \foo_D$ and a non-degenerate
pairing over~$\fff_D$:
\[
\sfL_r \times \sfL_{-r} \rightarrow \fff_D.
\]
In particular we have
\[
\Jump(\sL) = -\Jump(\sL).
\]
The structure of $\sfG_{\sL}$ is described in the following lemma. It
is well known so we omit its proof. Also see \Cref{app:split}
\begin{lemma} \label{lem:G0}
Let $\nu_D = \val(\varpi_D)$.
Then
\begin{equation}
\label{eq:sfG}
\sfG_{\sL} \cong \sfG_0 \times \sfG_{\rhalf} \times \prod_{r\in
\Jump(\sL)\cap (0,\rhalf)} \GL(\sfL_r)
\end{equation}
where $\sfG_0\cong \rU(\sfL_0)$ and
$\sfG_\rhalf \cong \rU(\sfL_{\rhalf})$ where $\sfL_{\rhalf}$ is
equipped with the form
$\inn{[v_1]}{[v_2]} = \inn{v_1}{v_2\varpi_D^{-1}}_V
\pmod{\fpp_D}$. \qed
\end{lemma}
\section{Tame base changes and epipelagic points}
\subsection{} \label{S51} Let $D$ be the division algebra over $k$ as in \Cref{SS21}.
Let $V$ be a $D$-module with an $\epsilon$-Hermitian sesquilinear form $\inn{}{}_V$ and let $\bfG$ be the classical group defined over $k$ such that $G := \bfG(k) = \rU(V,\inn{}{}_V)$. Suppose $E$ is a tamely ramified
finite extension of $k$ or $\kur$ such that $\bfG$ splits. In this
section we study the relations between buildings under tamely ramified
field extensions.
In all cases, it is standard to construct an $E$-vector space $\pV$
obtained by certain base change of $V$ so that
$\bfG(k) \subseteq \bfG(E)$ are subgroups of $\GL_E(\pV)$. By
\cite{GY} there is a canonical bijection
$\cB(\bfG,k) \iso
\cB(\bfG,E)^{\Gal(E/k)}$.
The next proposition describes this bijection in terms of splittable
norms on $V$ and $\pV$.
\begin{prop} \label{lem:TBC}
We identify buildings of classical groups with the corresponding set of splittable norms.
\begin{asparaenum}[(i)]
\item Suppose $D = k$. Let $\pV = V\otimes_k E$ and let
$i_V \colon V \rightarrow \pV$ be given by $v\mapsto v\otimes 1$. Let
$\inn{}{}_\pV$ be the $E$-linear extension of $\inn{}{}_V$. Then
$\bfG(E) = \rU(\pV,\inn{}{}_\pV)$. The bijection
$\cB(\bfG,k) \iso \cB(\bfG,E)^{\Gal(E/k)}$
is given by $l_V \mapsto l_V \otimes_k (\val|_E)$ and its inverse map is
$l_\pV \mapsto l_\pV \circ i_V$.
\item Suppose $D$ is a quadratic extension of $k$. We fix a field
embedding $\iota \in \Hom_k(D,E)$ and view $D$ as a subfield of
$E$. Let $\pV = V\otimes_D E$ and let $i_V : V \rightarrow \pV$ be
given by $v\mapsto v\otimes 1$. Then $\bfG(E) \cong \GL_E(\pV)$. The bijection
$\cB(\bfG,k) \iso \cB(\bfG,E)^{\Gal(E/k)}$
is given by $l_V \mapsto l_V\otimes_D (\val|_E)$ and its inverse map is
$l_\pV \mapsto l_\pV \circ i_V$.
\item Suppose $D$ is the quaternion algebra over $k$. We fix a
subfield $L$ of $E$ which is a quadratic extension of $k$. We identify $L$ with a subfield of $D$
and fix a $d \in D$ such that $d^2 \in k^\times$, $d^\tau = - d$ and
$\Ad(d)$ acts on $L$ by the non-trivial Galois action. Let
$\pr \colon D \rightarrow L$ be the projection of $D = L\oplus Ld$. Then
$Q(v_1,v_2):=\pr(\inn{v_1 d}{v_2}_V)$ defines an $L$-bilinear from
on $V$. Let $\pV = V\otimes_L E$ and let $\inn{}{}_{\pV}$ be the
$(-\epsilon)$-symmetric $E$-linear extension of $Q$. Then {$\bfG(E) \cong \rU(\pV,
\inn{}{}_{\pV})$.} The bijection
$\cB(\bfG,k) \iso \cB(\bfG,E)^{\Gal(E/k)}$
is given by $l_V \mapsto l_V\otimes_L (\val|_E)$ and its inverse map is
$l_\pV \mapsto l_\pV \circ i_V$.
\end{asparaenum}
\end{prop}
Before we give the proof of \Cref{lem:TBC}, we first recall the
uniqueness result stated in \cite[\SSS{1.2}]{GY}.
\begin{lemma} \label{lem:BBC}
Let $\cB$ and $\cB'$ be two $G$-sets satisfying the axioms of building of
$\bfG$ over $k$ (See \cite[\SSS{2.1}]{Tits} and
\cite[\SSS{1.9.1}]{GY}.). Let $j\colon \cB \to \cB'$ be a bijection such
that
\begin{enumerate}[(i)]
\item $j$ is $G$-equivariant, i.e. $j(g\cdot x) = g \cdot j(x)$ for all
$g \in G$, and
\item its restriction to an apartment $\cA$ is affine.
\end{enumerate}
Then $j$ is unique up to the translation by an element in
$X_*(\Cent(G)^\circ) \otimes \bR$.
\end{lemma}
In our cases, $\Cent(G)$ is anisotropic so the map $j$ is unique.
\begin{proof}[Proof of \Cref{lem:TBC}]
\begin{asparaenum}[(i)]
\item {
We consider the following diagram:
\[
\xymatrix{
\cB(\bfG,k) \ar@{==}[r] \ar@{^(->}[d]& \cB(\bfG,E)^{\Gal(E/k)} \ar@{^(->}[d] \\
\cB(\GL_k(V)) \ar@{=}[r] & \cB(\GL_E(\pV))^{\Gal(E/k)}.
}
\]
The buildings in the top row are the fixed point sets of the involutions $\sharp$ of the buildings in the bottom row. The bottom map is
$l_V \mapsto l_V \otimes (\val|_E)$. It is a $\GL_k(V)$-invariant map. It is a bijection, since it suffices to check this on an apartment, where it is obvious. The map
sends self-dual norms to self-dual norms, hence it induces the top row isomorphism, by restriction. It is the canonical
isomorphism $\cB(\bfG,k) \iso\cB(\bfG,E)^{\Gal(E/k)}$ by \Cref{lem:BBC}.
This proves~(i). }
\item
We refer to the computation in \cite[\SSS{1.13}]{BT3}.
Let $\iota_1, \iota_2$ be two $k$-embeddings of $D$ into $E$ (so $\iota$ is one of the two).
Let $\pV^{\iota_i} = V\otimes_{D,\iota_i} E$.
Then $V\otimes_k E\cong \pV^{\iota_1} \oplus \pV^{\iota_2}$, so we have a natural action of $\Gal(E/k)$ on $\pV^{\iota_1} \oplus \pV^{\iota_2}$.
Now Part (ii) follows by applying a similar argument as in (i) to the following diagram:
\[
\xymatrix{
\cB(\bfG,k) \ar@{==}[r] \ar@{^(->}[d]& \cB(\bfG,E)^{\Gal(E/k)} \ar@{^(->}[d] \\
\cB(\GL_D(V)) \ar@{=}[r] & \cB(\GL_E(\pV^{\iota_1})\times \GL_E(\pV^{\iota_2}))^{\Gal(E/k)}.
}
\]
\item Note that
$H(v_1,v_2) = \pr(\inn{v_1}{v_2}_V)$ defines a Hermitian form on
$V$. Moreover, $\rG(k) = \rU(V,Q) \cap \rU(V,H)$.
Part (iii) follows by applying a similar argument as in (i) to the following diagram:
\[
\xymatrix{
\cB(\bfG,k) \ar@{==}[r] \ar@{^(->}[d]& \cB(\rU(\pV,
\inn{}{}_{\pV}))^{\Gal(E/k)} \ar@{^(->}[d] \\
\cB(\rU(V,H)) \ar@{=}[r] & \cB(\GL_L(\pV))^{\Gal(E/k)}. }
\]
\end{asparaenum}
\end{proof}
For an $\foo_D$-lattice function $\sL$ corresponding to
$x \in \cB(\bfG,k)$, we will denote by $\sL^E$ the $\foo_E$-lattice
function in $\pV$ corresponding to $x \in \cB(\bfG,E)^{\Gal(E/k)}$ in
the above proposition. We need the following application of \Cref{lem:BBC}
in our study of the epipelagic points.
\begin{lemma} \label{L621} Let $\sL$ be the self-dual lattice function
corresponding to a point $x$ in $\cB(\bfG,k)$ of order $m$. Suppose
$\bfG$ splits under a tamely ramified extension $E$ with
ramification index $m$. Then $\Jump(\sL)$ is contained in either
$\frac{1}{m} \bZ$ or $\frac{1}{2 m} + \frac{1}{m} \bZ$.
\end{lemma}
\begin{proof}
We have $\val(E) = \fracmm \bZ$. Let $\sL^E$ be the
$\foo_E$-lattice function. Let $J = \Jump(\sL)$ and
$J^E=\Jump(\sL^E)$. By \cite[\SSS{4.2}]{RY}, $\sL^E$ corresponds to
a hyperspecial point in $\bfG(E)$. Hence $J^E = j_0 + \fracmm \bZ$
for some $j_0\in [0,\fracmm)$.
By \Cref{lem:TBC}, we have $J \subseteq J^E = J+\fracmm
\bZ$. Since $\sL$ is self-dual, $J = -J$. Hence
$-j_0+\fracmm \bZ =-J^E \subseteq - J +
\fracmm\bZ \subseteq J +\fracmm \bZ \subseteq J^E + \fracmm \bZ
= j_0 +\fracmm \bZ$. Therefore $j_0 = 0$ or $\fracdmm$.
The lemma follows.
\end{proof}
\subsection{Epipelagic points} \label{sec:EP}
Let $x \in \cB(\bfG,k)$ be an epipelagic point of order $m$.
We recall that $\kur$ is
the maximal unramified extension of $k$. Let~$E$ be the totally
ramified extension of~$\kur$ of degree~$m$. We fix a uniformizer
$\varpi_E$ such that $\varpi_E^m = \varpi$. Let
$\Gal(E/\kur) = \langle \sigma \rangle$ where
$\sigma(\varpi_E) = \zeta \varpi_E$ and $\zeta$ is a primitive $m$-th
root of unity. Let $\Fr \in \Gal(E/k)$ denote the lift of the Frobenius
automorphism in $\Gal(\kur/k)$ such that $\Fr(\varpi_E) = \varpi_E$. Now
$\Gal(E/k) = \braket{\Fr, \sigma}$.
The group $\bfG$ splits over $E$ \cite[\SSS{4.1}]{RY}. We recall \Cref{sec:N}
that $\bfT$ is a maximally $\kur$-split torus in $\bfG$ containing $\bfS$
and defined over $k$. Let $\bfY = \Cent_{\bfG}(\bfT)$ be a Cartan
subgroup of $\bfG$. Then $x$ is a hyperspecial point in
$\cA(\bfY,E)$. We have $\bfG(E)_x^\sigma = \bfG(\kur)_x$,
$\bfG(E)_{x,r}^\sigma = \bfG(\kur)_{x,r}$ for $r > 0$, and
$\varpi_E^d \fgg(E)_{x,0} = \fgg(E)_{x,\frac{d}{m}}$. The building
$\cB(\bfG,\kur)$ embeds into $\cB(\bfG,E)$ as the $\sigma$-invariant set
and
\[
\cB(\bfG,k) = \cB(\bfG,\kur)^\Fr = (\cB(\bfG,E)^\sigma)^\Fr.
\]
We set $D(E) = D\otimes_k E$. We equip $D(E)$ with the tensor product norm
of valuations of $D$ and $E$. Let $\sE$ and
$\sD$ be the lattice functions in $E$ and $D$ respectively defined by their valuations. Then
$\sD\otimes \sE$ is the corresponding lattice function on $\DE$. Let
$\foo_{D(E)} = (\sD\otimes \sE)_0$,
$\fpp_{D(E)} = (\sD\otimes \sE)_{0^+}$ and
$\fff_{D(E)} = \foo_{D(E)}/\fpp_{D(E)}$ which is a semisimple algebra over $\fff_E$.
Let $\sL$ be the self-dual lattice function in $V$ corresponding to
$x \in \cB(\bfG,k)$. Let $\sL_E := \sL \otimes \sE$ be the self-dual
$\foo_k$-lattice function in $V(E) := V \otimes_k E$. In fact it is an
$\fooDE$-lattice function. We have following situations.
\begin{enumerate}[(i)]
\item If $D = k$, then {$\sL_E = \sL^E$.}
\item If $D$ is a quadratic extension of $k$, then
$D(E) \cong E\times E$, $V(E) = \pV^{\iota_1}\oplus \pV^{\iota_2}$,
$\sL_E = \sL^{\iota_1, E} \oplus \sL^{\iota_2, E}$ where
$\pV^{\iota_1}$ and $\pV^{\iota_2}$, $\sL^{\iota_1,E}$ and
$\sL^{\iota_2,E}$ correspond to the two different $k$-embeddings of
$D$ into $E$.
\item Suppose $D$ is the quaternion algebra over $k$. We fix a quadratic
extension $L$ of $k$ in $E$. Then $D(E) \cong \Mat_2(E)$,
$V(E) = \pV^{\iota_1}\oplus \pV^{\iota_2}$ and
$\sL_E = \sL^{\iota_1, E} \oplus \sL^{\iota_2, E}$ where
$\sL^{\iota_1,E}$ and $\sL^{\iota_2,E}$ are $\foo_E$-lattice
functions corresponding to the two different $k$-embeddings of $L$
into $E$.
\end{enumerate}
Clearly, $\Jump(\sL_E) = \Jump(\sL^{\iota_i, E}) = \Jump(\sL) + \fracmm\bZ$. The
Galois group $\Gal(E/k)$ acts on $\VE$ by $s(v \otimes x) = v \otimes s(x)$ for
$s \in \Gal(E/k)$, $v \in V$ and $x \in E$. For $g\in \bfG(E)$
and $s \in \Gal(E/k)$, we have $s(g) = s \circ g\circ s^{-1}$ as $\DE$-linear
automorphism on $\VE$. In Cases (ii) and (iii), under the
decomposition, $\bfG(E)$ acts diagonally on
$V(E) = \pV^{\iota_1}\oplus \pV^{\iota_2}$.
Extending the notation in \Cref{def:DLattic},
we have $\bfG(E)_x = \bfG(E)_{\sL_E}$, $\bfG(E)_{x,r} = \bfG(E)_{\sL_E,r}$ and
$\fgg(E)_{x,r} = \fgg(E)_{\sL_E,r}$ by \Cref{lem:TBC}.
\subsection{Kac-Vinberg gradings} \label{sec:Kac-Vinberg}
{In \cite[\SSS{4}]{RY}, Reeder and Yu connect the Moy-Prasad filtration at an epipelagic point with the Kac-Vinberg gradings of Lie algebras over the residue fields.
We review their results here.}
By the classification of hyperspecial points for split classical groups (see Remark in \Cref{sec:apartment}), we can pick a point $x_0 \in \cA(\bfT, \kur)$ following the recipe in \cite[Section 3.2]{RY} such that $\Jump(\sLEz) = \Jump(\sLE)$ where $\sLEz$ is the lattice function in $V(E)$ corresponding to $x_0$.
The action of the generator $\sigma$ of $\Gal(E/\kur)$ on the Cartan subgroup $\bfY$ induces an action $\vartheta$ on $X_*$.
Then $x = x_0 + \frac{1}{m}\crho$ where $\crho \in X_*^\vartheta$.
Let $t:=\crho(\varpi_E) \in \bfG(E)$.
The $\foo_{D(E)}$-lattice function corresponding to $x_0$ is $\sLEz = t^{-1}
\sLE$.
We have isomorphisms
\[
\bfG(E)_{x_0} \xrightarrow{\Ad(t)} \bfG(E)_x \quad \text{and} \quad
\fgg(E)_{x_0,0} \xrightarrow{\Ad(t)} \fgg(E)_{x,0}\xrightarrow{\varpi_E^j}
\fgg(E)_{x,\frac{j}{m}}.
\]
Let $\sfGE = \bfG(E)_{x_0}/ \bfG(E)_{x_0,0^+}$ and
$\sfggE = \fgg(E)_{x_0,0}/ \fgg(E)_{x_0,0^+}$.
Let $\vartheta$ be the automorphisms on $\sfG(E)_{x_0}$ and $\sfg(E)_{x_0}$ induced by the $\sigma$ actions
on $\bfG(E)_{x_0}$ and $ \fgg(E)_{x_0,0}$ respectively.
Let $\theta :=\Ad(t^{-1})\circ \sigma \circ \Ad(t)$ be the automorphisms on
$\sfGE$ and on $\sfggE$ induced by the $\sigma$ actions
on $\bfG(E)_x$ and $\fgg(E)_{x,0}$.
Let $\sfggE^{\theta,\zeta^{-j}}$
be the $\zeta^{-j}$-eigenspace of $\theta$ on $\sfggE$. Then
\begin{enumerate}[(a)]
\item $\theta = \Ad(\tt) \vartheta$ where $\tt =
t^{-1} t^\sigma = \crho(\zeta) \pmod{G(E)_{x_0,0^+}}$;
\item
$\Ad(t) \colon \sfGE^\theta \iso \sfG_x := \bfG(\kur)_{x}/\bfG(\kur)_{x,0^+}$
is an isomorphism and
\item
$ \varpi_E^j\ad(t) \colon \sfggE^{\theta,\zeta^{-j}} \iso
\sfg_{x,\frac{j}{m}}$
is $\sfGE^\theta$-equivariant with $\sfGE^\theta$ acting on the right hand
side via~(b). Here we recall $\sfg_{x,\frac{j}{m}}$ in \Cref{sec:EpiP}.
\end{enumerate}
By putting $j = -1$ in (c), we define
$\iotag :=\varpi_E^{-1} \ad(t)\colon \sfggE^{\theta, \zeta} \iso \sfg_{x,-\fracmm}$.
Then $\iotag$ is a bijection between the set of stable vectors for the
$\sfGE^\theta$ action on $\sfggE^{\theta,\zeta}$ and the set of stable
vectors for the action $\sfGKx$ on $\sfg_{x,-\fracmm}$. The former
was studied by Vinberg \cite{Vin} and Levy \cite{Levy}.
\section{Moment maps}\label{sec:mm}
\subsection{} \label{sec:MM.gen} Let $W = V \otimes_D V'$. Using the
sesquilinear forms, we define $\Psi\colon W \iso \Hom_D(V,V')$ and
$\Psi' \colon W \iso \Hom_D(V',V)$ by
\[
\Psi(v \otimes v') (v_1) = v' \inn{v}{v_1}_{V} \mbox{ and } \Psi'(v
\otimes v') (v_1') = v \inn{v'}{v_1'}_{V'}
\]
for all $v, v_1 \in V$ and $v', v_1' \in V'$. Now $g\in G$ and
$g'\in G'$ acts on $\Hom_D(V',V)$ by the formula
$(g,g')\cdot w = g'w g^{-1}$.
\begin{definition}\label{def:Parings}
\begin{asparaenum}[1.]
\item We define a non-degenerate $G$-invariant symmetric $k$-bilinear
form\footnote{We warn that our trace form $\Bg$ has a factor of $\frac{1}{2}$.} $\Bg \colon \fgg\times \fgg\to k$ by $\Bg(X_1,X_2) = \half
\tr_{D/k}\tr(X_2^*X_1)$.
\item
We define an operator $\star\colon \Hom_D(V,V') \rightarrow
\Hom_D(V',V)$ by
\[
\inn{w(v)}{v'}_{V'} = \inn{v}{w^\star(v')}_V \quad \forall
w\in \Hom_D(V,V'), v\in V, v'\in V'.
\]
We note that if $x \in W$ and $\Psi(x) = w$, then $\Psi'(x) = w^\star$.
\item We define the moment map $M\colon W \cong \Hom_D(V,V') \to \fgg$
and $M'\colon W\to \fgg'$ by
\[
M(w) = w^\star w \mbox{ and } M'(w) = w w^\star.
\]
\end{asparaenum}
\end{definition}
By definition $M$ and $M'$ are $G \times G'$-equivariant.
\begin{lemma}\label{lem:MPsi}
Suppose $w_1,w_2, w\in W$, $X\in \fgg$ and $X'\in \fgg'$. Then
\begin{asparaenum}[(a)]
\item $\inn{w_1}{w_2} = \tr_{D/k}\tr(w_2^\star w_1)$,
\item $\inn{X\cdot w}{w} = 2 \Bg(M(w),X)$ and $\inn{X'\cdot
w}{w} = 2 \Bgp(-M'(w),X')$.
\end{asparaenum}
\end{lemma}
The proof is a straightforward computation using \eqref{equ4} and the
definition of $\star$. We will leave it to the reader.
\subsection{}
Let $\sL$ and $\sL'$ be two self-dual lattice functions on $V$ and
$V'$ respectively. Let $\sB =\sL\otimes \sL'$ on $W=V\otimes_D V'$.
\begin{lemma} \label{lem:MB}
\begin{enumerate}[(i)]
\item We have $\Jump(\sB) = \Jump(\sL) + \Jump(\sL')$.
\item The lattice function $\sB$ is self-dual in $W$,
i.e. $\sB_r^\sharp = \sB_{-r^+}$.
\item Under the isomorphism $\Psi\colon W \iso \Hom_D(V,V')$,
\[
\Psi(\sB_r) = \Set{w\in \Hom_D(V,V')| w\sL_s \subseteq \sL'_{s+r}\ \forall s\in \bR}.
\]
\item We have $(\Psi(\sB_r))^\star = \Psi'(\sB_r)$.
\item \label{lem:MB.M} We have $M(\sB_r) \subseteq \fgg_{\sL,2r}$ and
$M'(\sB_r) \subseteq \fgg'_{\sL',2r}$.
\end{enumerate}
\end{lemma}
\begin{proof}
Part (i) is \Cref{eqJLL}. Part (ii) is explained in the Remark in \Cref{S43}.
By~\eqref{eqhomr} the right hand side of (iii) is the lattice function on $\Hom_D(V,V') = V^* \otimes_D V'$.
On the other hand $\Psi$ maps $V \otimes V'$ to $V^* \otimes V'$.
We have seen in \Cref{S43} that the norm $l^*$ on $V^*$ translates to the norm $l^\sharp = l$ on $V$. It follows that the lattice function on the right hand side of (iii) corresponds to the lattice function $\sB$ under $\Psi$. This proves~(iii). If $w \in W$ then $\Psi'(w) = \Psi(w)^\star$. This proves (iv).
Part (v) follows directly from (iii) and~(iv).
\end{proof}
\def\bfGsL{{\bfG_{\sL}}}
\def\bfGsLp{{\bfG'_{\sL'}}}
\subsection{}
We could view $\sB_s$, $\fgg_{\sL,2s}$ and $\fgg'_{\sL',2s}$ as
schemes over $\foo_k$. Since $\star$ is $\foo_k$-linear, the moment
maps defined over the generic fibers as in \Cref{sec:MM.gen} extend to
morphisms between these $\foo_k$-schemes. The $\foo_k$-group scheme
$\bfGsL\times \bfGsLp$ acts on all these objects and the moment maps
are equivariant maps.
Let $\sfW_s = \sB_s/\sB_{s^+}$,
$\sfg_{\sL,2s} = \fgg_{\sL,2s}/\fgg_{\sL,2s^+}$ and
$\sfg'_{\sL',2s} = \fgg'_{\sL',2s}/\fgg'_{\sL',2s^+}$. We get
morphisms, as certain quotients of the moment maps over the special
fiber,
\begin{equation} \label{eq:MM.s}
\sfM_s \colon \sfW_s \rightarrow \sfg_{\sL,2s}
\quad \text{and} \quad \sfM'_s \colon \sfW_s \rightarrow \sfg'_{\sL',2s}.
\end{equation}
The actions of $\bfGsL\times \bfGsLp$ reduce to
$\sfG_{\sL}\times \sfG'_{\sL'}$ actions on $\sfW_s$, $\sfg_{\sL,2s}$
and $\sfg_{\sL',2s}'$. These are the moment maps over the residual
field $\fff$ which we will study later.
\section{Theta correspondences I} \label{sec:thetaI}
\subsection{} \label{sec:bfX}
In this section we let $(G,G')$ be a reductive dual pair
in $\Sp(W)$ as in \Cref{sec:dualpair}.
Let $\Sigma = (x,\lambda, \chi)$ and $\Sigma' = (x',\lambda',\chi')$
be epipelagic supercuspidal data for $G$ and $G'$ respectively. Let
$\piSigma = \pi_x^{\tG}(\lambda,\chi)$ (resp.
$\piSigmap = \pi_{x'}^{\tG'}(\lambda',\chi')$) be the corresponding epipelagic
representation of $\tG$ (resp. $\tG'$). From now on, we assume
$\theta(\piSigma) = \piSigmap$. As discussed in \Cref{S14}, $x$ and $x'$ are
both epipelagic points of order $m$ for some $m\geq 2$.
Let $\sL$ and $\sL'$ be self dual lattice functions corresponding to $x$ and
$x'$ respectively. Let $\sB = \sL \otimes \sL'$. Let
$\sfX = \sfW_{-\fracdmm} = \sB_{-\fracdmm}/ \sB_{-\fracdmm^+}$. We denote the
moment maps defined in \eqref{eq:MM.s} by
$\sfM \colon \sfX \rightarrow \sfg_{\sL,-\fracmm}$ and
$\sfM' \colon \sfX \rightarrow \sfg'_{\sL',-\fracmm}$.
\begin{prop} \label{prop:M.w} Suppose that
$\theta(\piSigma) = \piSigmap$. Then there exists a
$\bar{w} \in \sfX$ such that $ \lambda = \sfM(\bar{w}) $ and
$\lambda' = -\sfM'(\bar{w})$.
\end{prop}
\begin{proof}
By \cite{Pan} there exists
$(y,y') \in \cB(\bfG,k)\times \cB(\bfG',k)$,
$\sB' = \sL_y \otimes \sL_{y'}'$ and $w \in \sB_{-\frac{1}{2m}}'$
such that $M(w)+\fgg_{y,-\fracmm^+}$ is an unrefined minimal
$K$-type of $\piSigma$, and $-M'(w)+\fgg'_{y',-\fracmm^+}$ is an
unrefined minimal $K'$-type of $\piSigmap$. The moment maps $M$ and $M'$
commute with the action of $G \times G'$-conjugation. By
\Cref{prop:MKT} and conjugating by $G \times G'$, we may assume
that $y = x$, $y' = x'$, $w \in \sB_{-\frac{1}{2m}}$,
$M(w) + \fgg_{x,-\frac{1}{m}^+} = \sfM(\bar{w}) = \lambda$ and
$-M'(w)+\fgg_{x',-\frac{1}{m}^+}' = -\sfM'(\bar{w}) = \lambda'$
where $\bar{w} = w + \sB_{-\frac{1}{2m}^+} \in \sfX$.
\end{proof}
\begin{cor} \label{cor:JP} We have $\Jump(\sB) \subseteq \frac{1}{2m} +
\frac{1}{m} \bZ$. Moreover either
\begin{enumerate}[(i)]
\item $\Jump(\sL) \subseteq \frac{1}{m} \bZ$ and $\Jump(\sL')
\subseteq \frac{1}{2m} + \frac{1}{m} \bZ$ or \item $\Jump(\sL)
\subseteq \frac{1}{2m} + \frac{1}{m} \bZ$ and $\Jump(\sL')
\subseteq \frac{1}{m} \bZ$.
\end{enumerate}
\end{cor}
\begin{proof}
By \Cref{L621}, $\Jump(\sL)$ (resp. $\Jump(\sL')$) is a subset of
$\frac{1}{m} \bZ$ or $\frac{1}{2m} + \frac{1}{m} \bZ$. From the
proof of the last proposition, $\bar{w}$ is a nonzero element in
$\sB_{-\frac{1}{2m}}/\sB_{-\frac{1}{2m}^+}$. In particular
$-\frac{1}{2m} \in \Jump(\sB) = \Jump(\sL) + \Jump(\sL')$. The
corollary follows.
\end{proof}
\subsection{} \label{sec:MVinberg}
Reeder and Yu connect the Moy-Prasad filtration at an epipelagic point with the Kac-Vinberg grading of Lie algebras over the residue fields, as
described previously in \Cref{sec:Kac-Vinberg}. Now we relate this with the moment maps.
Let $\sLEz = t^{-1} \sLE$ and $\sLEzp = {t'}^{-1} \sLEp$ denote
the $\fooDE$-lattice functions corresponding to $x_0$ and $x_0'$ respectively as
in \Cref{sec:EP}. By \Cref{cor:JP} we are in one of the following two cases.
\begin{enumerate}[(a)]
\item We have $\Jump(\sLEz) = \fracmm\bZ$ and
$\Jump(\sLEzp) = \fracdmm + \fracmm \bZ$. In this case we set
$\sfV :=\sLEz_{,0:0^+}$ and $\sfV' :=\sLEzp_{,-\fracdmm:-\fracdmm^+}$.
\item We have $\Jump(\sLEz) = \fracdmm + \fracmm\bZ$ and
$\Jump(\sLEzp) = \fracmm \bZ$. In this case we set
$\sfV :=\sLEz_{,-\fracdmm:-\fracdmm^+}$ and $\sfV' :=\sL'^0_{E,0:0^+}$.
\end{enumerate}
Both $\sfV$ and $\sfV'$ are $\fffDE$-modules. In Case (a), we
assign a non-degenerate Hermitian forms $\sfV$ by
$\iinn{}{}_{\sfV} := \inn{}{}_V \pmod{\fpp_{D(E)}}$ and a non-degenerate
Hermitian form on $\iinn{}{}_{\sfV'} := \inn{}{}_{V'}\varpi_E \pmod{\fpp_{D(E)}}$.
In Case (b), the bilinear forms are defined similarly.
Thus $\sfGE = \rU(\sfV)$ (resp. $\sfGE' = \rU(\sfV')$) as $\fff_{\DE}$-linear transformations on $\sfV$ (resp. $\sfV'$)
preserving the form.
Let $\sfL^0(E)_r = \sL_{E,r:r^+}^0$.
The actions $\sigma$ and $t^{-1} \circ \sigma \circ t$ on
$\sL_{E,r}^0$ induce actions on $\sfL^0(E)_r$ which we denote by
$\vartheta$ and $\theta$ respectively. It is compatible with the
$\vartheta$ and $\theta$ actions on $\sfGE$ defined in \Cref{sec:Kac-Vinberg} in the sense that for
$g\in \sfGE$ we have
$\vartheta(g) = \vartheta \circ g \circ \vartheta^{-1}$ and
$\theta(g) = \theta \circ g\circ \theta^{-1}$ as linear
transformations on $\sfL^0(E)_r$.
Let $\sB(E) = \sB \otimes_{\foo_k} \sE$ and
$\sB^0(E) = (t^{-1}, t'^{-1}) \sB(E) = \sLEz \otimes_{\foo_{D(E)}}
\sLEzp$.
\def\sfMz{\sfM^0}
Let $\sfW^0_{-\fracdmm} = \sB^0(E)_{-\fracdmm}/\sB^0(E)_{-\fracdmm^+}$ and
$\tsfW := \sfV \otimes_{\fffDE} \sfV' \cong \Hom_{\fffDE}(\sfV,\sfV')$. We
observe the following diagram:
\begin{equation} \label{eq:tMM}
\vcenter{
\xymatrix@C=3em{
\tsfW \ar[d]_{\tsfM} &\ar[l]_\sim \sfW^0_{-\fracdmm} \ar[d]^{\sfMz}
& \ar[l]
\sB^0(E)_{-\fracdmm} \ar[d]^{M} \ar[r]_{\sim}^{(t,t')}
&
\sB(E)_{-\fracdmm} \ar[d]^{M} \\
\sfggE & \sfg^0_{-\fracmm} \ar[l]_{\varpi_E}^\sim
& \ar[l]
\fgg(E)_{x_0,-\fracmm}
\ar[r]_<>(.5){\sim}^<>(.5){\Ad(t)}
&
\fgg(E)_{x,-\fracmm}.
}}
\end{equation}
In the above diagram, the
$\bfG(E)_{x}\times \bfG'(E)_{x'}$-equivariant map $M$ on the far right
translates to the $\sfGE \times \sfGEp$-equivariant map $\tsfM$ on the
far left.
Here $\sfMz$ is defined by \eqref{eq:MM.s} with respect to lattices $\sLEz$ and
$\sLEzp$, and $s=-\fracdmm$.
The explicit formula for the map $\tsfM$ is exactly the same as that for $M$ with respect to the forms
$\iinn{}{}_{\sfV}$ and $\iinn{}{}_{\sfV'}$. We remark that
$\sfW^0_{-\fracdmm}$ is not equipped with any sesquilinear form. On
the other hand $\tsfW$ has isomorphic vector space structure as
$\sfW^0_{-\fracdmm}$ but it is equipped with a tensor product form.
\subsection{The ranks of $G$ and $G'$}
Let $r$ be the rank of $\fgg$ and let $P = P(X)$ be the coefficient of
$z^r$ in $\det(z {\mathrm{I}}_\fgg +\ad X)$. Then $P$ is an $\Ad(G)$-invariant homogeneous
rational function on $\fgg$ defined over $k$ such that the set of
regular semisimple elements in $\fgg$ is $P^{-1}(\bA-\set{0})$ where $\bA$ is the
affine line.
The following lemma is a consequence of \cite[Lemma~13]{Gross}.
\begin{lemma}\label{lem:reg}
Let $\lambda$ be a stable vector in
$\fgg_{x,-\frac{1}{m}:-\frac{1}{m}^+}$ and
$\gamma\in \fgg_{x,-\frac{1}{m}}$ be a lifting of $\lambda$. Then
$\gamma$ is a regular semisimple element. Let
$\bfT = \Cent_\bfG(\gamma)$. Then $\gamma$ is a good element of depth
$-\fracdmm$ with respect to $\bfT$, i.e. for every root $\alpha$ of
$\bfG(\barkk)$ with respect to $\bfT(\barkk)$, $\dalpha(\gamma)$ is
nonzero and $\val(\dalpha(\gamma)) = -\fracmm$.
\end{lemma}
\begin{proof}
Let $f(X) = \varpi_E \Ad(t)(X)$ and let $\barff$ be the induced map
in the following diagram:
\[
\xymatrix@R=1em{
\gamma \ar@{|->}[d] \ar@{}[r]|<>(.5){\in}&\fgg_{x,-\fracmm}\ar[d]
\ar@{}[r]|*{\subset} &\fgg(E)_{x,-\fracmm} \ar@{->>}[d] \ar[r]^{f} &\fgg(E)_{x_0,0} \ar@{->>}[d]\\
\lambda \ar@{}[r]|<>(.5){\in}&\fgg_{x,-\frac{1}{m};-\frac{1}{m}^+}\ar@{}[r]|*{\subset}&\sfg_{x,-\fracmm}(E) \ar[r]^<>(.5){\barff} & \sfg_{x_0}(E)\makebox[0em][l]{$=\sfggE$.}
}
\]
Since $\lambda$ is a stable vector, $\barff(\lambda)$ is a regular
semisimple element in $\sfggE$ and
$[P(f(\gamma))]\neq 0 \in \fff_E$. Therefore, $P(\gamma) \neq 0$ and
$\gamma$ is a regular semisimple element in $\fgg$.
Next we prove that $\gamma$ is good. Let
$R(\bfG(\barkk),\bfT(\barkk))$ be the set of roots. Let
$T'(\barkk) := \Ad(t)T(\barkk)$. Then
$\alpha \mapsto \alpha' := \alpha\circ \Ad(t^{-1})$ gives a map
$R(\bfG(\barkk),\bfT(\barkk)) \rightarrow
R(\bfG(\barkk),\bfT'(\barkk))$.
Moreover $\alpha'$ reduces to a root $\overline{\alpha'}\in \sfggE$.
For any $\alpha \in R(\bfG(\barkk),\bfT(\barkk))$,
$\overline{d\alpha'}([f(\gamma)]) \neq 0$ since
$[f(\gamma)] = \bar{f}(\lambda)$ is regular semisimple in
$\sfggE$. This implies that $\val(\dalpha(\gamma)) = -\fracmm$ which
proves the lemma.
\end{proof}
\begin{prop}\label{prop:RSP}
Suppose $\theta(\piSigma) =\piSigmap$.
Then $(\bfG,\bfG')$ or $(\bfG',\bfG)$ is one of the following types:
\begin{inparaenum}[(i)]
\item $(\rD_n,\rC_n)$,
\item $(\rC_n,\rD_{n+1})$,
\item $(\rC_n,\rB_n)$,
\item $(\rA_n,\rA_n)$,
\item $(\rA_n,\rA_{n+1})$.
\end{inparaenum}
\end{prop}
\begin{proof}
By \Cref{prop:M.w} there exists $w \in \sB_{-\fracdmm} \subset W$ such that
$[M(w)] = \sfM(\bar{w}) \in \fgg_{x,-\frac{1}{m}:-\frac{1}{m}^+}$
and
$[M'(w)] = \sfM'(\bar{w}) \in
\fgg'_{x',-\frac{1}{m}:-\frac{1}{m}^+}$
are stable vectors. By \Cref{lem:reg}, both $M(w)$ and $M'(w)$ are
regular semisimple elements.
Now we show that (i) to (v) list all possible cases which satisfy
the following condition.
\begin{equation}\label{equ17}
\text{There is a
$w\in W$ such that $M(w)$ and $M'(w)$ are both regular semisimple.}
\end{equation}
We may base change to the algebraic closure $\barkk$ so that $G = \bfG(\barkk)$
and $G' = \bfG'(\barkk)$ are split groups. We fix a maximal (split) torus $Y$
and identify its Lie algebra $\fyy$ with $\bA^{\dim\fyy}$. We list
$P|_\fyy = \prod_{\alpha\in \Phi(G,Y)}d\alpha$ explicitly when
$G$ is one the following groups.
\begin{description}
\item[$\GL(n)$] $\fyy = \bA^n$,
$P(a_1,\cdots, a_n) = \prod_{i \neq j} (a_i-a_j)$,
\item[$\Sp(2n)$] $\fyy = \bA^n$,
$P(a_1,\cdots, a_n) = \prod_{i\neq j}(a_i^2- a_j^2)\prod_j
(-4a_j^2)$,
\item[$\rO(2n)$] $\fyy = \bA^n$,
$P(a_1,\cdots, a_n) = \prod_{i\neq j}(a_i^2-a_j^2)$,
\item[$\rO(2n+1)$] $\fyy = \bA^n$,
$P(a_1,\cdots, a_n) = \prod_{i\neq j}(a_i^2-a_j^2)\prod_j (-a_j^2)$.
\end{description}
By the classification of dual pairs, we only have to consider one
of the following reductive dual pairs:
\begin{enumerate}[(a)]
\item $(\rO(2n),\Sp(2n'))$,
\item $(\Sp(2n), \rO(2n'+1))$ and
\item $(\GL(n),\GL(n'))$.
\end{enumerate}
We may assume that
${\mathrm{rank}} \, G \leq {\mathrm{rank}} \, G'$.
Let $\Wy$ and $\Wy'$ be the Weyl groups of $G$ and $G'$
with respect to $Y$ and $Y'$
respectively. By the first and second fundamental theorems of classical invariant theory $M$ induces an isomorphism $W/G \cong \fgg$ (see \cite{Howe95, GW}) which in turn induces an isomorphism $W/G \times G' \cong \fgg/G'$. We get following
diagram:
\[
\xymatrix{
\fyy \ar@{^(->}[r] \ar[d] & \fgg \ar[d]& \ar@{->>}[l]_{M} W \ar[r]^{M'} \ar[d]& \fgg'\ar[d] &
\fyy' \ar@{_(->}[l] \ar[d] \\
\fyy/\Wy \ar[r]^{\sim}& \fgg/G & \ar[l]_<>(.5){\sim} W/ G \times G'
\ar[r]& \fgg'/G' &\ar@{_(->}[l]_{\sim} \fyy'/\Wy'.
}
\]
We remind the readers that in the lemma below all the vector spaces
and algebraic groups are defined over $\barkk$.
\begin{lemma} \label{LJ1} Suppose
${\mathrm{rank}} \, G \leq {\mathrm{rank}} \, G'$.
Then there is a vector subspace $\AAA$ of $W$ which is stable under
the action of $Y \times Y'$ and such that the following
diagram below commutes.
\begin{equation} \label{eq:cd}
\vcenter{
\xymatrix{ \fyy \ar@{->>}[d] & & \AAA / Y \times Y'
\ar@{_(->>}[ll]_<>(.5){\sim}
\ar@{^(->}[rr] \ar@{->>}[d] & & \fyy' \ar@{->>}[d] \\
\fyy / \Wy \ar[r]^{\sim} & \fgg/ G & W/ G \times G'
\ar@{_(->}[l]_<>(.5){\sim} \ar@{^(->}[r] &
\fgg/ G' & \fyy'/\Wy' \ar[l]_{\sim}.}
}
\end{equation}
\end{lemma}
The proof of the lemma is given in \Cref{app:M0}.
\medskip
The top row of \eqref{eq:cd} defines an inclusion map
$\Upsilon_\fyy \colon \fyy\hookrightarrow
\fyy'$.
This map is the natural inclusion
$\bA^{\rank G} \hookrightarrow \bA^{\rank G'}$ which is well known to
the experts (for example see~\cite{Ad}).
Using the bottom row in \eqref{eq:cd}, we have an inclusion
$\fyy/\Wy \hookrightarrow \fyy'/\Wy'$ induced by $\Upsilon_\fyy$. Let
$P$ and $P'$ be the invariant polynomials for $G$ and $G'$ defined
before \Cref{lem:reg}. Then \eqref{equ17} is equivalent to the
following statement:
\[
\text{There is an $X\in \fyy$ such that $P|_\fyy(X) \neq 0$ and
$P'|_{\fyy'}(\Upsilon_\fyy(X)) \neq 0$.}
\]
It follows by inspection that (i) to (v) are all the possible cases.
\end{proof}
\section{Theta correspondences II} \label{sec:thetaII}
In this section we study theta correspondences of epipelagic representations.
By \Cref{prop:RSP}, \Cref{prop:M.w} and \Cref{cor:JP}, it is enough to consider the following situations:
\begin{enumerate}[(C1)]
\item \label{it:C0} The dual pair $(\bfG,\bfG')$ is one of the
following types:
\begin{inparaenum}[(i)]
\item $(\rD_n,\rC_n)$,
\item $(\rC_n,\rD_{n+1})$,
\item $(\rC_n,\rB_n)$,
\item $(\rA_n,\rA_n)$,
\item $(\rA_n,\rA_{n+1})$.
\end{inparaenum}
\item \label{it:C1} The points $x \in \cB(\bfG,k)$ and $x' \in \cB(\bfG',k)$
are epipelagic points of order $m$. In particular, $m\geq 2$. Let
$\sL$ and~$\sL'$ denote the corresponding $\foo_D$-lattice
functions.
\item \label{it:C2} $\sB = \sL \otimes \sL'$ where
$\Jump(\sB) \subseteq \frac{1}{2m} + \frac{1}{m} \bZ$.
\item \label{it:C3} There exists a
$\bw \in \sfX = \sB_{-\fracdmm} / \sB_{\fracdmm}$ such that
$\sfM(\bw) = \lambda \in \sfg_{x,-\fracmm}$ and
$-\sfM'(\bw) = \lambda' \in \sfg'_{x',-\fracmm}$ are stable vectors.
\end{enumerate}
\subsection{}
We now study the geometry of $\sfX$ and the moment maps which will
eventually determine the local theta correspondences.
By (C4) $\lambda$ is a stable vector so
$\tlambda := \iotag^{-1}(\lambda)$ is a regular semisimple element in
$\sfggE$ (c.f. \Cref{sec:EP}). When
$\tlambda\in \sfggE\subseteq \End_{\fff_\DE}(\sfV)$ is not of full
rank, it requires a special treatment. By (C1) and the classification
of epipelagic points in \cite{RY,Gross}, this is exactly in the
following situation:
\begin{equation}
\label[case]{case:E} \tag{E} \parbox[c][][c]{0.8\textwidth}{
$D$ is a ramified quadratic extension of $k$, the dual pair
$(G, G')$ is a pair of unitary groups of the same rank $n$, and
$\tlambda \in \sfggE$ has rank $n-1$.}
\end{equation}
\def\bSw{{\bS_\bw}}
\def\bSl{{\bS_\lambda}}
Define
\begin{align*}
\sfXll & := \sfM^{-1}(\lambda)\cap \sfM'^{-1}(-\lambda') \text{ and } \\
\sfSw & := \Stab_{\sfSl \times \sfSlp}(\bw).
\end{align*}
We note that $\sfS_\lambda$ is abelian in all our cases so all its
irreducible representations are one dimensional characters.
\begin{lemma} \label{lem:Xll}
\begin{asparaenum}[(i)]
\item \label{lem:Xll.1} The set $\sfXll$
is the $\sfSl$-orbit of $\bw$ in $\sfX$.
\item There is a group homomorphism $\alpha\colon \sfSlp \rightarrow \sfSl$ such
that $\trSp := \set{ (\alpha(g'),g') | g' \in \sfSlp }$ is a subgroup of
$\sfS_{\bw}$.
\item If we are not in \Cref{case:E}, then $\sfSl$ acts freely on
$\sfXll$ and $\sfS_{\bw} = \trSp$.
\item In \Cref{case:E}, let $\bSw = \Stab_{\sfSl}(\bw)$ and
$\bSl := \set{g\in \sfSl| g\circ \lambda = \lambda }$\footnote{Let
$h \in G_x$ and $\gamma\in \fgg_{x,-\fracmm}$ be any lifts of $g$ and
$\lambda$ respectively. We consider $h$ and $\gamma$ as elements in $\Hom_k(V,V)$.
Then $g\circ \lambda := h\circ \gamma + \fgg_{x,-\fracmm^+}\in
\fgg_{x,-\fracmm:-\fracmm^+}$
is well defined.}. Then $\bSw = \bSl$ and so the character $\chi$
of $\sfS_{\lambda}$ occurs in $\bC[\sfXll]$ if and only if
$\chi|_{\bSl}$ is trivial.
\end{asparaenum}
\end{lemma}
The proof is given in \Cref{app:Xll}.
\medskip
\noindent \remark\
\begin{inparaenum}
\item The homomorphism $\alpha$ induces a map
$\alpha^* \colon \widehat{\sfS_{\lambda}} \rightarrow
\widehat{\sfS_{\lambda'}'}$
given by $\alpha^*(\chi) = \chi \circ \alpha$. The definition
depends on the choice of $\bw\in \sfXll$. On the other hand, it is
well-defined up to conjugation by the proof in \Cref{app:Xll}. Hence
the map between the Grothendieck groups induced by $\alpha^*$ is
independent of the choice of $\bw$.
\item In the exceptional \Cref{case:E}, \Cref{lem:Xll}~(iv) will lead to the
fact that not all epipelagic representations can occur in this local theta
correspondence. The extreme case is the well known fact that not all
characters of $\rU(1)$ occur in the oscillator representation
of~$\Mp(2)$(see~\cite{Moen93}).
\end{inparaenum}
\begin{thm}\label{thm:main0}
Suppose (C1) to (C4) hold.
For any
character $\chi$ of $\sfS_{\lambda}$, let $\Sigma = (x,\lambda, \chi)$ and
$\Sigma' = (x',\lambda', \chi')$ where $\chi' = \chi^*\circ \alpha$ and
$\chi^*$ is the contragredient representation of $\chi$.
\begin{enumerate}[(i)]
\item Suppose we are not in the exceptional \cref{case:E}. Then
\begin{equation} \label{eq:CEpi}
\theta( \piSigma) = \piSigmap.
\end{equation}
In particular the theta lift is nonzero.
\item Suppose we are in the exceptional \Cref{case:E}. Then \eqref{eq:CEpi}
holds for $\chi \in \widehat{\sfSl}$ such that $\chi|_{\bSl}$ is
trivial.
\end{enumerate}
\end{thm}
The above theorem gives \Cref{thm:intro} (i).
\subsection{} \label{sec:sfX}
We set $B = \sB_{-\fracdmm}$ and $A = \sB_{\fracdmm}$. By (C3),
$A^\sharp = A$ and $B^\sharp = \sB_{\fracdmm^+} = \sB_{\fractdmm}$.
Since $m\geq 2$, $\sB_{-\fracdmm}\fpp_k \subseteq \sB_{\fractdmm}$ and
$B^\sharp$ is a good lattice. We form the following exact sequences of $\fff$-vector
spaces:
\begin{equation} \label{equ9}
\xymatrix{ 0\ar[r]& \sfY \ar[r]& \sfW \ar[r] & \sfX \ar[r] & 0}
\end{equation}
where
\[
\sfW = B/B^\sharp = \sB_{-\fracdmm:\fractdmm},\ \
\sfX = B/A = \sB_{-\fracdmm:\fracdmm} \text{ \ and \ }
\sfY = A/B^\sharp = \sB_{\fracdmm:\fractdmm}.
\]
The symplectic form on $W$ induces a non-degenerate $\fff$-symplectic form on $\sfW$ and
$\sfY$ is a maximal isotropic subspace in $\sfW$. Let
\begin{align*}
\sfP & = (G_{\sL}/G_{\sL,\frac{2}{m}}) \times
(G'_{\sL'}/G'_{\sL',\frac{2}{m}}) \text{ and } \\
\sfJ & = (G_{\sL}/G_{\sL,\fracmm}) \times
(G'_{\sL'}/G'_{\sL',\fracmm})=\sfGx\times \sfGxp.
\end{align*}
Then \eqref{equ9} is an exact sequence of $\sfP$-modules.
The proof of following lemma is given in \Cref{app:split}.
\begin{lemma} \label{lem:split}
The natural quotient $\sfP \twoheadrightarrow \sfJ$ has a splitting such that
the exact sequence \eqref{equ9} splits as
$\sfJ$-modules. We denote the splitting by $\sfW = \sfY\oplus \sfX$.
\end{lemma}
\subsection{Proof of \Cref{thm:main0}}
We recall the lattice model $\sS(A)$ in \Cref{sec:LM}. Let $\sS(A)_B$
be the subspace of functions in $\sS(A)$ with support in $B$. For
$f\in \sS(A)_B$, $w\in B$ and $b'\in B^\sharp$,
$f(w+b') = \psi(\half\inn{w}{b'}) f(w) = f(w)$. Therefore, we could
view $\sS(A)_B$ as a subspace in $\bC[\sfW]$. We fix the splitting
$\sfW = \sfY\oplus \sfX$ in \Cref{lem:split}. Let $\sfJ$ act on
$\bC[\sfX]$ by translation. Since
$f(w+a) = \psi(\frac{1}{2}\inn{w}{a})f(w)$ for all $a \in A$, the
restriction map $R_\sfX$ from $\sfW$ to $\sfX$ induces a $\sfJ$-module
isomorphism
\begin{equation} \label{eq:RX}
\xymatrix{
R_{\sfX} : \sS(A)_B \ar[r]^{\ \ \ \ \ \sim} & \bC[\sfX]
}
\end{equation}
whose inverse map $R_\sfX^{-1}$ is given by
\[
(R_\sfX^{-1} F)(w) = \psi \left( \half \inn{x}{y} \right) F(x)
\]
for all $w\in B$ such that $w \equiv x+y \pmod{B^\sharp}$ with
$x + B^\sharp \in \sfX$ and $y + B^\sharp \in \sfY$.
\medskip
For $f\in \sS(A)_B$, $w\in B$ and $g\in G_{\sL,\fracmm}$, we have $(g^{-1}-1)w \in
A$. By \eqref{eq:LA} and \Cref{lem:MPsi}, we have\footnote{See also the proof of \cite[Theorem~5.5]{Pan}}
\begin{equation} \label{eq:gf}
\begin{split}
\omegaA(g) f(w) & = f(g^{-1} w)
= f((g^{-1}-1)w+ w) \\
& = \psi\left(\half\inn{w}{(g^{-1}-1)w} \right)f(w) =
\psi\left(\frac{1}{2}\inn{(g-1)w}{w} \right)f(w) \\
& = \psi(\Bg(M(w),c(g)))f(w) = \psi_{\sfM(\bw)}(g)f(w)
\end{split}
\end{equation}
where $\bw$ is the image of $w$ in $\sfX = B/A$.
Similarly $\omegaA(g')f(w) = \psi_{-\sfM'(\bw)}(g')f(w)$ for all $g'\in G_{\sL,\fracmm}$.
Let $\sS(A)_B^\lambda$ be the
subspace of functions in $\sS(A)_B$ such that $G_{\sL,\fracmm}$ acts by
$\psi_\lambda$. Then it follows from~\eqref{eq:gf} and \eqref{eq:RX} that
\[
\sS(A)_B^\lambda = R_\sfX^{-1}(\bC[\sfM^{-1}(\lambda)]).
\]
A similar consideration applies to $\lambda' \in \sfg_{\sL',-\fracmm}'$ too.
Let
$\sS^{\lambda, \lambda'} := \sS(A)_B^{\lambda} \cap
\sS(A)_B^{\lambda'}$. Then
\[
\sS^{\lambda,\lambda'} = R_\sfX^{-1}(\bC[\Xllp]).
\]
We fix a $\barww$ in $\Xllp$. By \Cref{lem:Xll}, $\sfSl \twoheadrightarrow
\sfXll$ given by $s\mapsto s\cdot \bw$ is a surjection
of $\sfSl\times \sfSlp$-set.
Here $(s,s') \in \sfSl\times \sfSlp$ acts on $\sfSl$ by $(s,s') \cdot s_0 = s s_0 \alpha(s')^{-1}$ for all $s_0\in \sfSl$.
By the decomposition of regular representation of $\sfS_{\lambda}$, we have
\begin{equation}\label{eq:dec}
\bC[\Xllp] \subseteq \bC[\sfSl]= \bigoplus_{\chi\in \widehat{\sfS_{\lambda}}}
\bC_{\chi}\otimes \bC_{\chi^*\circ \alpha}
\end{equation}
as $\sfS_{\lambda}\times \sfS'_{\lambda'}$-modules. In (i),
\eqref{eq:dec} is an equality. In (ii) the summand
$\bC_{\chi}\otimes \bC_{\chi^*\circ \alpha}$ occurs in $\bC[\sfXll]$
if and only if $\chi|_{\bSl}$ is trivial by \Cref{lem:Xll}.
Fix any $\sfS_\lambda$-character $\chi$ which occurs in $\bC[\sfXll]$. It is
clear that the $H_{\lambda}$ and $H'_{\lambda'}$ in
\eqref{eq1} act on the space
$R_\sfX^{-1}(\bC_{\lambda}\otimes \bC_{\chi^*\circ \alpha})$ by the
characters $\psi_\lambda \otimes \chi$ and
$\psi_{\lambda'} \otimes \chi^*\circ \alpha$ respectively.
By Frobenius reciprocity, the subspace $R_\sfX^{-1}(\bC_{\lambda}\otimes \bC_{\chi^*\circ \alpha}) \subseteq \sS(A)$
induces a non-zero intertwining map
\[
\xymatrix{
\pi_x^{\tG}(\lambda,\chi) \boxtimes
\pi_{x'}^{\tG'}(\lambda',\chi^*\circ \alpha) \ar[r]& \sS(A).}
\]
The left hand side is irreducible so the above
map is an injection. Since the left hand side is also supercuspidal, by the smoothness of $\sS(A)$, we conclude that the
left hand side is a direct summand in $\sS(A)$ and we have a
projection map from $\sS(A)$ to the left hand side.
This completes the proof of \Cref{thm:main0}. \qed
\subsection{Proof of \Cref{thm:intro}~(ii)} \label{sec:Converse}
Part (a) is a restatement of \Cref{prop:M.w}. By \Cref{thm:intro} $\piSigma$ has a nonzero theta lift and $\theta(\piSigma) = \pi'_{\Sigma''}$ where $\Sigma'' = (x',\lambda',\chi^* \circ \alpha)$. By the uniqueness of the theta lift \cite{GT,Wa}, $\piSigmap = \pi'_{\Sigma''}$ and by \Cref{prop:MKT} $\chi' = \chi^* \circ \alpha$. This proves (b).
\qed
|
\section{Introduction}
Let us consider an $m$-dimensional Riemannian manifold $ ( \mc{M}^m, g ^m )$ isometrically immersed in a Riemannian manifold $ ( \bar \mc{M}^n , \bar g ^ n )$. The fundamental objects describing the \emph{intrinsic} geometry of $( \mc{M}^m, g^m )$ are the metric $ g^m$, the curvature tensor $ R _{ ijkl} $, and the Levi-Civita connection. On the other hand, the fundamental quantities describing the \emph{extrinsic} geometry of $ ( \mc{M}^m, g^m) $ as submanifold of $ ( \bar \mc{M}^n , \bar g ^ n ) $ are the second fundamental form $ h _{ij } ^ \alpha $, the normal connection $ \nabla ^ \perp$, and the normal curvature $\bar R ^ \perp_{ ij\alpha\beta} $, where Roman indices indicate tangential directions and Greek indices indicate normal ambient directions. It is well known (see Section \ref{sec:BM} for more details) that these geometric quantities are not mutually independent but must satisfy some compatibility conditions, the so called Gauss-Codazzi-Mainardi-Ricci equations. A natural way to define \emph{geometric scalars} out of this list of tensors is by taking tensor products and then contracting using the metric $\bar{g}$. More precisely, we first take a finite number of tensor products, say
\begin{equation}\nonumber
R_{i_1 j_1 k_1 l_1} \otimes \ldots \otimes R^\perp_{i_r, j_r, \alpha_r, \beta_r} \otimes \ldots \otimes h_{i_s j_s} \quad,
\end{equation}
thus obtaining a tensor of rank $4+\ldots+4+\ldots+2+\ldots+2$. Then, we repeatedly pick out pairs of indices in the above expression and contract them against each other using the metric $\bar{g}^{\alpha \beta}$ (of course, in case of contractions not including the normal curvature $R^\perp$ it is enough to contract using $g^{ij}$). This can be viewed in the more abstract perspective of Definition \ref{def:ComplContr} by saying that we consider a \emph{geometric complete contraction}
\begin{equation}\nonumber \label{eq:ContrGen0}
C(\bar{g},R,R^\perp,h) = \contr(\bar{g}^{\alpha_1 \beta_1}\otimes \ldots \otimes R_{i_1 j_1 k_1 l_1} \otimes \ldots \otimes R^\perp_{i_r, j_r, \alpha_r, \beta_r} \otimes \ldots \otimes h_{i_s j_s}) \quad.
\end{equation}
Let us stress that a complete contraction is determined by the \emph{pattern} according to which different indices contract agains each other; for example, the complete contraction $R_{ijkl}\otimes R^{ijkl}$ is different from $R^i_{ikl} \otimes R_{s}^{ksl}$. By taking linear combinations of geometric complete contractions (for the rigorous meaning see Definition \ref{def:LinCombCC}), we construct \emph{geometric scalar quantities}
\begin{equation}\nonumber \label{eq:PgRh0}
P(\bar{g},R, R^\perp, h):= \sum_{ l \in L } a _ l C ^l (\bar{g},R, R^\perp, h) \quad.
\end{equation}
The goal of the present paper is to classify those geometric scalar quantities which, once integrated over arbitrary submanifolds $(\mc{M}^m, g^m)$ of arbitrary manifolds $(\bar{\mc{M}}^n,\bar{g}^n)$, give rise to \emph{global conformal invariants}. More precisely, we say that the geometric scalar quantity $P(\bar{g},R, R^\perp, h)$ is a \emph{global conformal invariant for $m$-submanifolds in $n$-manifolds} if the following holds: for any ambient Riemannian manifold $\bar{\mc{M}}^n$, any compact orientable $m$-dimensional immersed submanifold $\mc{M}^m$ of $\bar{\mc{M}}^n$ and any $\phi \in C^\infty(\bar{\mc{M}})$, if one considers the conformal deformation $\hat{\bar{g}}:= e ^{2 \phi(x)} \bar{g}$ and calls $\hat{R}, \hat{R}^\perp, \hat{h}$ the tensors computed with respect to the conformal metric $\hat{\bar{g}}$, then
\begin{equation}\nonumber \label{eq:ConInv0}
\int_{\mc{M}^m} P(\hat{\bar{g}},\hat{R}, \hat{R}^\perp, \hat{h}) \, d \mu_{\hat{g}}= \int_{\mc{M}^m} P(\bar{g},R, R^\perp, h)\, d \mu_g \quad.
\end{equation}
Let us mention that the corresponding classification for \emph{intrinsic} global conformal invariants of Riemannian manifolds was a classical problem in conformal geometry motivated also by theoretical physics (the goal being to understand the so called conformal anomalies): indeed it is the celebrated Deser-Schwimmer conjecture \cite{DeserSchwimmer} which has recently been solved in a series of works by Alexakis \cite{AlexI,AlexII,AlexPf1, AlexPf2, AlexIV, AlexBook}. Inspired by the aforementioned papers, we address the problem of an analogous classification for global conformal invariants, but this time, \emph{for submanifolds}. Of course, as explained above, if one considers global conformal invariants for \emph{submanifolds} many other \emph{extrinsic} terms appears, namely the second fundamental form, the curvature of the normal bundle, and the normal connection; therefore the zoology of global conformal invariants is more rich and the classification more complicated.
\medskip
A well-known example of a global conformal invariant for two-dimensional submanifolds (called from now on surfaces) is the \emph{Willmore energy}. For an immersed surface $ f : \mc{M} ^ 2 \rightarrow ( \bar \mc{M}^ n , \bar g^n ) $ this is defined by
\begin{equation}\label{eq:defW}
\mathcal {W} ( f ) = \int _{ \mc{M} } | H | ^ 2 d \mu _ g + \int _{ \mc{M} } \bar {K}_{ \bar \mc{M}} (Tf(\mc{M})) d \mu _g \quad,
\end{equation}
where $H= \frac{1}{2} g^{ij} h_{ij}$ is the mean curvature vector and $\bar {K}_{ \bar \mc{M}} (Tf(\mc{M}))$ is the sectional curvature of the ambient manifold computed on the tangent space of $f(\mc{M})$. Clearly, in case $\bar{\mc{M}}^n=\mathbb{R}^n$, \eqref{eq:defW} reduces to the familiar Euclidean Willmore energy as $ \bar K _{ \mathbb{R} ^ n } = 0$. It is well known that the Willmore energy in Euclidean space is invariant under conformal transformations of the ambient manifolds, that is M\"obius transformations where the inversion is centered off the submanifold. In fact, more generally the conformal Willmore energy is invariant under conformal deformations of the ambient background metric. This can be seen by the following decomposition,
\begin{align*}
\|h^\circ \| ^ 2 = \| h \| ^ 2 - 2 |H| ^ 2, \quad K_{\mc{M}} = \frac{1}{2} (4 |H| ^ 2 - \| h \| ^ 2 ) + \bar K _{ \bar \mc{M} },
\end{align*}
where $h^\circ_{ij}:=h_{ij}-H g_{ij}$ is the traceless second fundamental form, $K_{\mc{M}}$ is the Gauss curvature of $(\mc{M},g)$, and in the second identity we just recalled the classical Gauss equation.
It follows that the conformal Willmore energy can be written as
\begin{align*}
|H| ^ 2 + \bar K_{\bar \mc{M} } = \frac 12 \| h^\circ \| ^ 2 + K_\mc{M}.
\end{align*}
Since $ \| h^\circ \| ^ 2 d \mu_g $ is a pointwise conformal invariant and $ \int _{\mc{M}} K_{\mc{M}} d \mu_ g= 2 \pi \chi(\mc{M})$ is a topological (hence, a fortiori, global conformal) invariant by the Gauss-Bonnet theorem, clearly any linear combination of the two is a global conformal invariant. A natural question is whether the Willmore functional is the unique energy having such an invariance property, up to topological terms.
Let us briefly mention that the Willmore energy has recently received much attention \cite{Blaschke, CaMo, BeRi, KMS, KS, LMS, LY, Mon2, MR2, MontielUrbano, Riv, Sim, Will}, and in particular the Willmore conjecture in codimension one has been solved \cite{MN}.
\medskip
We will show that, for codimension one surfaces, any global conformal invariant of a surface must be a linear combination of the norm squared of the traceless second fundamental form and the intrinsic Gauss curvature, that is the Willmore energy is the unique global conformal invariant up to the Gauss-Bonnet integrand which is a topological quantity (see Theorem \ref{thm:CD1surf}).
A similar statement holds also for codimension two surfaces (see Theorem \ref{thm:CD2surf}), once taking into account an additional topological term given by the Chern-Gauss-Bonnet integrand of the normal bundle.
For general submanifolds of codimension one, we show that if the global conformal invariant is a polynomial in the second fundamental form only, then it must be a contraction of the \emph{traceless} second fundamental form, that is it must be the integral of a \emph{pointwise} conformal invariant (see Theorem \ref{thm:P(g,h)}). Combining this with a theorem of Alexakis \cite[Theorem 1]{AlexI}, we show that if $m$ is even and the global conformal invariant has no mixed contractions between the intrinsic and the extrinsic curvatures then the integrand must be a linear combination of contractions of the (intrinsic) Weyl curvature, contractions of the traceless second fundamental form and the integrand of the Chern-Gauss-Bonnet formula, see Theorem \ref{thm:P=P1+P2}.
As an application of these ideas, in the last Section \ref{sec:genWill}, we introduce a higher dimensional analogue of the Willmore energy for hypersurfaces in Euclidean spaces. Such new energies are conformally invariant and attain the strictly positive lower bound only at round spheres, with rigidity, regardless of the topology of the hypersurface (see Theorem \ref{thm:3} and Theorem \ref{thm:4}).
\medskip
\begin{center}
{\bf Acknowledgments}
\\This work was written while the first author was visiting the {\it Forschungsinstitut f\"ur Mathematik} at the ETH Z\"urich. He would like to thank the Institut for the hospitality and the excellent working conditions.\\ The second author is supported by the ETH fellowship.
\end{center}
\medskip
\section{Background Material}\label{sec:BM}
\subsection{Complete contractions: abstract definition}
Following \cite{AlexI}, in this short section we define the notion of complete contractions.
\begin{definition}[Complete Contractions]\label{def:ComplContr}
Any complete contraction
\begin{align*}
C = \contr( ( A ^ 1 ) _{ i_ 1\dots i_s} \otimes ( A ^ t ) _{ j_1 \dots j _ q })
\end{align*}
will be seen as a formal expression. Each factor $ ( A ^ l ) _{ i _ 1\dots i_s } $ is an ordered set of slots. Given the factors $ A ^ 1 _{ i _1\dots i _ s }, A ^ l _{ j_1\dots j _ q } $ a complete contraction is then a set of pairs of slots $ ( a _1, b_1 ), \dots, (a _ w, b _ w ) $ with the following properties:
if $ k \neq l , \{a _ l, b _ l \}\cap \{ a _k ,b _k \} = \emptyset $, $ a _k \neq b _k $ and $ \bigcup_{ i = 1 } ^ { w } \{ a _i, b _i \} = \{ i _ 1 , \dots, j_q \}.$ Each pair corresponds to a particular contraction.
Two complete contractions
\begin{align*}
\contr(( A^1 ) _{ i_1\dots i _ s } \otimes \dots \otimes ( A ^ t ) _ { j _1\dots j _w} )
\end{align*}
and
\begin{align*}
\contr( (B^1)_{f _1\dots f _ q} \otimes \dots \otimes ( B ^ {t'} )_{v _ 1\dots v _ z } )
\end{align*}
will be identical if $ t = t ', A ^ l = B ^ l$ and if the $ \mu$-th index in $ A ^ l$ contracts against the $ \nu$-th index in $ A^r$ then the $\mu$-th index in $ B^ l $ contracts against the $\nu$-th in $ B ^ r $.
For a complete contraction, the \emph{length} refers to the number of factors.
\end{definition}
\begin{definition}[Linear combinations of complete contractions]\label{def:LinCombCC}
Linear combinations of complete contractions are defined as expressions of the form
\begin{align*}
\sum_{ l \in L } a _ l C ^l_1, \quad \sum_{ r \in R } b _r C_2 ^ r \quad,
\end{align*}
where each $ C ^ l _i$ is a complete contraction.
Two linear combinations are identical if $ R = L$, $ a _ l = b _l $ and $ C_1 ^ l = C_2 ^ l$. A linear combination of complete contractions is identically zero if for all $ l \in L $ we have $ a _l =0 $. For any complete contraction, we will say that a factor $ ( A)_{ r _1\dots r _{s_l} } $ has an internal contraction if two indices in
\begin{align*}
A _{ r_1\dots r _ {s_l} }
\end{align*}
are contracting amongst themselves.
\end{definition}
\subsection{Riemannian and Submanifold Geometry} \label{SSS:RSG}
Consider an $n$-dimensional Riemannian manifold $(\bar{\mc{M}} ^ n ,\bar{g} ^ n )$. Given $ x _ 0 \in \bar{\mc{M}}^ n$, let $ ( x ^1, \dots , x ^ n)$ be a local coordinate system with associated coordinate vector fields denoted by $ X^ \alpha $, that is $ X ^ \alpha = \frac { \partial }{ \partial x ^ \alpha } $. Called $\bar{\nabla}$ the Levi-Civita connection associated to $(\bar{\mc{M}} ^ n ,\bar{g} ^ n )$, the covariant derivative $ \bar{\nabla} _{ \frac { \partial }{\partial x ^ \alpha} }$ will be shortly denoted by $\bar{\nabla}_\alpha$.
The curvature tensor $ \bar{R} _{ \alpha \beta \gamma \eta } $ of $ \bar{g} ^ {n} _{ \alpha \beta } $ is given by the commutator of the covariant derivatives, that is
\begin{align}\label{eq:defR}
[ \bar{\nabla} _\alpha \bar{\nabla} _\beta - \bar{\nabla} _\beta \bar{\nabla} _\alpha ] X_\gamma = \bar{R} _{ \alpha\beta \gamma \eta} X ^ \eta \quad,
\end{align}
which in terms of coordinate systems may be expressed by Christoffel symbols,
\begin{align*}
\bar{R} _{ \alpha \beta \gamma} ^ \eta = \partial_{ \beta } \Gamma _{ \alpha \gamma } ^ \eta - \partial _\gamma \Gamma ^ \eta _{ \alpha \beta } + \sum_{ \mu } (\Gamma_{ \alpha \gamma } ^ \mu \Gamma _{ \mu \beta } ^ \eta -\Gamma _{ \alpha \beta } ^ \mu \Gamma _{ \mu \gamma } ^ \eta ).
\end{align*}
The two Bianchi identities are then
\begin{align*}
&\bar{R}_{ \alpha \beta \gamma\eta} + \bar{R} _{ \gamma\alpha \beta \eta } + \bar{R} _{\beta \gamma\alpha \eta } = 0\\
&\bar{\nabla} _ \alpha \bar{R} _{ \beta \gamma\eta \mu } + \bar{\nabla} _\gamma \bar{R}_{\alpha \beta \eta \mu } + \bar{\nabla} _{ \beta } \bar{R} _{ \gamma\alpha \eta \mu } = 0.
\end{align*}
Recall that, under a conformal change of metric $\hat{\bar{g}}^n= e ^{2 \phi (x)} \bar{g}^n(x)$, the curvature transforms as follows (see for instance \cite{Eastwood}):
\begin{eqnarray}
\bar{R}^{\hat{g}^n}_{\alpha \beta \gamma\eta }&=& e ^{2\phi (x)} \big[\bar{R}^{\bar{g}^n}_{\alpha \beta \gamma\eta }+\bar{\nabla}_{\alpha \eta }\phi \bar{g}^n_{\beta \gamma}+ \bar{\nabla}_{\beta \gamma} \phi \bar{g}^n_{\alpha \eta }-\bar{\nabla}_{\alpha \gamma}\phi \bar{g}^n_{\beta \eta }- \bar{\nabla}_{\beta \eta } \phi \bar{g}^n_{\alpha \gamma} \nonumber \\
&& \quad \quad \quad+ \bar{\nabla}_\alpha \phi \bar{\nabla}_\gamma \phi \bar{g}^n_{\beta \eta } + \bar{\nabla}_\beta \phi \bar{\nabla}_\eta \phi \bar{g}^n_{\alpha \gamma} - \bar{\nabla}_\alpha \phi \bar{\nabla}_\eta \phi \bar{g}^n_{\beta \gamma} - \bar{\nabla}_\beta \phi \bar{\nabla}_\gamma \phi \bar{g}^n_{\alpha \eta } \nonumber \\
&&\quad \quad \quad + |\bar{\nabla} \phi|^2 \bar{g}^n_{\alpha \eta } \bar{g}^n_{\beta \gamma} - |\bar{\nabla} \phi |^2 \bar{g}^n_{\alpha \gamma} g_{\eta \beta } \big] \quad .\label{eq:Rhatg}
\end{eqnarray}
Now let briefly introduce some basic notions of submanifold geometry. Given an $m$-dimensional manifold $\mc{M}^m$, $2\leq m <n$, we consider $f:\mc{M}^m \hookrightarrow \bar{\mc{M}}^n$, a smooth immersion. Recall that for every fixed $\bar{x} \in \mc{M}^m$ one can find local coordinates $(x^1, \ldots , x^n)$ of $\bar{\mc{M}}^n$ on a neighborhood $V_{f(\bar{x})}^{\bar{\mc{M}}}$ of $f(\bar{x})$ such that $\big((x^1\circ f), \ldots, (x^m\circ f) \big)$ are local coordinates on a neighborhood $U_{\bar{x}}^{\mc{M}}$ of $\bar{x}$ in $\mc{M}^m$ and such that
$$f\big(U_{\bar{x}}^{\mc{M}}\big)=\Big\{(x^1, \ldots, x^n)\in V_{f(\bar{x})}^{\bar{\mc{M}}} \; : \; x^{m+1}=\ldots =x^n=0\Big \} \quad. $$
Such local coordinates on $\bar{\mc{M}}$ are said to be \emph{adapted} to $f(\mc{M})$. We use the convention that latin index letters vary from $1$ to $m$ and refer to geometric quantities on $\mc{M}^m$, while greek index letters vary from $1$ to $n$ (or sometimes from $m+1$ to $n$ if otherwise specified) and denote quantities in the ambient manifold $\bar{\mc{M}}^n$ (or in the orthogonal space to $f(\mc{M})$ respectively).
In adapted coordinates, it is clear that $X_1, \ldots, X_m$ are a bases for the tangent space of $f(\mc{M}^m)$ and that the restriction of the ambient metric $\bar{g}^n$ defines an induced metric on $\mc{M}^m$, given locally by
$$g^m_{i j}:= \bar{g}^n(X_i, X_j)\quad .$$
Using standard notation, $(g^m)^{ij}$ denotes the inverse of the matrix $(g_m)_{ij}$, that is $(g^m)^{ik} (g^m)_{kj}=\delta_{ij}$.
For every $\bar{x} \in \mc{M}$, the ambient metric $\bar{g}^n$ induces the orthogonal splitting
$$T_{f(\bar{x})} \bar{\mc{M}}= T_{f(\bar{x})} f(\mc{M}) \oplus [T_{f(\bar{x})} f(\mc{M})]^\perp \quad, $$
where, of course, $ [T_{f(\bar{x})} f(\mc{M})]^\perp$ is the orthogonal complement of the $m$-dimensional subspace $T_{f(\bar{x})} f(\mc{M}) \subset T_{f(\bar{x})} \bar{\mc{M}}$.
We call $\pi_T: T_{f(\bar{x})} \bar{\mc{M}} \to T_{f(\bar{x})} f(\mc{M})$ and $\pi_N=Id-\pi_T: T_{f(\bar{x})} \bar{\mc{M}} \to [T_{f(\bar{x})} f(\mc{M})]^\perp$ the tangential and the normal projections respectively, one can define the \emph{tangential} and the \emph{normal connections} (which correspond to the Levi-Civita connections on $(\mc{M}, g)$ and on the normal bundle respectively) by
\begin{eqnarray}\label{eq:tgNConn}
\nabla_{X_i} X_j&:=&\pi_T(\bar{\nabla}_{X_i} X_j), \; i,j=1,\ldots,m, \label{eq:tgConn} \\
\nabla^\perp _{X_i} X_{\alpha} &:=&\pi_N(\bar{\nabla}_{X_i} X_\alpha), \; i=1,\ldots,m, \alpha=m+1, \ldots, n. \label{eq:NConn}
\end{eqnarray}
Associated to the tangential and normal connections we have the tangential and normal Riemann curvature tensors (which correspond to the curvature of $(\mc{M}, g)$ and of the normal bundle respectively) defined analogously to \eqref{eq:defR}:
\begin{eqnarray}
[ \nabla _i \nabla _j - \nabla _j \nabla _i ] X_k &= &R _{ i j k l} X ^ l \;, \quad i,j,k,l=1,\ldots, m \label{eq:defRt} \\
\left[{\nabla^\perp_i \nabla^\perp_j - \nabla^\perp_j \nabla^\perp_i} \right] X_{\alpha} &= &R^\perp _{ i j \alpha \beta} X ^ \beta \;, \quad i,j=1,\ldots, m, \; \alpha, \beta=m+1, \ldots, n. \label{eq:defRn}
\end{eqnarray}
The transformation of $R_{ijkl}$ and $R^\perp_{ij\alpha \beta}$ under a conformal change of metric is analogous to \eqref{eq:Rhatg}, just replacing $\bar{\nabla}$ with $\nabla$ or with $\nabla^{\perp}$ respectively.
The second fundamental form $h$ of $f$ is defined by
\begin{equation}\label{eq:defh}
h(X_i, X_j):= \pi_{N} (\bar{\nabla}_{ X_i} X_j)= \bar{\nabla}_{X_i} X_j- \nabla_{X_i} X_j \quad .
\end{equation}
It can be decomposed orthogonally into its trace part, the \emph{mean curvature}
\begin{equation}\label{eq:defH}
H:=\frac{1}{m} (g^m)^{ij} h_{ij} \quad,
\end{equation}
and its trace free part, the \emph{traceless second fundamental form}
\begin{equation}\label{eq:hH}
h^\circ_{ij}:=h_{ij}-H g^m_{ij} \quad,
\end{equation}
indeed it is clear from the definitions that
\begin{equation}\label{eq:Splith}
h_{ij}= h_{ij}^\circ + H g^m_{ij} \quad \text{ and } \quad \langle h^\circ, Hg^m \rangle = (g^m)^{ik} (g^m)^{jl} \; h^\circ_{ij} \; H g_{kl}=H \; \Tr_{g^m} (h^\circ) =0 \quad.
\end{equation}
Under a conformal change of the ambient metric $\hat{\bar{g}}^n= e ^{2 \phi (x)} \bar{g}^n(x)$, the above quantities change as follow:
\begin{equation}\label{eq:CChH}
h^{\hat{g}}_{ij}= e^{\phi} \big[h_{ij}-g^{m}_{ij}\; \pi_{N} (\bar{\nabla} \phi) \big], \quad H^{\hat{g}}=e^{- \phi} \left[ H - \pi_{N} (\bar{\nabla} \phi) \right] \quad\text{and}\quad (h^{\hat{g}})^{\circ}_{ij}= e^{\phi}\; h^{\circ}_{ij} \quad.
\end{equation}
Observe that, in particular, the endomorphism of $T f(\mc{M})$ associated to $h^{\circ}$ is invariant under conformal deformation of the ambient metric, that is
\begin{equation}
[(h^{\hat{g}})^{\circ}]^i_j= [h^{\circ}]^i_j \quad.
\end{equation}
Finally let us recall the fundamental equations of Gauss and Ricci which link the ambient curvature $\bar{R}$ computed on $T f (\mc{M})$ (respectively on $T f (\mc{M})^{\perp}$) with the second fundamental form and the intrinsic curvature $R$ (respectively with the second fundamental form and the normal curvature $R^\perp$):
\begin{eqnarray}
\bar{R}_{ijkl}&=&R_{ijkl}+(h_{il})_{\alpha} \; (h_{jk})^{\alpha}- (h_{ik})_{\alpha} \; (h_{jl})^{\alpha} \quad, \label{eq:Gauss} \\
\bar{R}_{ij \alpha \beta}&=&R_{ij\alpha \beta}^\perp + (h_{ik})_\alpha \; (h_j^k)_{\beta}-(h_{ik})_{\beta} \; (h_j^k)_{\alpha} \quad, \label{eq:Ricci}
\end{eqnarray}
where, of course, $(h_{il})^{\alpha}$ denotes the $\alpha$-component of the vector $h_{il}\in T f (\mc{M})^{\perp}\subset T \bar{\mc{M}}$ and where we adopted Einstein's convention on summation of repeated indices.
\subsection{Geometric complete contractions and global conformal Invariants of Submanifolds}
Given an immersed submanifold $f:\mc{M}^m\hookrightarrow \bar{\mc{M}}^n$, of course the above defined Riemannian curvature $R$, the curvature of the normal bundle $R^\perp$, and the extrinsic curvatures $h, h^\circ, H$ are geometric objects, that is they are invariant under change of coordinates and under isometries of the ambient manifold. So they give a list of \emph{geometric tensors}. A natural way to define \emph{geometric scalars} out of this list of tensors is by taking tensor products and then contracting using the metric $\bar{g}$. More precisely, we first take a finite number of tensor products, say,
\begin{equation}
R_{i_1 j_1 k_1 l_1} \otimes \ldots \otimes R^\perp_{i_r, j_r, \alpha_r, \beta_r} \otimes \ldots \otimes h^{\circ}_{i_s j_s} \otimes \ldots \otimes H g_{i_t j_t} \quad,
\end{equation}
thus obtaining a tensor of rank $4+\ldots+4+\ldots+2+\ldots+2$. Then, we repeatedly pick out pairs of indices in the above expression and contract them against each other using the metric $\bar{g}^{\alpha \beta}$ (of course, in case of contractions not including the normal curvature $R^\perp$ it is enough to contract using $g^{ij}$). This can be viewed in the more abstract perspective of Definition \ref{def:ComplContr} by saying that we consider a \emph{geometric complete contraction}
\begin{equation}\label{eq:ContrGen}
C(\bar{g},R,R^\perp,h) = \contr(\bar{g}^{\alpha_1 \beta_1}\otimes \ldots \otimes R_{i_1 j_1 k_1 l_1} \otimes \ldots \otimes R^\perp_{i_r, j_r, \alpha_r, \beta_r} \otimes \ldots \otimes h^{\circ}_{i_s j_s} \otimes \ldots \otimes H g_{i_t j_t} ) \quad.
\end{equation}
Let us stress that a complete contraction is determined by the \emph{pattern} according to which different indices contract agains each other; for example, the complete contraction $R_{ijkl}\otimes R^{ijkl}$ is different from $R^i_{ikl} \otimes R_{s}^{ksl}$. By taking linear combinations of geometric complete contractions (for the rigorous meaning see Definition \ref{def:LinCombCC}), we construct \emph{geometric scalar quantities}
\begin{equation}\label{eq:PgRh}
P(\bar{g},R, R^\perp, h):= \sum_{ l \in L } a _ l C ^l (\bar{g},R, R^\perp, h) \quad.
\end{equation}
\begin{remark}
Notice that thanks to the Gauss \eqref{eq:Gauss} and Ricci \eqref{eq:Ricci} equations, one can express the ambient curvature $\bar{R}$ restricted on the tangent space of $\mc{M}$ or restricted to the normal bundle (that is $\bar{R}_{ijkl}$ and $\bar{R}_{ij\alpha \beta}$) as quadratic combination of $R, R^\perp$ and $h$. This is the reason why we can assume it is not present in the complete contractions \eqref{eq:ContrGen} without losing generality. Analogously, thanks to \eqref{eq:hH}, we can assume that $h$ is not present in the complete contractions but just $h^\circ$ and $H g$.
\end{remark}
\begin{definition}[Weight of a geometric scalar quantity]
Let $P(\bar{g},R, R^\perp, h)$ be a geometric scalar quantity as in \eqref{eq:PgRh} and consider the homothetic rescaling $\bar{g} \mapsto t^2 \bar{g}$ of the ambient metric $\bar{g}$, for $t \in \mathbb{R}_{+}$. By denoting $R_{t^2\bar{g}}, R^\perp_{t^2\bar{g}}, h_{t^2\bar{g}}$ the tensors computed with respect to the rescaled metric $t^2 \bar{g}$, if
$$P(t^2\bar{g},R_{t^2\bar{g}}, R^\perp_{t^2\bar{g}}, h_{t^2\bar{g}})= t^K P(\bar{g},R_{\bar{g}}, R^\perp_{\bar{g}}, h_{\bar{g}}), \quad \text{for some } K \in \mathbb Z \quad,$$
we then say that $P(\bar{g},R, R^\perp, h)$ is a geometric scalar quantity of \emph{weight $K$}.
\end{definition}
Recall that, under a general conformal deformation $\hat{\bar{g}}= e ^{2 \phi(x)} \bar{g}$ of the ambient metric $\bar{g}$ on $\bar{\mc{M}}^n$, the volume form of the immersed $m$-dimensional submanifold $f(\mc{M})$ rescales by the formula $d\mu_{\hat{g}}=e ^{m \phi(x)} d\mu_{g}$, where of course $\hat{g}= e ^{2 \phi(x)} g$ is the induced conformal deformation on $f(\mc{M})$. In particular, for every constant $t\in \mathbb{R}_+$, we have $d\mu_{t^2 g}= t^m d \mu_{g}$. Thus, for any scalar geometric quantity $P(\bar{g},R, R^\perp, h)$ of weight $-m$, the integral $\int_{\mc{M}^m} P(\bar{g},R, R^\perp, h) \, d\mu_g$ is scale invariant for all compact orientable $m$-dimensional immersed submanifolds in any $n$-dimensional ambient Riemannian manifold. We are actually interested in those scalar geometric quantity $P(\bar{g},R, R^\perp, h)$ of weight $-m$ which give rise to integrals which are invariant not only under constant rescalings, but under general conformal rescalings. Let us give a precise definition.
\begin{definition}[Global conformal invariants of submanifolds]\label{def:GCI}
Let $P(\bar{g},R, R^\perp, h)$ be a geometric scalar quantity as in \eqref{eq:PgRh} and consider the conformal rescaling $\hat{\bar{g}}:= e ^{2 \phi(x)} \bar{g}$ of the ambient metric $\bar{g}$, for $\phi \in C^\infty(\bar{\mc{M}})$. By denoting $\hat{R}, \hat{R}^\perp, \hat{h}$ the tensors computed with respect to the conformal metric $\hat{\bar{g}}$, if
\begin{equation}\label{eq:ConInv}
\int_{\mc{M}^m} P(\hat{\bar{g}},\hat{R}, \hat{R}^\perp, \hat{h}) \, d \mu_{\hat{g}}= \int_{\mc{M}^m} P(\bar{g},R, R^\perp, h)\, d \mu_g
\end{equation}
for any ambient Riemannian manifold $\bar{\mc{M}}^n$, any compact orientable $m$-dimensional immersed submanifold $f(\mc{M}^m)\subset \bar{\mc{M}}^n$ and any $\phi \in C^\infty(\bar{\mc{M}})$, we then say that $\int_{\mc{M}^m} P(\bar{g},R, R^\perp, h)\, d \mu_g$ is a \emph{global conformal invariant for $m$-submanifolds in $n$-manifolds}.
\end{definition}
In this paper we address the question of classifying (at least in some cases) such global conformal invariants of submanifolds.
\subsection{The Operator $ I _{\bar{g},R, R^\perp, h} (\phi) $ and its Polarisations} \label{SS:Ig}
Inspired by the work of Alexakis \cite{AlexI,AlexII,AlexPf1, AlexPf2, AlexIV, AlexBook} on the classification of global conformal invariants of Riemannian manifolds, we introduce the useful operator $I_{\bar{g},R, R^\perp, h}(\phi)$, which ``measures how far the scalar geometric invariant $P(\bar{g},R, R^\perp, h)$ is to give rise to a global conformal invariant''.
\begin{definition}\label{def:Iphi}
Let $P(\bar{g},R, R^\perp, h)$ be a linear combination
$$P(\bar{g},R, R^\perp, h):= \sum_{ l \in L } a _ l C ^l (\bar{g},R, R^\perp, h) \quad,$$
where each $C ^l (\bar{g},R, R^\perp, h)$ is in the form \eqref{eq:ContrGen} and has weight $-m$, and assume that $P(\bar{g},R, R^\perp, h)$ gives rise to a global conformal invariant for $m$-submanifolds, that is it satisfies \eqref{eq:ConInv}. We define the differential operator $I _{\bar{g},R, R^\perp, h} (\phi)$, which depends both on the geometric tensors $\bar{g},R, R^\perp, h$ and on the auxiliary function $\phi\in C^{\infty}(\bar{\mc{M}}^n)$ as
\begin{equation}\label{eq.defI}
I _{\bar{g},R, R^\perp, h} (\phi)(x):= e^{m \phi(x)} P(\hat{\bar{g}},\hat{R}, \hat{R}^\perp, \hat{h})(x)-P(\bar{g},R, R^\perp, h)(x)\quad,
\end{equation}
where we use the notation of Definition \ref{def:GCI}.
\end{definition}
Notice that, thanks to \eqref{eq:ConInv}, it holds
\begin{equation}\label{eq:IntIInv}
\int_{\mc{M}^m} I _{\bar{g},R, R^\perp, h} (\phi) \, d \mu_g =0 \quad,
\end{equation}
for every Riemannian $n$-manifold $\bar{\mc{M}}^n$, every compact orientable $m$-submanifold $f(\mc{M}^m)\subset \bar{\mc{M}}^n$, and every function $\phi \in C^\infty(\bar{\mc{M}})$.
\\
By using the transformation laws for $R, R^\perp, h$ under conformal rescalings recalled in Section \ref{SSS:RSG}, it is clear that $I _{\bar{g},R, R^\perp, h} (\phi)$ is a differential operator acting on the function $\phi$. In particular we can polarize, that is we can pick any $A>0$ functions $\psi_1(\cdot), \ldots, \psi_{A}(\cdot)$, and choose
$$\phi(x):=\sum_{l=1}^A \psi_l(x)\quad.$$
Thus, we have a differential operator $I _{\bar{g},R, R^\perp, h} (\psi_1, \ldots, \psi_A)(\cdot)$ so that, by \eqref{eq:IntIInv}, it holds
$$\int_{\mc{M}^m} I _{\bar{g},R, R^\perp, h} (\psi_1, \ldots, \psi_A) \, d \mu_g =0 \quad, $$
for every Riemannian $n$-manifold $\bar{\mc{M}}^n$, every compact orientable $m$-submanifold $f(\mc{M}^m)\subset \bar{\mc{M}}^n$, and any functions $\psi_1, \ldots, \psi_A \in C^\infty(\bar{\mc{M}})$.
\\
Now, for any given functions $\psi_1(\cdot),\ldots, \psi_{A}(\cdot)$, we can consider the rescalings
$$\lambda_1 \psi_1(\cdot), \ldots, \lambda_{A} \psi_{A}(\cdot), $$
and, as above, we have the equation
\begin{equation}\label{eq:IntLambda}
\int_{\mc{M}^m} I _{\bar{g},R, R^\perp, h} (\lambda_1 \psi_1, \ldots, \lambda_A \psi_A) \, d \mu_g =0 \quad.
\end{equation}
The trick here is to see $\int_{\mc{M}^m} I _{\bar{g},R, R^\perp, h} (\lambda_1 \psi_1, \ldots, \lambda_A \psi_A) \, d \mu_g$ as a polynomial $\Pi(\lambda_1, \ldots, \lambda_A)$ in the independent variables $\lambda_1, \ldots, \lambda_A$. But then, equation \eqref{eq:IntLambda} implies that such polynomial $\Pi(\lambda_1, \ldots, \lambda_A)$ is identically zero. Hence, each coefficient of each monomial in the variables $\lambda_1, \ldots, \lambda_A$ must vanish.
We will see later in the proofs of the results how to exploit this crucial trick.
\section{Global Conformal Invariants of Surfaces}
It is well known that in the Euclidean space $ \mathbb{R} ^ n $, the Willmore energy $ \mathcal {W} ( f):=\int |H|^2 d \mu_g$ of a surface is invariant under conformal maps, that is under M\"obius transformations with inversion centred off the surface. It can be shown that this is a consequence of the fact that the conformal Willmore energy of a surface immersed in a general Riemannian manifold $ f :\mc{M}^2 \hookrightarrow \bar{\mc{M}} ^ n $
\begin{align*}
\mathcal W_{conf}(f) = \int_{ \mc{M}} | H| ^ 2 d \mu _g + \int _{ \mc{M}} \bar K d \mu_g
\end{align*}
where $ \overline K $ is the sectional curvature of the ambient space restricted to the surface, is invariant under conformal deformations of the ambient metric. Recall that by the Gauss equation one can write the intrinsic Gauss curvature $K$ of $(\mc{M}^2,g)$ as $K = ( |H| ^ 2 - \frac{1}{2} |h^\circ|^2 ) + \overline K $, where $ | h^\circ | ^ 2 = g^{ik}g^{jl} h^\circ_{ij} h^\circ_{kl}$ is the squared norm of the traceless second fundamental form, we can rewrite $\mathcal W_{conf}$ as
$$\mathcal W_{conf}(f)=\frac{1}{2} \int _{ \mc{M}} | h^\circ | ^ 2 d \mu _g + \int _{ \mc{M} } K d \mu _g \quad . $$
Notice that both $\int _{ \mc{M}} | h^\circ | ^ 2 d \mu _g$ and $\int _{ \mc{M} } K d \mu _g$ are natural conformal invariants: the first integral is conformally invariant as $ |h^\circ |^ 2 d \mu _g $ is a pointwise conformally invariant thanks to formula \eqref{eq:CChH}, and the second integral is conformally invariant by the Gauss-Bonnet theorem (more generally it is a \emph{topological} invariant). It trivially follows that any linear combination of the two integrands gives rise to a global conformal invariant. Our next result is that actually in codimension one there are no other global conformal invariants and in codimension two the situation is analogous once also the normal curvature is taken into account.
\subsection{Global Conformal Invariants of Codimension One Surfaces}
As announced before, as first result we show that, for codimension one surfaces, any global conformal invariant must be a linear combination of the squared norm of the traceless second fundamental form and the intrinsic Gauss curvature.
\begin{theorem}\label{thm:CD1surf}
Let $ P ( \bar{g} ,R, R^\perp, h )=\sum_{l\in L} a_l C^l(\bar{g},R,R^\perp,h) $ be a geometric scalar quantity for two-dimensional submanifolds of codimension one made by linear combinations of complete contractions in the general form
\begin{equation}\label{eq:ContrGen1}
C^l(\bar{g},R,R^\perp,h) = \contr(\bar{g}^{\alpha_1 \beta_1}\otimes \ldots \otimes R_{i_1 j_1 k_1 l_1} \otimes \ldots \otimes R^\perp_{i_r, j_r, \alpha_r, \beta_r} \otimes \ldots \otimes h^{\circ}_{i_s j_s} \otimes \ldots \otimes H g_{i_t j_t} ) \quad,
\end{equation}
and assume that $ \int _{\mc{M}}P ( \bar{g} ,R, R^\perp, h ) d \mu_{g}$ is a global conformal invariant, in the sense of Definition \ref{def:GCI}.
Then there exist $a,b \in \mathbb{R}$ such that $ P$ has the following decomposition
\begin{align*}
P ( \bar{g} ,R, R^\perp, h )= a K + b | h^\circ |^ 2 \quad.
\end{align*}
\end{theorem}
\begin{proof}
First of all notice that since by assumption here we are working in codimension one, the normal curvature $R^\perp$ vanishes identically so it can be suppressed without losing generality. Recall that by taking a constant rescaling function $\phi\equiv \log t$, for some $t>0$, the volume form rescales as $t^2$:
$d\mu_{t^2 g}= t^2 d\mu_{g}$. Therefore, if $P ( \bar{g} ,R ,h ) $ gives rise to a global conformal invariant it must necessarily be a linear combination of complete contractions $C^ l (\bar{g},R,h)$ each of weight $-2$. Observing that any contraction of $g^{-1}\otimes g^{-1} \otimes R$ is already of weight $-2$ and that any contraction of $g^{-1} \otimes h$ is of weight $-1$, the only possibility for $P ( \bar{g} ,R ,h )$ to be of weight $-2$ is that it decomposes as
\begin{eqnarray}
P(\bar{g},R,h)&=& \quad a \contr(g^{i_1j_1}\otimes g^{i_2 j_2} \otimes R_{i_3 j_3 i_4 j_4}) \nonumber \\
&&+ b \contr(g^{i_1j_1}\otimes g^{i_2 j_2}\otimes H g_{i_{3} j_{3}}\otimes \ldots \otimes H g_{i_{r+2} j_{r+2}}\otimes h^\circ_{i_{r+3} j_{r+3}}\otimes \ldots \otimes h^\circ_{i_{r+4} j_{r+4}} ), \label{eq:Psplits}
\end{eqnarray}
where in the above formula we intend that $r=0,1,2$ is the number of $H$ factors, and $2-r$ is the number of $h^\circ$ factors.
Clearly, since by assumption $\mc{M}$ is a 2-d manifold, the term $\contr(g^{i_1j_1}\otimes g^{i_2 j_2} R_{i_3 j_3 i_4 j_4})$ is a (possibly null) multiple of the Gauss curvature. To get the thesis it is therefore sufficient to prove that $r=0$, that is the second summand in \eqref{eq:Psplits} is completely expressed in terms of the traceless second fundamental form. Indeed any complete contraction of $g^{-1}\otimes g^{-1} \otimes h^\circ$ is a (possibly null) multiple of $|h^\circ|^2$.
To this aim observe that, since by the Gauss Bonnet Theorem $\int_{\mc{M}} K d \mu_g$ is a global conformal invariant, called
\begin{eqnarray}
P_1(g, h)&:=& P(\bar{g},R,h)- a K \nonumber \\
&=& b \contr(g^{i_1j_1}\otimes g^{i_2 j_2}\otimes H g_{i_{3} j_{3}}\otimes \ldots \otimes H g_{i_{r+2} j_{r+2}}\otimes h^\circ_{i_{r+3} j_{r+3}}\otimes \ldots \otimes h^\circ_{i_{r+4} j_{r+4}} ), \label{eq:defP1gh}
\end{eqnarray}
we have that $\int_{\mc{M}} P_1(g,h) d \mu_g$ is a global conformal invariant for compact surfaces immersed in 3-manifolds, as difference of such objects.
\\Consider then an arbitrary compact surface $f(\mc{M}^2)$ immersed into an arbitrary Riemannian $3$-manifold $(\bar{\mc{M}}^3, \bar{g}^3)$, and an arbitrary conformal rescaling $\hat{\bar{g}}^3=e^{2 \phi(x)} \bar{g}^3$ of the ambient metric by a smooth function $\phi\in C^\infty(\bar{\mc{M}})$. All the hatted geometric tensors $\hat{g}, \hat{h}, \hat{h}^\circ, \hat H$ denote the corresponding tensors computed with respect to the rescaled metric $\hat{\bar{g}}$. In Section \ref{SS:Ig} we defined the operator
$$I^1_{(g,h)}(\phi):= e^{2 \phi} P_1(\hat{g}, \hat{h})- P_1(g,h)\quad ,$$
and we observed that the conformal invariance of the integrated quantities implies (see \eqref{eq:IntIInv})
\begin{equation}\label{eq:IntI=00}
\int_{\mc{M}} I^1_{(g,h)}(\phi) \, d \mu_{g}=0 \quad.
\end{equation}
From the formulas \eqref{eq:CChH} of the change of $h^\circ$ and $H$ under a conformal deformation of metric, it is clear that $I^1_{(g,h)}(\phi)$ does not depend directly on $\phi$ but just on $\pi_N(\bar{\nabla} \phi)= (\partial_N \phi) N $, the normal derivative of $\phi$. More precisely $I^1_{(g,h)}(\phi)$ is a polynomial in $\partial_N \phi$ exactly of the same degree $0\leq r\leq 2$ as $P_1(g, h^\circ, H)$ seen as a polynomial in $H$.
By considering $t\phi$ for $t \in \mathbb{R}$, we therefore get that
$$I^1_{g, h}(t \phi)= \sum_{k=1}^r a_k C^k(g, h^\circ, H) \; t^k \left(\partial_N \phi \right)^k \quad.$$
Recalling now \eqref{eq:IntI=00}, we obtain that $ \int_{\mc{M}} I^1_{(g,h)}(t\phi) \, d \mu_{g} $ vanishes identically as a polynomial in $t$, so
\begin{equation}\label{eq:IntIk0}
0=\frac{d^k}{dt^k}\big|_{{t=0}} \int_{\mc{M}} I^1_{(g,h)}(t\phi) \, d \mu_{g} = k! \int_{\mc{M}} a_k C^k(g, h^\circ, H) \left( \partial_N \phi \right)^k \, d \mu_{g},\quad \forall k=0,\ldots,r.
\end{equation}
Pick an arbitrary point $x \in \mc{M}$; by choosing local coordinates in $\bar{\mc{M}}^3$ adapted to $f(\mc{M})$ at $f(x)$, it is easy to see that for any given function $\psi \in C^{\infty}_c(\mc{M})$ supported in such coordinate neighborhood of $x$, there exists $\phi \in C^\infty(\bar{\mc{M}})$ such that
$$\psi = \frac{\partial \phi}{\partial x^3} \circ f= \partial_N \phi \circ f \quad. $$
By plugging such arbitrary $C^\infty_c(\mc{M})$ function $\psi$ in place of $\partial_N \phi$ in \eqref{eq:IntIk0}, we obtain that not only the integrals but the integrands themselves must vanish, that is $a_k C^k(g, h^\circ, H) \equiv 0$ on $\mc{M}$. It follows that $I^1_{g,h}(\phi)\equiv 0$ or, in other words, the degree of $I^1_{g, h}(t \phi)$ as a polynomial in $t$ is $0$. By the above discussion we have then that $r=0$, which was our thesis.
\end{proof}
\subsection{Global Conformal Invariants of Codimension Two Surfaces}
Let $ (\bar{\mc{M}} ^ 4, \bar{g} _{ \alpha \beta } ) $ be a four dimensional Riemannian manifold and $ f : \mc{M} \hookrightarrow \bar{\mc{M}}$ an immersion of an oriented compact surface $\mc{M}^2$.
In local coordinates, if $ \{ e _1, e _ 2, e _ 3 , e _ 4 \} $ is an adapted orthonormal frame, so that $ \{ e _1, e _2 \} $ is a frame on $ T f(\mc{M})$ and $ \{ e _3, e _ 4\} $ is a frame on $ [T f (\mc{M})]^\perp$ we define
\begin{align*}
K ^ \perp = R ^ \perp( e _1, e _2, e _3, e _ 4 )\quad,
\end{align*}
where $ R ^ \perp$ is the curvature tensor of the normal connection defined in \eqref{eq:defRn}. Note that $K^\perp$ is well defined up to a sign depending on orientation, indeed by the symmetries of the normal curvature tensor, for a codimension two surface $R^\perp$ has only one non-zero component, namely $\pm K ^ \perp$.
We also denote by
$$ \bar K^ \perp = \bar R ( e _1, e _2 , e _ 3 ,e _4 )$$
the ambient curvature evaluated on the normal bundle.
For a codimension two surface, we note that we have two topological invariants: the integral of the Gauss curvature $K$ giving the Euler Characteristic of $\mc{M}$ via the Gauss-Bonnet Theorem, and the integral of the normal curvature $K^\perp$ which gives the Euler characteristic of the normal bundle,
\begin{equation}\label{eq:TopInv}
\int_{\mc{M}} K d \mu _g = 2 \pi \chi(\mc{M}) , \quad \int _{\mc{M}} K ^ \perp d \mu _g = 2 \pi \chi^ \perp (f(\mc{M})).
\end{equation}
As already observed, $|h^ \circ|^2 d \mu_g$ is a \emph{pointwise} conformal invariant. Hence any linear combination of these three integrands is a global conformal invariant and in the next theorem we show that there are no others.
\begin{theorem}\label{thm:CD2surf}
Let $ P ( \bar{g} ,R, R^\perp, h )=\sum_{l\in L} a_l C^l(\bar{g},R,R^\perp,h) $ be a geometric scalar quantity for two-dimensional submanifolds of codimension two made by linear combinations of complete contractions in the general form
\begin{equation}\label{eq:ContrGen1}
C^l(\bar{g},R,R^\perp,h) = \contr(\bar{g}^{\alpha_1 \beta_1}\otimes \ldots \otimes R_{i_1 j_1 k_1 l_1} \otimes \ldots \otimes R^\perp_{i_r, j_r, \alpha_r, \beta_r} \otimes \ldots \otimes h^{\circ}_{i_s j_s} \otimes \ldots \otimes H g_{i_t j_t} ) \quad,
\end{equation}
and assume that $ \int _{\mc{M}}P ( \bar{g} ,R, R^\perp, h ) d \mu_{g}$ is a global conformal invariant, in the sense of Definition \ref{def:GCI}.
Then there exist $a,b,c \in \mathbb{R}$ such that $ P$ has the following decomposition
\begin{align*}
P ( \bar{g} ,R, R^\perp, h )= a K +b K^\perp + c | h^\circ |^ 2 \quad.
\end{align*}
\end{theorem}
\begin{proof}
Exactly as in the proof of Theorem \ref{thm:CD1surf}, if $P ( \bar{g} ,R, R^\perp ,h ) $ gives rise to a global conformal invariant it must necessarily be a linear combination of complete contractions $C^ l (\bar{g},R, R^\perp, h)$ each of weight $-2$. Observing that any contraction of $g^{-1}\otimes g^{-1} \otimes R$ and of $\bar{g}^{-1}\otimes \bar{g}^{-1} \otimes R^\perp$ is already of weight $-2$, and that any contraction of $g^{-1} \otimes h$ is of weight $-1$, the only possibility for $P ( \bar{g} ,R, R^\perp, h )$ to be of weight $-2$ is that it decomposes as
\begin{eqnarray}
P(\bar{g},R, R^\perp, h)&=& \quad a \contr(g^{i_1j_1}\otimes g^{i_2 j_2} \otimes R_{i_3 j_3 i_4 j_4}) + b \contr(\bar{g}^{\alpha_1 \beta_1}\otimes g^{\alpha_2 \beta_2} \otimes R^\perp_{\alpha_3 \beta_3 \alpha_4 \beta_4}) \nonumber \\
&&+ c \contr(g^{i_1j_1}\otimes g^{i_2 j_2}\otimes H g_{i_{3} j_{3}}\otimes \ldots \otimes H g_{i_{r+2} j_{r+2}} \nonumber \\
&& \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \; \otimes h^\circ_{i_{r+3} j_{r+3}}\otimes \ldots \otimes h^\circ_{i_{r+4} j_{r+4}} ), \nonumber
\end{eqnarray}
where in the above formula we intend that $r=0,1,2$ is the number of $H$ factors, and $2-r$ is the number of $h^\circ$ factors.
Clearly, since by assumption $\mc{M}$ is a 2-d manifold, the term $\contr(g^{i_1j_1}\otimes g^{i_2 j_2} R_{i_3 j_3 i_4 j_4})$ is a (possibly null) multiple of the Gauss curvature. Analogously, since $f(\mc{M})\subset \bar{\mc{M}}$ is a codimension two submanifold, the term $\contr(\bar{g}^{\alpha_1 \beta_1}\otimes g^{\alpha_2 \beta_2} \otimes R^\perp_{\alpha_3 \beta_3 \alpha_4 \beta_4}) $
is a (possibly null) multiple of the normal curvature $K^\perp$. But, by \eqref{eq:TopInv}, we already know that $\int_{\mc{M}} K d \mu_g$ and $\int_{\mc{M}} K^\perp d \mu_g$ are global conformal invariants so, called
\begin{eqnarray}
P_1(g, h)&:=& P(\bar{g},R,R^\perp, h)- a K- bK^\perp \nonumber \\
&=& b \contr(g^{i_1j_1}\otimes g^{i_2 j_2}\otimes H g_{i_{3} j_{3}}\otimes \ldots \otimes H g_{i_{r+2} j_{r+2}}\otimes h^\circ_{i_{r+3} j_{r+3}}\otimes \ldots \otimes h^\circ_{i_{r+4} j_{r+4}} ), \nonumber
\end{eqnarray}
also $\int_{\mc{M}} P_1(g,h) d \mu_g$ is a global conformal invariant for compact surfaces immersed in 3-manifolds, as difference of such objects. The thesis can be now achieved by repeating verbatim the second part of the proof of Theorem \ref{thm:CD1surf}.
\end{proof}
\subsubsection{Two examples: the complex projective plane and the complex hyperbolic plane}
Two particular cases of interest (apart from the spaces forms) are $ \mathbb{CP } ^ 2 $ and $ \mathbb{CH}^ 2$ the complex projective plane and the complex hyperbolic plane respectively. These are K\"ahler manifolds with their standard K\"ahler form $ \Omega$ of constant holomorphic sectional curvature and unlike their real counterparts, $\mathbb{S}^4$ and $ \mathbb H ^ 4$, they are not locally conformal to $ \mathbb {C} ^ 2 $.
Let us consider an immersion $ \phi : \mc{M} \hookrightarrow \bar{\mc{M}}$ of an oriented surface, where $ \bar{\mc{M}} = \mathbb {CP}^ 2, \mathbb{CH}^2$. The K\"ahler function $C$ on $\mc{M} $ is defined by $ \phi ^ * \Omega = C d \mu_g$. Only the sign of $ C$ depends on the orientation, hence $ C ^2 $ and $ |C|$ are well defined for non-orientable surfaces.
The K\"ahler function satisfies
\begin{align*}
-1 \leq C \leq 1 \quad.
\end{align*}
By direct computation, we find that the Willmore functional is equal to
\begin{align*}
\mathcal {W}_{\mathbb{CP}^ 2 }( \phi) = \int _{\mc{M}} \left( | H | ^ 2 + \overline K \right) \, d \mu_g = \int _{ \mc{M}} \left( | H| ^ 2 + 1 + 3 C ^ 2 \right) d \mu_g \quad ,
\end{align*}
and
\begin{align*}
\mathcal{W}_{\mathbb{CH} ^ 2 } ( \phi )= \int _{\mc{M}} \left( | H | ^ 2 + \overline K \right) d \mu_g = \int _{ \mc{M}} \left( | H| ^ 2 - 1 -3 C ^ 2 \right) d \mu_g \quad.
\end{align*}
By the Ricci equation \eqref{eq:Ricci}, we also have
\begin{align*}
\overline K ^ \perp = \overline R_{1234} = K^\perp - \left(( \accentset{\circ} h_{1p})_{3} (\accentset{\circ} h _{ 2p})_{4} - ( \accentset{\circ} h _{ 2 p})_{ 3 }( \accentset{\circ} h _{ 1 p})_{4 } \right)
\end{align*}
which, applying the symmetries of the curvature tensor, can be written as a complete contraction. Thanks to the last formula, it is clear that $\int \overline K ^ \perp d \mu_g$ is a global conformal invariant for codimension two surfaces (since linear combination of such objects); therefore, the following energies are all global conformal invariant:
\begin{align}
\mathcal W ^ +_{\mathbb {CP}^2} ( \phi) &= \int _{ \mc{M}} | H| ^ 2 + \overline K - \overline K ^ \perp = \int_{\mc{M}} (|H|^ 2 + 6 C ^ 2 ) d \mu_g \label{eq:WCP21} \\
\mathcal W ^ -_{\mathbb {CP}^2} ( \phi) &= \int _{ \mc{M}} | H| ^ 2 + \overline K + \overline K ^ \perp = \int_{\mc{M}} (|H|^ 2 + 2) d \mu_g \quad, \label{eq:WCP22}
\end{align}
and
\begin{align}
\mathcal W ^ +_{\mathbb {CH}^2} ( \phi) &= \int _{ \mc{M}} | H| ^ 2 + \overline K - \overline K ^ \perp = \int_{\mc{M}} (|H|^ 2 - 6 C ^ 2 ) d \mu_g \label{eq:WCH21} \\
\mathcal W ^ -_{\mathbb {CH}^2} ( \phi) &= \int _{ \mc{M}} | H| ^ 2 + \overline K + \overline K ^ \perp = \int_{\mc{M}} (|H|^ 2 - 2) d \mu_g \quad. \label{eq:WCH22}
\end{align}
Let us remark that the energies \eqref{eq:WCP21},\eqref{eq:WCP22} have already been object of investigation in \cite{MontielUrbano}, where it was shown that $ \mathcal {W}^-_{\mathbb {CP} ^ 2} ( \phi) \geq 4 \pi \mu - 2 \int_{\Sigma}|C| d \mu_g $ where $\mu$ is the maximum multiplicity. Moreover in the same paper it was shown that equality holds if and only if $ \mu =1 $ and $\mc{M}$ is either a complex projective line or a totally geodesic real projective plane, or $ \mu = 2 $ and $ \mc{M}$ is a Lagrangian Whitney sphere.
\begin{remark}
Since the goal of the present paper is to investigate the structure of global conformal invariants, we recalled the definition of $\mathcal W ^{\pm}_{\mathbb {CP}^2}$ given in \cite{MontielUrbano} and we defined the new functionals $\mathcal W ^{\pm}_{\mathbb {CH}^2}$ in order to give interesting examples of Willmore-type energies in codimension 2. In a forthcoming work we will address the question wether $\mathcal W ^{\pm}_{\mathbb {CH}^2}$ satisfy analogous properties as $\mathcal W ^{\pm}_{\mathbb {CP}^2}$.
\end{remark}
\section{Global Conformal Invariants of Submanifolds}
Let us consider a geometric scalar quantity $ P (g^m, h^m)$ of the form
\begin{align} \label{eqn_star}
\sum_{ l \in L } a _l C^ l ( g^m , h^m) \quad,
\end{align}
where each $C ^ l$ is a complete contraction
\begin{align} \label{eqn_contract}
\contr(g^{i_1 j_1}\otimes\ldots \otimes g^{i_s j_s} \otimes h_{ i _{s+1} j_{s+1}}\otimes \dots \otimes h _{ i _{2s} j_{2s}}) \quad ;
\end{align}
that is we consider complete contractions as defined in \eqref{eq:ContrGen} but depending just on the second fundamental form $h^m$ and the induced metric $g^m$ for immersed $m$-submanifolds $f(\mc{M}^m)$ in Riemannian $n$-manifolds $(\bar{\mc{M}}^n,\bar{g}^n)$.
Our first goal in this section is to understand the structure of geometric scalar quantities \eqref{eqn_star} giving rise to global conformal invariants for submanifolds, in the sense of Definition \ref{def:GCI}.
This is exactly the content of the next result.
\begin{theorem}\label{thm:P(g,h)}
Let $ P ( g^m ,h^m ) $ be as in \eqref{eqn_star} with each $ C^l$ of the form \eqref{eqn_contract}, and assume that $ \int _{\mc{M}}P( g ^ m ,h ^ m) d \mu_{g^m}$ is a global conformal invariant, in the sense of Definition \ref{def:GCI}. Then there exists a \emph{pointwise} conformal invariant $W ( g ^ m, \accentset{\circ} h^m )$ of weight $ -m$ depending only on the traceless second fundamental form $\accentset{\circ} h$ contracted with the induced metric $g^m$ so that
\begin{align*}
P( g ^ m , h ^ m ) = W ( g ^ m, \accentset{\circ} h ^m )\quad .
\end{align*}
In other words, for every $l \in L$ one has that $ C ^l$ in \eqref{eqn_star} is a complete contraction of weight $-m$ of the form
\begin{align} \label{eqn_contract}
\contr(g^{i_1 j_1}\otimes\ldots \otimes g^{i_m j_m} \otimes \accentset{\circ} h_{ i _{m+1} j_{m+1}}\otimes \dots \otimes \accentset{\circ} h _{ i _{2m} j_{2m}}) \quad .
\end{align}
\end{theorem}
\begin{proof}
First of all notice that by taking a constant rescaling function $\phi\equiv \log t$, for some $t>0$, the volume form rescales with a power $m$, that is
$d\mu_{{t^2 g}^m}= t^m d\mu_{{g}^m}$. Therefore, if $P ( g^m ,h^m ) $ gives rise to a global conformal invariant it must necessarily be a linear combination of complete contractions $C^ l ( g^m , h^m)$ each of weight $-m$. But observing that any contraction of $g^{-1} \otimes h$ is of weight $-1$, this implies that in \eqref{eqn_contract} we have $s=m$, that is $C ^ l(g^m, h^m)$ is a complete contraction of the form
\begin{align} \label{eqn_contract1}
\contr(g^{i_1 j_1}\otimes\ldots \otimes g^{i_m j_m} \otimes h_{ i _{m+1} j_{m+1}}\otimes \dots \otimes h _{ i _{2m} j_{2m}}) \quad .
\end{align}
Recalling \eqref{eq:Splith}, that is the orthogonal splitting $h_{ij}=H g_{ij}+h _{ij}^\circ$ , we can rewrite such a complete contraction as
\begin{align} \label{eqn_contract1}
\contr(g^{i_1 j_1}\otimes\ldots \otimes g^{i_m j_m} \otimes H g_{ i _{m+1} j_{m+1}}\otimes \dots \otimes H g_{ i_{m+r} j_{m+r}} \otimes h_{ i _{m+r+1} j_{m+r+1}}^\circ \otimes \dots \otimes h _{ i _{2m} j_{2m}}^\circ ) \quad .
\end{align}
Our goal is to prove that $r=0$, that is there are no $H$ factors, so $P(g^m,h^m)=P(g^m,\accentset{\circ} h^m)$ is expressed purely as complete contractions of traceless fundamental forms, which are \emph{pointwise} conformal invariants once multiplied by $d\mu_g$ thanks to \eqref{eq:CChH}.\\
To that aim consider an arbitrary compact $m$-dimensional immersed submanifold $f(\mc{M}^m)$ of an arbitrary Riemannian $n$-manifold $(\bar{\mc{M}}^n, \bar{g}^n)$, and an arbitrary conformal rescaling $\hat{\bar{g}}^n=e^{2 \phi(x)} \bar{g}^n$ of the ambient metric by a smooth function $\phi\in C^\infty(\bar{\mc{M}})$. All the hatted geometric tensors $\hat{g}, \hat{h}, \hat{h}^\circ, \hat H$ denote the corresponding tensors computed with respect to the deformed metric $\hat{\bar{g}}$. In Section \ref{SS:Ig} we defined the operator
$$I_{(g,h)}(\phi):= e^{m \phi} P(\hat{g}, \hat{h})- P(g,h)\quad ,$$
and we observed that the conformal invariance of the integrated quantities implies (see \eqref{eq:IntIInv})
\begin{equation}\label{eq:IntI=0}
\int_{\mc{M}} I_{(g,h)}(\phi) \, d \mu_{g}=0 \quad.
\end{equation}
From the formulas \eqref{eq:CChH} of the change of $h^\circ$ and $H$ under a conformal deformation of metric, it is clear that $I_{(g,h)}(\phi)$ does not depend directly on $\phi$ but just on $\pi_N(\bar{\nabla} \phi)$, the projection of $\bar{\nabla} \phi$ onto the normal space of $f(\mc{M})$. More precisely $I_{(g,h)}(\phi)$ is polynomial in the components of $\pi_N(\bar{\nabla} \phi)$ exactly of the same degree $0\leq r\leq m$ as $P(g, h^\circ, H)$ seen as a polynomial in $H$.
By considering $t\phi$ for $t \in \mathbb{R}$, we get that
$$I_{g, h}(t \phi)= \sum_{k=1}^r a_k C^k(g, h^\circ, H, \pi_N(\bar{\nabla} \phi)) \; t^k\quad ,$$
where $C^k(g, h^\circ, H, \pi_N(\bar{\nabla} \phi))$ is an homogeneous polynomial of degree $k$ in the components of $\pi_N(\bar{\nabla} \phi)$. Recalling now \eqref{eq:IntI=0}, we obtain that $ \int_{\mc{M}} I_{(g,h)}(t\phi) \, d \mu_{g} $ vanishes identically as a polynomial in $t$, so
\begin{equation}\label{eq:IntIk}
0=\frac{d^k}{dt^k}\mid_{{t=0}} \int_{\mc{M}} I_{(g,h)}(t\phi) \, d \mu_{g} = k! \int_{\mc{M}} a_k C^k(g, h^\circ, H, \pi_N(\bar{\nabla} \phi)) \, d \mu_{g},\quad \forall k=0,\ldots,r.
\end{equation}
Pick an arbitrary point $x \in \mc{M}$; by choosing local coordinates in $\bar{\mc{M}}^n$ adapted to $f(\mc{M})$ at $f(x)$, it is easy to see that for any given functions $\psi^i,\ldots, \psi^{n-m} \in C^{\infty}_c(\mc{M})$ supported in such a coordinate neighborhood of $x$, there exists $\phi \in C^\infty(\bar{\mc{M}})$ such that $$\psi^i= (\bar{\nabla} \phi)^{m+i} \circ f \quad, $$
where of course thanks to this choice of coordinates we have
$$\pi_N(\bar{\nabla} \phi)=\Big((\bar{\nabla} \phi)^{m+1}, \ldots, (\bar{\nabla} \phi)^{n}\Big)\quad. $$
By plugging such arbitrary $C^\infty_c(\mc{M})$ functions $\psi^i,\ldots, \psi^{n-m}$ in place of $\pi_N(\bar{\nabla} \phi)$ in \eqref{eq:IntIk}, we obtain that not only the integrals but the integrands themselves must vanish, that is $a_k C^k(g, h^\circ, H, \pi_N(\bar{\nabla} \phi))\equiv 0$ on $\mc{M}$. It follows that $I_{g,h}(\phi)\equiv 0$ or, in other words, the degree of $I_{g, h}(t \phi)$ as a polynomial in $t$ is $0$. By the above discussion we have that $r=0$, which was our thesis. \end{proof}
We pass now to consider the more general geometric scalar quantity $ P (g^m, R^m, h^m)$, for $m\in \mathbb N$ \emph{even}, of the form
\begin{equation}\label{eq:PP1P2}
P (g^m, R^m, h^m)=P_1(g^m, h^m)+ P_2(g^m,R^m)\quad,
\end{equation}
where
\begin{align} \label{eq:P1P2}
P_1(g^m,h^m)=\sum_{ l \in L } a _l C^ l ( g^m , h^m) \quad \text{and} \quad P_2(g^m,R^m)=\sum_{ s \in S } b _s C^ s ( g^m , R^m) \quad,
\end{align}
where each $C ^ l( g^m , h^m)$ is a complete contraction
\begin{align} \label{eq:Cl}
\contr(g^{i_1 j_1}\otimes\ldots \otimes g^{i_s j_s} \otimes h_{ i _{s+1} j_{s+1}}\otimes \dots \otimes h _{ i _{2s} j_{2s}}) \quad
\end{align}
and each $C^ s ( g^m , R^m)$ is a complete contraction
\begin{align} \label{eq:Cs}
\contr(g^{i_1 j_1}\otimes\ldots \otimes g^{i_{2r} j_{2r}} \otimes R_{ i _{2r+1} j_{2r+1} k_{2r+1} l_{2r+1} }\otimes \dots \otimes R_{ i _{3r} j_{3r} k_{3r} l_{3r} }) \quad.
\end{align}
In other words we consider complete contractions as defined in \eqref{eq:ContrGen} which split in two parts: one depending just on the second fundamental form $h^m$ and the other one just on the intrinsic curvature $R^m$,
for immersed $m$-submanifolds $(f(\mc{M}^m), g^m)$ in Riemannian $n$-manifolds $(\bar{\mc{M}}^n,\bar{g}^n)$.
As usual, the goal is to understand the structure of geometric scalar quantities \eqref{eq:PP1P2} giving rise to global conformal invariants for submanifolds, in the sense of Definition \ref{def:GCI}.
This is exactly the content of the next result.
\begin{theorem}\label{thm:P=P1+P2}
Let $m \in \mathbb N$ be \emph{even} and let $P( g ^ m , R^m, h ^m )=P_1(g^m, h^m)+ P_2(g^m,R^m)$ be a geometric scalar quantity as above. Assume that $ \int _{\mc{M}}P( g ^ m, R^m, h ^ m) d \mu_{g^m}$ is a global conformal invariant, in the sense of Definition \ref{def:GCI}.
Then both $ \int _{\mc{M}}P_1( g ^ m, h ^ m) d \mu_{g^m}$ and $ \int _{\mc{M}}P_2( g ^ m, R ^ m) d \mu_{g^m}$ are global conformal invariants. It follows that
i) There exists a \emph{pointwise} conformal invariant $W_1 ( g ^ m, \accentset{\circ} h^m )$ of weight $ -m$ depending only on the traceless second fundamental form $\accentset{\circ} h$ contracted with the induced metric $g^m$ so that
\begin{align*}
P_1( g ^ m , h ^ m ) = W_1 ( g ^ m, \accentset{\circ} h ^m )\quad ;
\end{align*}
or, in other words, for every $l \in L$ one has that $ C ^l$ in \eqref{eq:Cl} is a complete contraction of weight $-m$ of the form
\begin{align} \label{eqn_contract2}
\contr(g^{i_1 j_1}\otimes\ldots \otimes g^{i_m j_m} \otimes \accentset{\circ} h_{ i _{m+1} j_{m+1}}\otimes \dots \otimes \accentset{\circ} h _{ i _{2m} j_{2m}}) \quad .
\end{align}
ii) Called $ \Pfaff(R^m)$ the Pfaffian of the intrinsic Riemann tensor $R^m$ and $W^m$ the Weyl tensor of $g^m$, $P_2(g^m, R^m)$ is of the form
\begin{align*}
P_2(g^m, R^m)= \tilde{P}_2(g^m, W^m)+ c \Pfaff(R^m) \;, \quad \text{for some } c\in \mathbb{R} \quad,
\end{align*}
where $\tilde{P}_2(g^m, W^m)$ is a \emph{pointwise} conformal invariant of weight $-m$ expressed as a linear combination of complete contractions of the form
$$\contr(g^{i_1 j_1}\otimes\ldots \otimes g^{i_{m} j_{m}} \otimes W_{ i _{m+1} j_{m+1} k_{m+1} l_{m+1} }\otimes \dots \otimes W_{ i _{\frac{3m}{2}} j_{\frac{3m}{2}} k_{\frac{3m}{2}} l_{\frac{3m}{2}} }) \quad. $$
\end{theorem}
\begin{remark}
It is well known that
\begin{itemize}
\item The Weyl tensor $W_{ijkl}(g^m)$ is a pointwise scalar conformal invariant of weight $2$, that is it satisfies $W_{ijkl}(e^{2\phi(x)} g^m(x))= e^{2 \phi(x) } W(g^m)(x)$ for every $\phi\in C^\infty(\mc{M})$ and every $x\in \mc{M}^m$. It follows that any complete contraction of the tensor $g^{-1}\otimes g^{-1} \otimes W$ is a pointwise scalar conformal invariant of weight $-2$.
\item The Pfaffian $\Pfaff(R_{ijkl})$ of the intrinsic curvature $R_{ijkl}=R^m$ integrated over the manifold gives rise to a \emph{topological} invariant:
$$\int_{\mc{M}^m} \Pfaff(R_{ijkl}) \, d\mu_g= \frac{2^m \pi^{m/2} (\frac{m}{2}-1)!}{2(m-1)!} \chi(\mc{M}^m)\quad, $$
where $\chi(\mc{M}^m)$ is the Euler Characteristic of $\mc{M}^m$.
\end{itemize}
Therefore, recalling the conformal invariance of the traceless second fundamental form \eqref{eq:CChH}, any linear combination of complete contractions
$$P(g^m,h^m,R^m)=W_1(g^m, \accentset{\circ} h^m)+ \tilde{P}_2(g^m, W^m)+ c \Pfaff(R^m)$$
as in the thesis of Theorem \ref{thm:P=P1+P2} gives rise to a ``trivial'' global conformal invariant. Thereom \ref{thm:P=P1+P2} states that, under the assumption that $P$ splits into the sum of an intrinsic part and of an extrinsic part depending just on the second fundamental form, this is actually the only possibility.
\end{remark}
\begin{proof}
Since by assumption $P( g ^ m , R^m, h ^m )=P_1(g^m, h^m)+ P_2(g^m,R^m)$ gives rise to a global conformal invariant, if we show that $\int_{\mc{M}} P_1(g^m, h^m) \, d \mu_g$ is a global conformal invariant, the same will be true for $\int_{\mc{M}} P_2(g^m, R^m) \, d \mu_g$. In order to prove that, consider an arbitrary compact $m$-dimensional immersed submanifold $f(\mc{M}^m)$ of an arbitrary Riemannian $n$-manifold $(\bar{\mc{M}}^n, \bar{g}^n)$, and an arbitrary conformal rescaling $\hat{\bar{g}}^n=e^{2 \phi(x)} \bar{g}^n$ of the ambient metric by a smooth function $\phi\in C^\infty(\bar{\mc{M}})$. All the hatted geometric tensors $\hat{g}, \hat{h}, \hat{h}^\circ, \hat H, \hat{R}$ denote the corresponding tensors computed with respect to the rescaled metric $\hat{\bar{g}}$. Analogously to Section \ref{SS:Ig}, define the operators
\begin{eqnarray}
I^1_{g,h}(\phi)&:=& e^{m \phi} P_1(\hat{g}, \hat{h})- P_1(g,h)\quad , \label{eq:I1} \\
I^2_{g,R}(\phi)&:=& e^{m \phi} P_2(\hat{g}, \hat{R})- P_2(g,R)\quad , \label{eq:I2} \\
I_{g,R,h}(\phi)&:=& I^1_{g,h}(\phi)+I^2_{g,R}(\phi) \quad . \label{eq:I=I1+I2}
\end{eqnarray}
As already observed in the proof of Theorem \ref{thm:P(g,h)}, $I^1_{g,h}(\phi)$ does not depend directly on $\phi$ but only on $\pi_N(\bar{\nabla} \phi)$, the projection of $\bar{\nabla} \phi$ onto the normal space of $f(\mc{M})$. More precisely $I^1_{g,h}(\phi)$ is polynomial in the components of $\pi_N(\bar{\nabla} \phi)$ exactly of the same degree $0\leq r\leq m$ as $P_1(g, h^\circ, H)$ seen as a polynomial in $H$.
\\On the other hand, recalling the formulas \eqref{eq:Rhatg} of the change of the intrinsic curvature $R_{ijkl}$ of $(\mc{M}^m, g^m)$ under conformal deformation of the metric, it is clear that $I^2_{g,R}$ does not depend directly on $\phi$ but only on $\nabla_\mc{M} \phi$, the projection of $\bar{\nabla} \phi$ onto the tangent space of $f(\mc{M})$, and on $\nabla^2_\mc{M} \phi$, the covariant Hessian of $\phi|_{\mc{M}}$. More precisely $I^2_{g,R}(\phi)$ is polynomial in the components of $\nabla_\mc{M} \phi, \nabla^2 _\mc{M} \phi$.
By considering $t\phi$ for $t \in \mathbb{R}$, we get that
$$I^1_{g, h}(t \phi)= \sum_{k=1}^r a_k C^k(g, h^\circ, H, \pi_N(\bar{\nabla} \phi)) \; t^k\quad ,\quad I^2_{g,R}(t \phi)= \sum_{l=1}^s b_l C^l(g, R, \nabla_\mc{M} \phi, \nabla^2 _{\mc{M}} \phi) \; t^k\quad $$
where $C^k(g, h^\circ, H, \pi_N(\bar{\nabla} \phi))$ (respectively $C^l(g, R, \nabla_\mc{M} \phi, \nabla^2 _{\mc{M}} \phi)$) is an homogeneous polynomial of degree $k$ in the components of $\pi_N(\bar{\nabla} \phi)$ (respectively of $\nabla_\mc{M} \phi, \nabla^2 _\mc{M} \phi$). Recalling now \eqref{eq:IntI=0}, we obtain that $ \int_{\mc{M}} I_{(g,h)}(t\phi) \, d \mu_{g} $ vanishes identically as a polynomial in $t$, so
\begin{eqnarray}
0&=&\frac{d^k}{dt^k}\bigg|_{{t=0}} \int_{\mc{M}} I_{(g,h)}(t\phi) \, d \mu_{g} \nonumber \\
&=& k! \int_{\mc{M}} a_k C^k(g, h^\circ, H, \pi_N(\bar{\nabla} \phi)) + b_k C^k(g, R, \nabla_\mc{M} \phi, \nabla^2 _{\mc{M}} \phi) \, d \mu_{g},\quad \forall k\in \mathbb N. \label{eq:akbk}
\end{eqnarray}
Pick an arbitrary point $x \in \mc{M}$; by choosing local coordinates in $\bar{\mc{M}}^n$ adapted to $f(\mc{M})$ at $f(x)$, it is easy to see that for any given functions $\psi^i,\ldots, \psi^{n-m} \in C^{\infty}_c(\mc{M})$ supported in such coordinate neighborhood of $x$, there exists $\phi \in C^\infty(\bar{\mc{M}})$ such that
$$\psi^i= (\bar{\nabla} \phi)^{m+i} \circ f \quad \text{and} \quad \nabla_{\mc{M}} (\phi\circ f)=0 \quad, $$
where of course thanks to the this choice of coordinates we have
$$\pi_N(\bar{\nabla} \phi)=\Big((\bar{\nabla} \phi)^{m+1}, \ldots, (\bar{\nabla} \phi)^{n}\Big)\quad. $$
With this choice of $\phi$, the second summand in the integral of \eqref{eq:akbk} disappears, and thanks to the arbitrariness of the $C^\infty_c(\mc{M})$ functions $\psi^i,\ldots, \psi^{n-m}$ in place of $\pi_N(\bar{\nabla} \phi)$ in \eqref{eq:akbk}, we obtain that not only the integrals but the integrands themselves must vanish, that is $a_k C^k(g, h^\circ, H, \pi_N(\bar{\nabla} \phi))\equiv 0$ on $\mc{M}$, so $I^1_{g,h}(\phi)\equiv 0$.
In particular $\int_{\mc{M}} P_1(g^m,h^m) d \mu_g$ is a global conformal invariant for $m$-dimensional submanifolds which implies that also $\int_{\mc{M}} P_2(g^m,R^m) d \mu_g$ is a global conformal invariant, since by assumption $\int_{\mc{M}} [P_1(g^m,h^m) + P_2(g^m,R^m) ] d \mu_g $ is that as well.
Claim i) follows then directly from Theorem \ref{thm:P(g,h)}. To get claim ii) observe that by construction $P_2(g^m, R^m)$ depends just on the \emph{intrinsic} Riemannian structure $(\mc{M},g)$ and not on the immersion $f$ into an ambient manifold $\bar{\mc{M}}$. Therefore we proved that $\int_{\mc{M}} P_2(g^m, R^m) d \mu_g$ is an intrinsic Riemannian conformally invariant quantity, which enters into the framework of the papers of Alexakis \cite{AlexI}, \cite{AlexII}. More precisely, by applying \cite[Theorem 1]{AlexI}, we obtain claim ii) and the proof is complete.
\end{proof}
\section{Generalized Willmore Energies in Higher Dimensions}\label{sec:genWill}
In this final section, we will introduce a higher dimensional analogue of the Willmore energy (actually we will construct a two-parameters family of such functionals). This new energy is conformally invariant and only attains its strictly positive lower bound at a round sphere. Let us start with some preliminaries about the Willmore functional.
Given an immersion $f:\mc{M}^2 \hookrightarrow \mathbb{R}^3$ of a closed surface $\mc{M}^2$ into the Euclidean space $\mathbb{R}^3$, the Willmore functional $\mathcal W(f)$ is defined by
$$\mathcal W (f):=\int_{\mc{M}} |H|^2 d \mu_g\quad. $$
A natural way to introduce such functional is via conformal invariance: by Gauss-Bonnet Theorem, $\int_{\mc{M}} K d \mu_g$ is a topological hence a fortiori conformal invariant quantity; moreover, by the formula \eqref{eq:CChH}, $|h^\circ|^2 d \mu_g$ is a \emph{pointwise} conformal invariant. It follows that \emph{any} functional
$$ W_{\alpha}(f):=\int_{\mc{M}} \left( K + \alpha |h^\circ|^2 \right) d \mu_g \text{ is a global conformally invariant quantity}$$
in the sense of Definition \ref{def:GCI}.
For immersions into $\mathbb{R}^3$ we have that $K+\frac{1}{2} |h^\circ|^2= |H|^2 $, so
$$\mathcal W(f) = W_{1/2}(f) \text{ is invariant under conformal trasformations of $\mathbb{R}^3$,} $$
that is under Moebius transformations centred out of $f(\mc{M})$.
Observing that $K=\detg(h)$ and that $\frac{1}{2}|h^\circ|^2=-\detg(h^\circ)$, we can also write
\begin{equation}\label{eq:Wdeth}
\mathcal W(f)=\int_{\mc{M}} \left( \detg(h) - \detg(h^\circ) \right) \, d \mu_g \quad.
\end{equation}
Of course the functional $\mathcal W$ is non-negative and vanishes exactly on minimal surfaces, which are therefore points of strict global minimum; the critical points of $\mathcal W$ can therefore be seen in a natural way as ``generalized conformal minimal surfaces''; this was indeed the starting point in the '20ies of the theory of Willmore surfaces by Blaschke \cite{Blaschke}, who was looking for a natural conformally invariant class of immersions which included minimal surfaces. Let us mention that such functional was later rediscovered in the 1960's by Willmore \cite{Will} who proved that round 2-spheres
are the points of strict global minimum for $\mathcal W$.
Motivated by this celebrated two dimensional theory, our goal is to investigate the case of 4-d hypersurfaces in $\mathbb{R}^5$, that is $f:(\mc{M}^4,g)\hookrightarrow (\mathbb{R}^5, \delta_{\mu\nu})$ isometric immersion. \\\\We address the following natural questions:
1) Is it possible to ``perturb'' the Pfaffian of the Riemann tensor of the induced metric $g$ on $\mc{M}^4$ in order to get a conformally invariant functional vanishing on minimal surfaces?
2) Is that functional positive definite? If not, how can we preserve the conformal invariance and make it positive definite?
3) Are round spheres of $\mathbb{R}^5$ strict global minimum of this conformally invariant functional?
\medskip
As in the 2-d case, the starting point is the Gauss-Bonnet Theorem. For $4$-d smooth closed (that is compact without boundary) immersed hypersurfaces $f:(\mc{M}^4,g) \hookrightarrow (\mathbb{R}^5, \delta_{\mu \nu})$, a well known generalization of the Gauss-Bonnet Theorem states that
\begin{equation}\label{eq:GenGB}
\int_{\mc{M}^4} \detg(h) \, d \mu_g=\frac{4 \pi^2 }{3} \; \chi(\mc{M}) = \frac{8 \pi^2 }{3} \deg(\gamma) \quad,
\end{equation}
where $\detg(h):=\det(g^{-1} h)$, $\chi(\mc{M})$ is the Euler Characteristic of $\mc{M}$ and $\gamma: \mc{M}^4 \to S^4\subset \mathbb{R}^5$ is the Gauss map associated to the immersion $f$.
By applying the classical Newton's identities for symmetric polynomials to the symmetric polynomials of the principal curvatures of the immersion $f$, it is an easy exercise to write $\det(g^{-1} h)$ as
\begin{equation}\label{eq:detgh}
\detg(h)=\frac{32}{3} H^4- 4 H^2 |h|^2+\frac{4}{3} H \Tr_g(h^3)+\frac{1}{8}|h|^4-\frac{1}{4} \Tr_g(h^4) \quad,
\end{equation}
where $\Tr_g(h^p):=\Tr[(g^{-1} h)^p]$. By using the orthogonal decomposition $h=\accentset{\circ} h+ H g$, see \eqref{eq:Splith}, iteratively we get that
\begin{eqnarray}
|h|^2&=& \Tr_g(h^2) = |h^\circ|^2 + 4 H^2 \nonumber \\
\Tr_{g}(h^3)&=& \Tr_{g}(\accentset{\circ} h^3)+ 3 H |\accentset{\circ} h|^2 + 4 H^3 \nonumber\\
\Tr_{g}(h^4)&=& \Tr_{g}(\accentset{\circ} h^4)+ 4 H \Tr_{g}(\accentset{\circ} h^3) + 6 H^2 |\accentset{\circ} h|^2 + 4 H^4 \quad, \nonumber
\end{eqnarray}
which, plugged into \eqref{eq:detgh}, give
\begin{equation}\label{eq:detg-1h}
\detg(h)= H^4- \frac{1}{2} H^2 |\accentset{\circ} h|^2+\frac{1}{3} H \Tr_g(\accentset{\circ} h ^3)+\frac{1}{8}|\accentset{\circ} h|^4-\frac{1}{4} \Tr_g(\accentset{\circ} h^4) \quad.
\end{equation}
Inspired by the 2-d case, in particular by formula \eqref{eq:Wdeth}, we wish to make appear a term $\detg(\accentset{\circ} h)$, this is possible thanks to the following lemma.
\begin{lemma}\label{lem:Trdet}
For any $4$-d hypersurface immersed into $\mathbb{R}^5$, we have that
\begin{equation}\label{eq:Trdet}
\Tr_g(\accentset{\circ} h^4)-\frac{1}{2} |\accentset{\circ} h|^4 = -4 \detg(\accentset{\circ} h) \quad.
\end{equation}
\end{lemma}
\begin{proof}
Since $\accentset{\circ} h$ is bilinear symmetric with respect to $g$ it can be diagonalized and its eigenvalues $\{\lambda_i\}_{i=1,\ldots,4} \subset \mathbb{R}$ satisfy $\lambda_4= -\lambda_1-\lambda_2-\lambda_3$, since $\Tr_g(\accentset{\circ} h)=0$. We then have that
\begin{eqnarray}
\Tr_g(\accentset{\circ} h^4)& = &\lambda_1^4+ \lambda_2^4 + \lambda_3^4 + (\lambda_1+\lambda_2+\lambda_3)^4 \nonumber\\
&=& 2 \sum_{i=1}^3 \lambda_i^4 + 4 \sum_{1\leq i\neq j \leq 3} \lambda_i^3 \lambda_j + 6 \sum_{1\leq i\neq j \leq 3} \lambda_i^2 \lambda_j^2 + 12 \sum_{1\leq i\neq j \neq k \leq 3} \lambda_i^2 \lambda_j \lambda_k \quad.\label{eq:Trh4}
\end{eqnarray}
On the other hand,
\begin{eqnarray}
|\accentset{\circ} h|^4 & = & \left[\lambda_1^2+ \lambda_2^2 + \lambda_3^2 + (\lambda_1+\lambda_2+\lambda_3)^2\right]^2 \nonumber\\
&=& 4 \sum_{i=1}^3 \lambda_i^4 + 8 \sum_{1\leq i\neq j \leq 3} \lambda_i^3 \lambda_j + 12 \sum_{1\leq i\neq j \leq 3} \lambda_i^2 \lambda_j^2 + 16 \sum_{1\leq i\neq j \neq k \leq 3} \lambda_i^2 \lambda_j \lambda_k \quad.\label{eq:h4}
\end{eqnarray}
Combining \eqref{eq:Trh4} and \eqref{eq:h4} and recalling that $\lambda_4= -\lambda_1-\lambda_2-\lambda_3$ we can conclude:
$$
\Tr_g(\accentset{\circ} h^4)- \frac{1}{2}|\accentset{\circ} h|^4 = 4 \sum_{1\leq i\neq j \neq k \leq 3} \lambda_i^2 \lambda_j \lambda_k = 4 \lambda_1 \lambda_2 \lambda_3 (\lambda_1+ \lambda_2+\lambda_3)= -4 \detg(\accentset{\circ} h) \quad.
$$
\end{proof}
Thanks to the identity \eqref{eq:Trdet}, we can rewrite \eqref{eq:detg-1h} as
\begin{equation}\label{eq:dethH}
\detg(h)- \detg(\accentset{\circ} h)= H^4- \frac{1}{2} H^2 |\accentset{\circ} h|^2+\frac{1}{3} H \Tr_g(\accentset{\circ} h ^3)\quad.
\end{equation}
Since for even dimensional hypersurfaces in the Euclidean space it is well know that the Pfaffian of the intrinsic Riemann tensor is a multiple of $\detg(h)$, we have just answered to question 1):
\begin{proposition}\label{prop:1}
Given an isometric immersion $f:(\mc{M}^4,g)\hookrightarrow (\mathbb{R}^5, \delta_{\mu\nu})$ of a closed $4$-manifold $(\mc{M}^4,g)$, define
$$\mathcal P (f):= \int_{\mc{M}} \left[ \detg(h)- \detg(\accentset{\circ} h) \right] d \mu_g \quad. $$
Then, the functional $\mathcal P$ is invariant under conformal transformations of $\mathbb{R}^5$ centered off of $f(\mc{M}^4)$.
\\Moreover $\mathcal P $ vanishes identically on minimal hypersurfaces, and the minimal hypersurfaces of $\mathbb{R}^5$ satisfying $\Tr_g(\accentset{\circ} h^3)\equiv 0$ are critical points for $\mathcal P$.
\end{proposition}
\begin{proof}
By using \eqref{eq:dethH}, we have that $\detg(h)- \detg(\accentset{\circ} h) $
clearly vanishes if $H\equiv 0$. Observe that the right hand side of \eqref{eq:dethH} has a linear term in $H$, so a priori a minimal hypersurface is not a critical point for $\mathcal P$, but if we assume also that $\Tr_g(\accentset{\circ} h ^3)$ vanishes, of course we obtain criticality. Since $\left[ \frac{1}{4} \Tr_g(\accentset{\circ} h^4)- \frac{1}{8}|\accentset{\circ} h|^4 \right] d \mu_g$ is a pointwise conformal invariant and $\detg(h)$ is a topological invariant by \eqref{eq:GenGB}, the functional $\mathcal P$ is then invariant under conformal transformation of $\mathbb{R}^5$ preserving the topology of $f(\mc{M})$, that is under conformal transformations of $\mathbb{R}^5$ centered off of $f(\mc{M}^4)$.
\end{proof}
\begin{remark}
Notice that, formally, the integrand of the 4-d functional $\mathcal P$ is exactly the same as the 2-d Willmore functional $\mathcal W$, written as in \eqref{eq:Wdeth}.
\end{remark}
Since in 2-d the quantity $\detg(h)-\detg(h)=H^2$ is non-negative, it is natural to ask if the same is true in the 4-d case, that is if $\detg(g)-\detg(\accentset{\circ} h)$ is non negative. Surprisingly this is not the case, for instance it is not difficult to compute that if the principal curvatures are $1,1,6,6,$ one has $\detg(h)-\detg(\accentset{\circ} h)=-\frac{49}{16}<0$.
A natural question is then if we can manipulate $\mathcal P$ it in order to obtain a new functional which is still conformally invariant but this time is nonnegative definite; this is exactly question 2) above. To this aim observe that, by Young's inequality, we have that
\begin{eqnarray}
-\frac{1}{2} H^2 |\accentset{\circ} h|^2 &\geq& -\frac{\alpha}{4} H^4 - \frac{1}{4 \alpha} |\accentset{\circ} h|^4 \quad \text{for every } \alpha>0 \quad\text{and }\nonumber \\
\frac{1}{3} H \Tr_g(\accentset{\circ} h ^3) &\geq& -\frac{\beta}{12} H^4 - \frac{1}{4 \beta^{1/3}} [\Tr_g(\accentset{\circ} h ^3)]^{4/3} \quad \text{for every } \beta>0 \quad. \nonumber
\end{eqnarray}
The last two lower bounds combined with \eqref{eq:detg-1h} give
\begin{equation}\label{eq:Pab}
P_{\alpha \beta} (g,h):= \detg(h)+ \frac{1}{4} \Tr_g(\accentset{\circ} h^4)+\left(\frac{1}{4 \alpha} -\frac{1}{8} \right) |\accentset{\circ} h|^4 + \frac{1}{4 \beta^{1/3}} [\Tr_g(\accentset{\circ} h ^3)]^{4/3} \geq \left(1- \frac{3\alpha+\beta}{12} \right) H^4.
\end{equation}
We have therefore answered to question 2) above:
\begin{proposition}\label{prop:2}
Given an isometric immersion $f:(\mc{M}^4,g)\hookrightarrow (\mathbb{R}^5, \delta_{\mu\nu})$ of a closed $4$-manifold $(\mc{M}^4,g)$, and let $ P_{\alpha \beta}$ be the following expression
\begin{equation}\label{eq:defPab}
P_{\alpha \beta} (g,h):= \detg(h)+ \frac{1}{4} \Tr_g(\accentset{\circ} h^4)+\left(\frac{1}{4 \alpha} -\frac{1}{8} \right) |\accentset{\circ} h|^4 + \frac{1}{4 \beta^{1/3}} [\Tr_g(\accentset{\circ} h ^3)]^{4/3} \quad.
\end{equation}
Then, the functional
$$\mathcal P_{\alpha \beta} (f):= \int_{\mc{M}} P_{\alpha \beta} (g,h) d \mu_g $$
is invariant under conformal transformations of $\mathbb{R}^5$ centered off of $f(\mc{M}^4)$.
\\Moreover $ P_{\alpha \beta} (g,h) \geq 0$ for every immersed hypersurface $f(\mc{M}^4)\subset \mathbb{R}^5$, provided $\alpha, \beta>0$ satisfy $3\alpha+\beta \leq 12$.
\end{proposition}
We now answer the last question 3) by proving that round spheres are the points of strict global minimum for the functional $\mathcal P_{\alpha, \beta}$, for $\alpha, \beta>0$ with $\alpha \leq 2$. This result may be seen as the 4-d analogue of the celebrated 2-d theorem of Willmore \cite{Will} asserting that the Willmore functional is strictly minimized by embedded round 2-d spheres of $\mathbb{R}^3$.
\begin{theorem} \label{thm:3}
For any isometric immersion $f:(\mc{M}^4,g)\hookrightarrow (\mathbb{R}^5, \delta_{\mu\nu})$ of a closed $4$-manifold $(\mc{M}^4,g)$, let
$$F(g,h):=\detg(h)+Z(\accentset{\circ} h), $$
where $Z(\accentset{\circ} h)$ is a non negative real function of the eigenvalues of $\accentset{\circ} h$, homogeneous of degree $4$ (that is $Z(t \accentset{\circ} h)=t^4 Z(\accentset{\circ} h)$, for every $t \in \mathbb{R}$), and such that $Z(\accentset{\circ} h)=0$ if and only if $\accentset{\circ} h=0$.
Then the functional
$$\mathcal F(f):= \int_{\mc{M}} \left[ \detg(h)+Z(\accentset{\circ} h) \right] \, d\mu_g $$
is invariant under conformal transformations of $\mathbb{R}^5$ centered off of $f(\mc{M}^4)$ and
\begin{align*}
\mathcal F (f) \geq \frac{ 8 \pi ^ 2 } { 3} \quad,
\end{align*}
with equality if and only if $f(\mc{M})$ is embedded in $ \mathbb{R} ^ 5 $ as a round sphere.
\\In particular, this holds for the functional $\mathcal P_{\alpha \beta}$ above provided $\alpha, \beta>0$ and $\alpha \leq 2$.
\end{theorem}
\begin{proof}
Since by assumption the function $Z(\accentset{\circ} h)$ is non-negative, denoting
$$\mc{M}^+:=\{x \in \mc{M}: \detg(h)\geq 0\} \subset \mc{M} \quad,$$
we have
\begin{equation}\label{eq:intPab}
\mathcal F (f)=\int_{\mc{M}} \detg(h)+Z(\accentset{\circ} h) d \mu_g \geq \int_{\mc{M}} \detg(h) \, d \mu_g \geq \int_{\mc{M}^+} \detg(h) \, d \mu_g \quad,
\end{equation}
with equality in the first estimate if and only if $\accentset{\circ} h\equiv 0$, but this happens if and only if $f(\mc{M}^4)$ is embedded as a round sphere (since by assumption $\mc{M}^4$ is closed so $f(\mc{M})$ cannot be an affine hyperplane).
Observing that $|\detg(h)|=|\detg(D\gamma)|= |J(\gamma)|$, where $\gamma: \mc{M}^4 \to S^4$ is the Gauss map of the immersion $f$ and $J(\gamma)$ is its Jacobian, we get
\begin{equation}\label{eq:gammaJ}
\int_{\mc{M}^+} \detg(h) \, d \mu_g = \int_{\mc{M}^+} |J(\gamma)| d \mu_g \geq Vol_{S^4}(\gamma(\mc{M}^+)) \quad,
\end{equation}
where $Vol_{S^4}(\gamma(\mc{M}^+))$ is the volume of $\gamma(\mc{M}^+)\subset S^4$ with respect to the volume form of $S^4$. But now it is well known that $\gamma(\mc{M}^+)=S^4$, since $f(\mc{M}^4)$ is a \emph{closed} hypersurface (the standard argument is to consider, for every $\nu \in S^4$, an affine hyperplane of $\mathbb{R}^5$ orthogonal to $\nu$ and very far from $f(\mc{M}^4)$; then one translates such affine hyperplane towards $f(\mc{M})$, keeping the orthogonality with $\nu$, up to the first tangency point $f(x)$. There, one has $\gamma(x)=\nu$ and moreover, by construction, $f(\mc{M}^4)$ must lie on just one side of its affine tangent space at $f(x)$, namely such translated hyperplane, therefore $\detg(h)|_{x}\geq 0$. Hence $\nu \in \gamma(\mc{M}^+)$).
Combining this last fact with \eqref{eq:intPab} and \eqref{eq:gammaJ}, we get that
$$\mathcal F (f) \geq |S^4| =\frac{8 \pi^2}{3}\quad. $$
Equality in the last formula of course implies equality in \eqref{eq:intPab} and \eqref{eq:gammaJ} but, as already observed above, equality in \eqref{eq:intPab} implies that $f(\mc{M}^4)$ is embedded as a round sphere in $\mathbb{R}^5$.
To see that $\mathcal P_{\alpha, \beta}$ satisfies the assumptions, observe that if $\alpha, \beta>0$ and $\alpha \leq 2$, then
$$
P_{\alpha \beta} (g,h):= \detg(h)+ \frac{1}{4} \Tr_g(\accentset{\circ} h^4)+\left(\frac{1}{4 \alpha} -\frac{1}{8} \right) |\accentset{\circ} h|^4 + \frac{1}{4 \beta^{1/3}} [\Tr_g(\accentset{\circ} h ^3)]^{4/3} \geq \detg(h),
$$
with equality if and only if $\Tr_g(\accentset{\circ} h^4)=|\accentset{\circ} h^2|^2 \equiv 0$; but that happens if and only if $\accentset{\circ} h\equiv 0$.
\end{proof}
By a similar proof one can show the following higher dimensional generalization, however obtaining a comparably explicit formula for $F(g,h)$ is less clear.
\begin{theorem} \label{thm:4}
For any isometric immersion $f:(\mc{M}^{2n},g)\hookrightarrow (\mathbb{R}^{2n+1}, \delta_{\mu\nu})$ of a closed $2n$-manifold $(\mc{M}^{2n},g)$, let
$$F(g,h):=\detg(h)+Z(\accentset{\circ} h), $$
where $Z(\accentset{\circ} h)$ is a non negative real function of the eigenvalues of $\accentset{\circ} h$, homogenous of degree $2n$ (that is $Z(t \accentset{\circ} h)=t^{2n} Z(\accentset{\circ} h)$, for every $t \in \mathbb{R}$), and such that $Z(\accentset{\circ} h)=0$ if and only if $\accentset{\circ} h=0$.
Then the functional
$$\mathcal F(f):= \int_{\mc{M}} \left[ \detg(h)+Z(\accentset{\circ} h) \right] \, d\mu_g $$
is invariant under conformal transformations of $\mathbb{R}^{2n+1}$ centered off of $f(\mc{M}^{2n})$ and
\begin{align*}
\mathcal F (f) \geq \omega_{2n} \quad,
\end{align*}
where $\omega_{2n}$ is the surface area of a unit sphere $ \mathbb{S}^{2n}\subset \mathbb{R} ^{2 n+1}$ with equality if and only if $f(\mc{M})$ is embedded in $ \mathbb{R} ^ {2n+1} $ as a round sphere.
\\Moreover there exists $C=C(n)>0$ such that, setting $Z(\accentset{\circ} h)=C \| \accentset{\circ} h \|^{2n}$, it holds $$F(g,h)= \detg(h)+ C \| \accentset{\circ} h \|^{2n} \geq 0\quad .$$
\end{theorem}
\begin{proof}
The argument for the first part of the theorem is analogous as above, so we will just show the last statement.
\\Using the orthogonal decomposition for the second fundamental form $ h = \accentset{\circ} h + \frac { H}{2n} g $, clearly one has
\begin{align*}
\detg (h) = \detg \left( \accentset{\circ} h + \frac{ H}{ 2n} g \right).
\end{align*}
Now, expanding the determinant, we have that
\begin{align*}
\detg\left( \accentset{\circ} h + \frac{ H}{ 2n} g \right) = \detg (\accentset{\circ} h) + \sum_{k = 1 } ^ { 2n -2 } H ^ k P_{k}(\accentset{\circ} h ) + \frac { H ^ {2n} }{ (2n) ^ { 2n-1}}\quad,
\end{align*}
where $ P_{k}( \accentset{\circ} h ) $ is a complete contraction of order $ 2n-k$. Note that the term corresponding to $k =2n-1$ does not appear as $\tr_g (\accentset{\circ} h) = 0 $. We can now apply Young's inequality to conclude
\begin{align*}
\sum_{ k =1 } ^{2n-2} H ^ k P_{ k } (\accentset{\circ} h ) \geq - \epsilon H ^ {2n} - C_\epsilon \|\accentset{\circ} h \| ^ {2n}.
\end{align*}
Furthermore, as $ \detg ( \accentset{\circ} h) \geq - C ' \| \accentset{\circ} h \| ^ {2n}$ we find
\begin{align*}
\detg (h) \geq \left( \frac { 1 }{( 2n) ^{2n-1} } - \epsilon \right) H ^ {2n} - C_\epsilon' \|\accentset{\circ} h \| ^ {2n } .
\end{align*}
Choosing $\epsilon=(2n)^{-2n}$ we then get that there exists $C=C(n)$ such that $\detg (h) + C \|\accentset{\circ} h \| ^ {2n } \geq 0$, as desired.
\end{proof}
|
\section{Introduction}
\label{SEC:Introduction}
The description of quantum many body systems is an intrinsically
hard problem due to the dimensionality of the Hilbert space,
which depends exponentially on the system size.
Despite the impressive success of numerical and approximate
techniques like Exact Diagonalization \cite{Laeuchli2011},
Quantum Monte Carlo \cite{Metropolis1953, Handscomb1962},
and the Density Matrix Renormalization Group \cite{White1992}, exactly solvable
models remain essential in studying the physics of quantum many body
systems. An analytical solution is not only indispensable for benchmarking
approximate methods, it can also elucidate the general structure of solutions
and underlying physical principles.
With the experimental advances in cooling and
controlling the interactions of atoms, interesting model
systems that have a closed theoretical solution
may even be engineered and studied in the laboratory.
Given the complexity of a generic quantum many body problem, the
direct construction of exactly solvable models appears as an appealing
approach. In the past years, $1+1$ dimensional Conformal Field Theory
(CFT) has proven to be a powerful tool to construct model Hamiltonians
and corresponding ground state wave functions for
continuum and lattice quantum Hall models
\cite{Moore1991, Cirac2010, Nielsen2013a}. In this approach, a CFT is taken as
the starting point to derive a quantum many body system. It can thus
contribute to a systematic understanding of systems that admit a
description in terms of a scale invariant theory, such as critical
systems or systems with topological order that have a gapless edge
spectrum. Theories for which such an analysis was carried out include
the $SU(2)_k$ \cite{Nielsen2011}, $SO(n)_1$ \cite{Tu2013c}, $U(1)_q$
\cite{Tu2014b}, and $SU(n)_1$ \cite{Tu2014, Bondesan2014}
Wess-Zumino-Witten (WZW) theories. Another correspondence between CFT
and a class of continuum and lattice quantum systems in two dimensions was
studied in Ref. \onlinecite{Ardonne2004}. In these models, the ground
state exhibits a $z=2$ Lifshitz scale invariance.
In this paper, we describe a method of constructing excited states
of a spin system that is derived from a CFT.
Correlation functions of fields in the CFT are interpreted as wave functions of states in the
spin system.
Following earlier studies \cite{Nielsen2011, Nielsen2012},
we consider the $SU(2)_1$ WZW model and
define a parent Hamiltonian for the state that
corresponds to the correlator of $N$ primary fields.
In the case of periodic boundary conditions in one
dimension, which we focus on here, the resulting spin system is equivalent to
the Haldane-Shastry model \cite{Haldane1988, Shastry1987}.
Based on solutions to the Calogero-Sutherland model\cite{Calogero1969,
Calogero1971, Sutherland1971, Sutherland1971a}, excited states of the
Haldane-Shastry Hamiltonian were first constructed as polynomials in
the particle basis\cite{Haldane1988, Haldane1991}. It was realized
later\cite{Haldane1992, Talstra1995PhD} that these states are the
highest weight states of the Yangian algebra, a hidden symmetry of the
Haldane-Shastry Hamiltonian. Given the close relation between the
Haldane-Shastry model and the $SU(2)_1$ WZW CFT, it was conjectured in
Ref.~\onlinecite{Haldane1992} that there should be a correspondence
between the states in the CFT and those of the spin chain. Here we
show that this is indeed the case. Since our ansatz is
based on the $SU(2)$ currents, it is manifestly $SU(2)$ invariant.
We relate our results both to the Yangian highest weight states and to
the spinon basis of CFT states \cite{Bouwknegt1994}.
Our ansatz mirrors the structure of the CFT, with the ground state
corresponding to the CFT vacuum and the excited states corresponding
to CFT descendant states. More precisely, we construct the excited
states from the ground state by insertion of current operator modes
into correlation functions of primary fields. We show that the
Hamiltonian of the spin system is block diagonal in these states, with
each block corresponding to a fixed number of inserted current
operators. This allows us to obtain excited states by successively
adding current operators and diagonalizing the blocks. As the
Hamiltonian, the Yangian is closed in the subspaces of states with a certain number
of current operators. This implies that one can construct representations
of the Yangian algebra in these subspaces. We explicitly construct
the highest weight states for the analytically obtained eigenstates. Our ansatz for
the excited states and its relation to descendant states of the CFT
is summarized in Table \ref{table:structure-of-excited-states}.
\begin{table}[htb]
\caption{\label{table:structure-of-excited-states}
Structure of the states in the CFT and the spin system. The
ground state of the spin system is constructed from the product of $N$
primary fields $\Phi_{\mathbf{s}}(\mathbf{z}) = \phi_{s_1}(z_1) \dots \phi_{s_N}(z_N)$
and corresponds to the CFT vacuum $\Ket{0}$. Excited states are
constructed from the ground state by insertion of current operator
modes $J^{a_k}_{-1}\dots J^{a_1}_{-1}$.
}
\centering
\begin{tabular}{llcl}
& CFT & & Spin system\\
\hline
\cellblock{Ground\\state} & $\Ket{0}$ & $\leftrightarrow$
& $\textcolor{blue}{\langle 0| \Phi_{\mathbf{s}}(\mathbf{z})} | 0 \rangle$\\
& $\downarrow$ & & $\downarrow$\\
\cellblock{Excited\\states} & $\textcolor{red}{(J^{a_k}_{-1} \dots J^{a_1}_{-1})(0)} \Ket{0}$
& $\leftrightarrow$ & $\textcolor{blue}{\langle 0| \Phi_{\mathbf{s}}(\mathbf{z})} \textcolor{red}{(J^{a_k}_{-1} \dots J^{a_1}_{-1})(0)} | 0 \rangle$
\end{tabular}
\end{table}
We perform the diagonalization of the Hamiltonian analytically for up
to eight current operators, construct Yangian highest weight states
in the obtained eigenstates, and confirm numerically for small system
sizes that the complete spectrum can be obtained in this way. We show
that this construction can be done both for an even and for an odd
number of spins in the chain.
This paper is structured as follows: In Section
\ref{SEC:states-from-conformal-blocks}, we review some properties of
the $SU(2)_1$ WZW model and describe the states that form the basis of
our construction. We introduce the parent Hamiltonian in Section
\ref{SEC:Parent-Hamiltonian-and-Spectrum} and construct excited
states, both analytically (Section \ref{SEC:Eigenstates-Analytically})
and numerically (Section \ref{SEC:Numerics}). We conclude in Section
\ref{SEC:Conclusion}.
\section{States From Conformal Blocks}
\label{SEC:states-from-conformal-blocks}
In this section, we briefly review some properties of the $SU(2)_1$ WZW model
and describe the correspondence between conformal blocks and spin system
wave functions.
In addition to the identity, the model has one primary field
with scaling dimension $h = 1/4$, the vertex operator $\V{s}{z}$. It
can be constructed from the chiral part $\varphi(z)$ of a free, massless boson as
\begin{align}
\label{eq:vertex-operator-from-free-boson}
\V{s}{z} & = e^{i \pi (q-1) (s+1) / 2} :e^{i s \varphi(z)/ \sqrt{2}}:.
\end{align}
Here $s=\pm1$ corresponds to the two components of the vertex
operator, and the colons denote normal ordering. (The holomorphic
field $\phi_s(z)$ has an anti-holomorphic counterpart
$\bar{\phi}_s(\bar{z})$. In the following we only consider the
holomorphic sector.) The value $q \in \{0, 1\}$ corresponds to the
two sectors of the CFT: $q=0$ if the operator $\V{s}{z}$ acts on a
state that has an even number of $h=1/4$ primary fields and $q=1$ for
a state with an odd number, respectively.
In addition to conformal invariance, the $SU(2)_1$ WZW model has an $SU(2)$
symmetry, which is generated by the current operator $J^a(z)$.
Its Laurent expansion defines modes $J^a_n$,
\begin{align}
J^a(z) &= \sum^{\infty}_{n = -\infty} z^{-n - 1} J^a_n,
\end{align}
where $a \in \{x, y, z\}$.
They satisfy the Kac-Moody algebra \cite{Knizhnik1984}
\begin{align}
\label{eq:Kac-Moody-current-algebra}
\Comm{J^a_m}{J^b_n} &= i \varepsilon_{abc} J^c_{m+n} + \frac{m}{2} \delta_{ab} \delta_{m+n, 0},
\end{align}
with $\varepsilon_{abc}$ being the Levi-Civita symbol and $\delta_{ab}$ the Kronecker delta.
A primary field $\V{s}{z}$ transforms covariantly with respect to conformal
and $SU(2)$ transformations. The latter is expressed by
the operator product expansion (OPE) between the current $J^a(z)$
and $\V{s}{z}$ \cite{DiFrancesco1997},
\begin{align}
\label{eq:ope-j-with-vertex-op}
J^a(z) \V{s}{w} &\sim -\sum_{s'}\frac{(t^a)_{s s'}}{z-w} \V{s'}{w}.
\end{align}
Here $t^a$ are the $SU(2)$ spin operators. They are related
to the Pauli matrices $\sigma^a$ by $t^a = \sigma^a / 2$.
The modes of the current operator
give rise to a tower of descendant states
\begin{align}
\label{eq:descendants-in-CFT}
&(J^{a_k}_{-1} \dots J^{a_1}_{-1})(0) \Ket{0},
\end{align}
with $\Ket{0}$ being the CFT vacuum. States of this form build up the spectrum of the CFT \cite{Gepner1986}.
Note that it suffices to consider states obtained from the $n=-1$ mode of
$J^a(z)$. This is so because one can successively rewrite a higher
order mode $J^a_{-n}$, $n > 0$, in terms of the lower order modes $J^a_{-n+1}$
and $J^a_{-1}$ by means of the Kac-Moody algebra (cf. Eq. \eqref{eq:Kac-Moody-current-algebra}),
\begin{align}
\label{eq:elim-higher-order-J}
J^a_{-n} &= \frac{i}{2} \varepsilon_{abc} \Comm{J^{c}_{-1}}{J^{b}_{-n+1}}, n \neq 0.
\end{align}
In this work, we show that the spectrum of the spin systems
organizes in the same way in terms of current operators.
The excited states we obtain are linear combinations
of conformal blocks containing current operator
modes $J^{a_k}_{-1} \dots J^{a_1}_{-1}$.
We give a summary of these states in Table \ref{table:summary-of-states}
and describe the construction of states from conformal blocks in the next subsections.
\begin{table}[htb]
\caption{\label{table:summary-of-states}
Summary of the different towers of states obtained by insertion of
current operator modes. $\Phi_{\mathbf{s}}(\mathbf{z})$ denotes the product of $N$
primary fields, $\Phi_{\mathbf{s}}(\mathbf{z}) = \phi_{s_1}(z_1) \dots \phi_{s_N}(z_N)$.
Using the OPE between $\phi_s(z)$ and $J^a(z)$, the wave functions for
these states can be written as the application of $k$ Fourier transformed
spin operators to the state without current operators, as
explained in Section \ref{SEC:States-from-Vertex-and-Current-Operators}.
}
\centering
\begin{tabular}{lll}
& Tower of states & See Eq. \\
\hline
$N$ even & $\Braket{ \Phi_{\mathbf{s}}(\mathbf{z}) (J^{a_k}_{-1}\dots J^{a_1}_{-1})(0)}$ & \eqref{eq:def-J-states} \\
$N$ even & $\Braket{ \phi_{s_\infty}(\infty) \Phi_{\mathbf{s}}(\mathbf{z}) (J^{a_k}_{-1}\dots J^{a_1}_{-1} \phi_{s_0})(0)}$ & \eqref{eq:def-tower-above-psi0inf-sgl} \\
$N$ odd & $\Braket{\phi_{s_\infty}(\infty) \Phi_{\mathbf{s}}(\mathbf{z}) (J^{a_k}_{-1}\dots J^{a_1}_{-1})(0)}$ & \eqref{eq:tower-above-psi0odd} \\
$N$ odd & $\Braket{\Phi_{\mathbf{s}}(\mathbf{z}) (J^{a_k}_{-1}\dots J^{a_1}_{-1} \phi_{s_0})(0)}$ & \eqref{eq:alpha0odd-tower}
\end{tabular}
\end{table}
\subsection{State Obtained from a String of Vertex Operators}
\label{SEC:State-from-Vertex-Operators}
We consider an even number of vertex operators $\Vi{i}$ ($i = 1, \dots, N$)
that each transform under a representation of $SU(2)$ generated by
spin operators $t^a_i$, as expressed by the OPE of
Eq. \eqref{eq:ope-j-with-vertex-op}. The key idea is to view
the correlation function of vertex operators in the CFT
as the wave function of a system of spin $1/2$
degrees of freedom on a lattice,
\begin{align}
\label{eq:def-ground-state}
\Ket{\psi_0} &= \sum_{s_1 \dots s_N} \psi_0(s_1, \dots, s_N) \Ket{s_1, \dots, s_N},
\intertext{where $\psi_0(s_1, \dots, s_N)$ is given by}
\psi_0(s_1, \dots, s_N) &= \Braket{\V{s_1}{z_1} \dots \V{s_N}{z_N}}.
\end{align}
Here,
\begin{align}
\label{eq:vertex-operators-with-sign-factor}
\Vi{j} &= e^{\pi i (j-1) (s_j + 1) / 2} :e^{i s_j \varphi(z_j)/ \sqrt{2}}:,
\end{align}
and $\Ket{s_1, \dots, s_N}$ with $s_i = \pm 1$ is the tensor product
of eigenstates $\Ket{s_i }$ of the $z$-component of the spin operator
$t^z_i$. The coordinates $z_i$ define the lattice positions in the
complex plane and are kept fixed. In contrast to the continuum case,
there is no spatial degree of freedom in the basis states $\Ket{s_1,
\dots, s_N}$ and therefore no integral over the positions in
Eq. \eqref{eq:def-ground-state}. This is why we use the notation
$\psi_0(s_1, \dots s_N)$ without the coordinates $z_i$ for the spin
wave function.
The correlation function of $N$ vertex operators is given by
\cite{DiFrancesco1997}
\begin{align}
\label{eq:vertex-op-correlator}
\psi_0(s_1, \dots, s_N) &= \Braket{\V{s_1}{z_1} \dots\V{s_N}{z_N}} \\
&= \delta_{\mathbf{s}} \chi_{\mathbf{s}} \prod_{i < j}^{N} (z_i - z_j)^{s_i s_j / 2}\notag,
\end{align}
where $\delta_{\mathbf{s}}$ is $1$ if $\sum_{i=1}^N s_i = 0$ and $0$
otherwise. $\chi_\mathbf{s}$ is the Marshall sign factor,
\begin{align}
\label{eq:marshall-sign-factor}
\chi_{\mathbf{s}} &= \prod_{p=1}^{N} e^{i \pi (p-1) (s_p+1)/2},
\end{align}
which ensures that the state $\psi_0$ is a spin singlet\cite{Cirac2010, Nielsen2012},
\begin{align}
\label{eq:psi0-singlet-property}
T^a \psi_0 &= 0, \quad T^a = \sum_{i=1}^{N} t^a_i.
\end{align}
Note that the condition $\delta_{\mathbf{s}}$ requires the
number of primary fields $\V{s_j}{z_j}$ in the correlator to be even.
The positions $z_j$ can, in principle, assume any value on the complex
plane. In the following we mostly consider $N$ spins uniformly distributed
on the circle,
\begin{align}
\label{eq:zi-on-unit-circle}
z_j &= z^j, \quad\text{with } z = e^{2 \pi i/N}.
\end{align}
In this case, the state $\psi_0$ has a momentum of $\pi$ if $N/2$ is odd and $0$
if $N/2$ is even\cite{Bernevig2001}, see also Appendix \ref{SEC:Lattice-momentum}.
\subsection{States Obtained from Vertex and Current Operators}
\label{SEC:States-from-Vertex-and-Current-Operators}
In analogy to the tower of CFT descendant states $(J^{a_k}_{-1} \dots J^{a_1}_{-1})(0) \Ket{0}$,
we define a tower of spin states by insertion of
modes of the current operator $J^a(z)$ into a correlation function
of vertex operators,
\begin{align}
\label{eq:def-J-states}
&\Ket{\psi_{a_k \dots a_1}} = \sum_{s_1 \dots s_N} \psi_{a_k \dots a_1}(s_1, \dots, s_N) \Ket{s_1, \dots, s_N}, \notag \\
\intertext{where $\psi_{a_k \dots a_1}(s_1, \dots, s_N)$ is given by}
&\psi_{a_k \dots a_1}(s_1, \dots, s_N) \notag \\
& \quad= \Braket{\V{s_1}{z_1} \dots \V{s_N}{z_N} (J^{a_k}_{-1} \dots J^{a_1}_{-1})(0)}.
\end{align}
It is possible to obtain the states $\psi_{a_k \dots a_1}$ by applying spin operators to $\psi_0$.
To see this, we first note that
\begin{equation}
\langle \Phi_{\mathbf{s}}(\mathbf{z}) (J^a_{-1}B)(0)\rangle =\frac{1}{2\pi i}\oint_0\frac{\mathrm{d} w}{w}\langle \Phi_{\mathbf{s}}(\mathbf{z}) J^a(w)B(0)\rangle,
\end{equation}
with $\Phi_{\mathbf{s}}(\mathbf{z}) \equiv \Vi{1} \dots \Vi{N}$ and $B$ being an arbitrary operator.
Applying this relation to the definition of $\psi_{a_k \dots a_1}$ we obtain
\begin{align}
\label{eq:derivation_of_Jpsi}
&\psi_{a_k \dots a_1}(s_1, \dots, s_N) \notag \\
&= \Braket{\Vi{1} \dots \Vi{N} (J^{a_k}_{-1} \dots J^{a_1}_{-1})(0)} \notag \\
&= \frac{1}{2\pi i}\oint_0\frac{\mathrm{d} w}{w}\langle \Phi_{\mathbf{s}}(\mathbf{z}) J^{a_k}(w)(J^{a_{k-1}}_{-1}\ldots J^{a_1}_{-1})(0)\rangle \notag \\
&=-\frac{1}{2\pi i}\sum_{j=1}^N\oint_{z_j}\frac{\mathrm{d} w}{w}\langle \Phi_{\mathbf{s}}(\mathbf{z}) J^{a_k}(w)(J^{a_{k-1}}_{-1}\ldots J^{a_1}_{-1})(0)\rangle. \\
\intertext{Using the OPE between a current operator and a primary field (cf. Eq. \eqref{eq:ope-j-with-vertex-op}), we get}
&\psi_{a_k \dots a_1}(s_1, \dots, s_N) \notag \\
&= \frac{1}{2\pi i}\sum_{j=1}^N\oint_{z_j}\frac{\mathrm{d} w}{w}\frac{t_j^{a_k}}{w-z_j} \langle \Phi_{\mathbf{s}}(\mathbf{z}) (J^{a_{k-1}}_{-1}\ldots J^{a_1}_{-1})(0)\rangle \notag \\
&= \sum_{j=1}^N\frac{t_j^{a_k}}{z_j} \psi_{a_{k-1}\ldots a_1}.
\end{align}
Successive application of the same argument results in
\begin{align}
\label{eq:result-Jpsi-from-Gpsi}
\psi_{a_k \dots a_1} &= \left(\sum_{j_k=1}^{N} \frac{t^{a_k}_{j_k}}{z_{j_k}}\right) \dots \left(\sum_{j_1=1}^{N} \frac{t^{a_1}_{j_1}}{z_{j_1}}\right) \psi_0.
\end{align}
If the positions $z_j$ are uniformly distributed on the circle, we can express this
result in terms of Fourier transformed spin operators $u^{a}_{l}$,
\begin{align}
\label{eq:ft-spin-ops-definition}
u^a_{l} &\equiv \sum_{j=1}^{N} t^a_j e^{2 \pi i j l / N}.
\end{align}
With $z=e^{2\pi i /N}$ in Eq. \eqref{eq:result-Jpsi-from-Gpsi}, we have
\begin{align}
\label{eq:def-I-spin-ops}
\psi_{a_k \dots a_1} &= \It{a_k}{-1} \dots \It{a_1}{-1} \psi_0.
\end{align}
Therefore, each additional insertion of $J^{a}_{-1}$ changes the momentum
of the state by $2 \pi / N$.
\subsection{States with Additional Spins at Zero and Infinity}
\label{SEC:definition-psi-zero-inf}
We define an additional class of states
by inserting two extra vertex operators into the correlator $\Braket{\Vi{1} \dots \Vi{N}}$,
one at $z=0$ and one at $z=\infty$,
\begin{align}
\label{eq:def-0inf-corr}
&\Ket{\psi^{s_0, s_\infty}_0} = \sum_{s_1, \dots, s_N} \psi_0^{s_0, s_\infty}(s_1, \dots, s_N) \Ket{s_1, \dots, s_N}, \\
&\psi_0^{s_0, s_\infty}(s_1, \dots, s_N)\notag \\
&\quad\quad= \Braket{\V{s_\infty}{\infty} \Vi{1} \dots \Vi{N} \V{s_0}{0}}. \notag
\end{align}
In the Riemann sphere picture, the additional spins are added
at the south and north pole, respectively, while the $N$ spins at the
unit circle are located at the equator (see Fig. \ref{fig:Riemann-sphere}).
\begin{figure}[htb]
\centering
\includegraphics[width=.6\linewidth]{sphere.pdf}
\caption{\label{fig:Riemann-sphere}
(Color online) The Riemann sphere with $N$ spins at the equator (unit circle) and two
additional spins, one at the north pole ($z=0$) and one at the south pole ($z=\infty$).
}
\end{figure}
The wave function is
\begin{align}
\label{eq:wave-function-alpha0}
\psi^{s_0, s_\infty}_0(s_1, \dots, s_N) &\propto \delta_{\bar{\mathbf{s}}} (-1)^{s_0 (1-s_\infty)/2} \chi_{\mathbf{s}} \prod_{n=1}^{N} z_n^{s_0 s_n / 2} \notag \\
&\quad \times \prod_{n < m}^N (z_n - z_m)^{s_n s_m / 2},
\end{align}
where $\delta_{\bar{\mathbf{s}}} = 1$ for $s_0 + s_\infty + \sum_{i=1}^N s_i = 0$ and
$\delta_{\bar{\mathbf{s}}} = 0$ otherwise.
Note that the extra fields inserted at zero and infinity
are primary fields. If we insert additional current operators,
we generate descendant states. We thus define
a tower of states on top of $\psi^{s_0 s_\infty}_0$,
\begin{align}
\label{eq:def-tower-above-psi0inf-sgl}
&\psi_{a_k \dots a_1}^{s_0, s_\infty}(s_1, \dots, s_N) \\
&= \Braket{\V{s_\infty}{\infty} \Vi{1} \dots \Vi{N} (J^{a_k}_{-1} \dots J^{a_1}_{-1} \Vs{s_0})(0)}. \notag
\end{align}
This ansatz thus corresponds to
the tower of CFT descendant states $(J^{a_k}_{-1} \dots J^{a_1}_{-1} \phi_s)(0) \Ket{0}$.
An argument similar to the one given for $\psi_{a_k \dots a_1}$ shows that
\begin{align}
\label{eq:Jpsi-zero-inf-from-spin-operators}
\psi_{a_k \dots a_1}^{s_0, s_\infty} &= \left(\frac{t^{a_k}_{\infty}}{z_\infty} + \sum_{j_k=1}^{N} \frac{t^{a_k}_{j_k}}{z_{j_k}}\right) \dots \notag \\
&\quad \times \left(\frac{t^{a_1}_{\infty}}{z_\infty} + \sum_{j_1=1}^{N} \frac{t^{a_1}_{j_1}}{z_{j_1}}\right) \psi^{s_0, s_\infty}_0.
\end{align}
Note that the terms $t^{a_j}_{\infty}/z_{\infty}$ do not contribute
in the limit $z_\infty \to \infty$. On the unit circle, we have
\begin{align}
\label{eq:Jpsi-zero-inf-from-FT-spin-operators}
\psi_{a_k \dots a_1}^{s_0, s_\infty} &= \It{a_k}{-1} \dots \It{a_1}{-1} \psi^{s_0, s_\infty}_0
\end{align}
in terms of Fourier transformed spin operators.
\subsection{Odd Number of Spins}
\label{SEC:odd-number-of-spins}
A correlation function of vertex operators $\Braket{\Vi{1} \dots \Vi{N}}$ is only non-zero
if the sum $\sum_{i=1}^{N} s_i $ vanishes (cf. Eq. \eqref{eq:vertex-op-correlator}).
This property, the \emph{charge neutrality condition},
implies that the vertex operators in the correlator need to have a net charge of zero.
As a consequence, the number of vertex operators, and therefore the number
of spins, needs to be even.
We can, however, still consider a model with an odd number
of spins at the circle, by adding an extra vertex operator
that compensates the excess charge at the circle.
Inserting an additional vertex operator at $z=\infty$,
we obtain the state
\begin{align}
\label{eq:state-with-vertex-operator-at-infinity}
&\psi^{s_\infty}_0 (s_1, \dots, s_N) \notag\\
&\quad= \Braket{\V{s_\infty}{\infty} \Vi{1} \dots \Vi{N}} \notag \\
&\quad\propto \delta_{\bar{\mathbf{s}}} (-1)^{(s_\infty+1)/2} \chi_{\mathbf{s}} \prod_{i < j}^{N} (z_i - z_j)^{s_i s_j / 2},
\end{align}
where $\delta_{\bar{\mathbf{s}}} = 1$ if $s_\infty + \sum_{i=1}^N s_i
= 0$ and $\delta_{\bar{\mathbf{s}}} = 0$ otherwise.
The wave function $\psi^{s_\infty}_0$ has spin $1/2$,
\begin{align}
\label{eq:psi0add-has-spin-1-2} T^a T^a \psi^{s_\infty}_0 = \frac{3}{4} \psi^{s_\infty}_0, \quad \text{with } T^a = \sum_{i=1}^{N} t^a_i.
\end{align}
This is a consequence of $\psi^{s_\infty}_0$ being a singlet of the total spin including the point
$z=\infty$, $(t^a_\infty + T^a) \psi^{s_\infty}_0 = 0$.
As in the case of an even number of spins,
we define a tower of states by insertion of current operators,
\begin{align}
\label{eq:tower-above-psi0odd}
&\psi^{s_\infty}_{a_k \dots a_1}(s_1, \dots, s_N) \notag \\
&\quad= \Braket{\V{s_\infty}{\infty} \Vi{1} \dots \Vi{N} (\Ja{k} \dots \Ja{1})(0)}.
\end{align}
We obtain a second class of states by inserting the additional vertex operator at $z=0$
instead of $z=\infty$,
\begin{align}
\label{eq:alpha0odd-definition}
&\psi^{s_0}_0(s_1, \dots, s_N) \notag\\
&\quad= \Braket{\Vi{1} \dots \Vi{N} \V{s_0}{0}} \notag \\
&\quad\propto \delta_{\bar{\mathbf{s}}} \chi_{\mathbf{s}} \prod_{i=1}^{N} z_i^{s_0 s_i/2} \prod_{i < j}^{N} (z_i - z_j)^{s_i s_j / 2},
\end{align}
and the corresponding tower
\begin{align}
\label{eq:alpha0odd-tower}
&\psi^{s_0}_{a_k, \dots, a_1}(s_1, \dots, s_N) \notag \\
&\quad = \Braket{\Vi{1} \dots \Vi{N} (\Ja{k} \dots \Ja{1} \Vs{s_0})(0)}.
\end{align}
On the unit circle, the two towers of states can be written as
\begin{align}
\label{eq:odd-N-tower-in-terms-of-FT-spin-operators-psi0odd}
\psi^{s_\infty}_{a_k, \dots, a_1}(s_1, \dots, s_N) &= \Ia{k} \dots \Ia{1} \psi^{s_\infty}_0
\end{align}
and
\begin{align}
\label{eq:odd-N-tower-in-terms-of-FT-spin-operators-alpha0odd}
\psi^{s_0}_{a_k, \dots, a_1}(s_1, \dots, s_N) &= \Ia{k} \dots \Ia{1} \psi^{s_0}_0,
\end{align}
respectively.
Comparing the wave functions of Eq. \eqref{eq:state-with-vertex-operator-at-infinity}
and Eq. \eqref{eq:alpha0odd-definition} with the case of an even number of spins
(Eq. \eqref{eq:vertex-op-correlator} and \eqref{eq:wave-function-alpha0}),
we conclude that $\psi^{s_\infty}_0$ is the analog of $\psi_0$ and $\psi^{s_0}_0$
of $\psi^{s_0 s_\infty}_0$, respectively.
\section{Parent Hamiltonian and Spectrum}
\label{SEC:Parent-Hamiltonian-and-Spectrum}
In this section, we introduce the parent Hamiltonian of $\psi_0$,
discuss its equivalence to the Haldane-Shastry model and
construct its spectrum from CFT current operators. Furthermore,
we relate our ansatz to the multiplets of the Yangian algebra
and the spinon construction of the CFT Hilbert space. From now
on we assume a uniform, one-dimensional lattice with
periodic boundary conditions,
\begin{align}
\label{eq:fix-z-to-unit-circle}
z_j = e^{2 \pi i j / N}
\end{align}
with $j \in \{1, \dots, N\}$.
\subsection{Parent Hamiltonian}
\label{SEC:Parent-Hamiltonian}
In earlier work on the Haldane-Shastry model \cite{Shastry1992},
an operator $\Co{a}{i}$ was constructed that annihilates the wave function $\psi_0$,
$\Co{a}{i} \psi_0 = 0$. This operator was later obtained in a more general setting from
the $SU(2)_1$ WZW model using null vectors\cite{Nielsen2011}.
In terms of spin operators,
\begin{align}
\label{eq:definition-of-C}
\Co{a}{i} &= \frac{2}{3} \sum_{j(\neq i)} w_{ij} \left(t^a_j + i \varepsilon_{abc} t^b_i t^c_j \right),
\end{align}
with
\begin{align}
\label{eq:definition-of-w}
w_{ij} &\equiv \frac{z_i + z_j}{z_i - z_j}.
\end{align}
We have used the notation $\sum_{i(\neq j)}$ for a sum over
all $i \in \{1, \dots, N\} \setminus \{j\}$.
This allows for the definition of a parent Hamiltonian $H$ of
$\psi_0$ \cite{Nielsen2011},
\begin{align}
\label{eq:def-Hamiltonian}
H &= \frac{1}{4} \sum_{i=1}^N (\Co{a}{i})^\dagger \Co{a}{i}.
\end{align}
Note that $H$ is
positive semidefinite and $\psi_0$ is an eigenstate of $H$ with zero energy.
It is known\cite{Nielsen2011} that $H$ is
closely related to the Haldane-Shastry Hamiltonian $H_{\text{HS}}$,
if the spins are uniformly distributed on the circle,
\begin{align}
\label{eq:H-and-HHS}
H_{\text{HS}} &= \frac{1}{2} \sum_{i \neq j} \frac{t^a_i t^a_j}{\mathrm{sin}^2\left(\frac{(i-j) \pi}{N}\right)}
= H + \frac{N+1}{6} T^a T^a + E_0.
\end{align}
Here $T^a = \sum_{i=1}^{N} t^a$ is the total spin and $E_0 =
-(N^3+5N)/24$ the ground state energy of the Haldane-Shastry
Hamiltonian. Note that $\psi_0$ is annihilated by $H$ and also by
$T^aT^a$ in Eq. \eqref{eq:H-and-HHS} since it is a singlet. Therefore,
$\psi_0$ is the ground state of the Haldane-Shastry Hamiltonian.
From now on, we will work with the Hamiltonian
\begin{align}
\label{eq:def-full-H} \mathcal{H} &= H + \frac{N+1}{6} T^a T^a,
\end{align}
dropping the constant $E_0$.
\subsection{Block Diagonal Form of the Hamiltonian}
\label{SEC:block-diagonal-form-of-the-Hamiltonian}
We now systematically construct excited states of $\mathcal{H}$ from
conformal correlation functions. Specifically, we build the
excited states as linear combinations of the states $\psi_{a_k \dots a_1}$
(cf. Eq. \eqref{eq:vertex-op-correlator} and \eqref{eq:result-Jpsi-from-Gpsi}).
The key to this construction is that the Hamiltonian does not couple
states with a fixed number of current operators to states with a
different number of current operators, i.e. the Hamiltonian is
block-diagonal in this basis.
Therefore, we can diagonalize the Hamiltonian in the subspaces of
states with a certain number of current operators.
It is not necessary to construct the Hamiltonian in the full Hilbert space of dimension $2^N$ in
order to find eigenstates beyond the ground state. Rather, we obtain
eigenstates by successively adding current operators and
diagonalizing the blocks.
Let us now show that $\mathcal{H} \psi_{a_k \dots a_1}$ is a linear combination of states
obtained from $\psi_0$ by insertion of $k$ current operators.
Recall that the insertion of $k$ current operator modes $J^{a_j}_{-1}$ $(j=1, \dots, k)$
into the correlation function of vertex operators is equivalent to the
successive application of Fourier transformed spin operators $\It{a_j}{-1}$ to
the ground state $\psi_0$ (cf. Eq. \eqref{eq:def-I-spin-ops}).
Therefore, we have computed the commutator between $\mathcal{H}$ and $\It{a}{-1}$ by an
explicit expansion of the Hamiltonian
in terms of Fourier modes
$\It{a}{l}$ (cf. Section
\ref{SEC:Commutator-H-J} of the Appendix). The result of this
calculation is
\begin{align}
\label{eq:commutator-Hf-with-I}
\Comm{\mathcal{H}}{\It{a}{-1}} &= (N-1) \It{a}{-1} + \sum_{i=1}^{N}\frac{3}{2} \frac{\Co{a}{i}}{z_i} + i \varepsilon_{abc} \It{b}{-1} T^{c}.
\end{align}
From this we can already conclude that the energy of a state
with one current operator is $N-1$,
\begin{align}
\label{eq:energy-single-Jpsi}
\mathcal{H} \It{a}{-1} \psi_0 &= \Comm{\mathcal{H}}{\It{a}{-1}}\psi_0 = (N-1) \It{a}{-1} \psi_0,
\end{align}
since $\mathcal{H}$, $\Co{a}{i}$, and $T^c$ annihilate the ground state $\psi_0$.
We need to know how $\Co{a}{i}$ acts on $\psi_{a_k\dots a_1}$ to determine the energy
of states with more than one current operator.
As we show in Section \ref{SEC:decoupling-equation-for-Jpsi} of the Appendix,
\begin{align}
\label{eq:decoupling-equation-Jpsi}
&\mathcal{C}_i^a \psi_{a_k\dots a_1} \notag \\
&= \sum_{q=1}^k\frac{(K_{a_q}^a)_i}{z_i} \psi_{a_k\dots a_{q+1}a_{q-1}\dots a_1} \notag \\
& \quad+(K_{b}^a)_i T^b\psi_{a_k\dots a_1} \notag \\
& \quad +2 (K_{b}^a)_i \sum_{q=2}^k\sum_{n=0}^{q-1} \frac{i\varepsilon_{ba_qc}}{z_i^{n+1}} \notag \\
& \quad\quad \times \langle \Phi_{\mathbf{s}}(\mathbf{z}) (J_{-1}^{a_k}\dots J_{-1}^{a_{q+1}} J_n^{c}J_{-1}^{a_{q-1}} \dots J_{-1}^{a_1})(0)\rangle,
\end{align}
with
\begin{align}
\label{eq:decoupling-equation-details}
\Phi_{\mathbf{s}}(\mathbf{z}) &= \Vi{1} \dots \Vi{N}, \quad \text{and} \\
(K^{a}_b)_{i} &= \frac{2}{3} (\delta_{ab} - i \varepsilon_{abc} t^c_i).
\end{align}
Combining Eq. \eqref{eq:commutator-Hf-with-I} and
Eq. \eqref{eq:decoupling-equation-Jpsi}, we obtain
\begin{align}
\label{eq:derivation-H-Jpsi}
&\mathcal{H} \psi_{a_k\dots a_1} \notag \\
&= \sum_{r=1}^{k} \Ia{k} \dots \Ia{r+1} \Comm{\mathcal{H}}{\Ia{r}} \Ia{r-1} \dots \Ia{1} \psi_0 \notag \\
&= k (N-1) \psi_{a_k\dots a_1} + \sum_{2 \le q<r \le k} \sum_{n=0}^{q-1} F_{a_k \dots a_1}^{qr,n}\notag \\
& \quad + \sum_{1 \le q<r \le k} \Big(2 \psi_{a_k\dots a_{r+1} a_q a_{r-1} \dots a_{q+1} a_{r} a_{q-1} \dots a_1} \notag \\
&\phantom{\quad + \sum_{q<r}^{k} \Big(} - 2 \delta_{a_r a_q} \psi_{a_k\dots a_{r+1} c a_{r-1} \dots a_{q+1} c a_{q-1} \dots a_1} \notag \\
&\phantom{\quad + \sum_{q<r}^{k} \Big(} + \psi_{a_k \dots a_{r+1} a_{q} a_{r} a_{r-1} \dots a_{q+1} a_{q-1} \dots a_{1}} \notag \\
&\phantom{\quad + \sum_{q<r}^{k} \Big(} - \psi_{a_k \dots a_{r+1} a_{r} a_{q} a_{r-1} \dots a_{q+1} a_{q-1} \dots a_{1}} \notag \\
&\phantom{\quad + \sum_{q<r}^{k} \Big(} + 2 \delta_{N 2} \delta_{a_r a_q} \psi_{a_k \dots a_{r+1} a_{r-1} \dots a_{q+1} a_{q-1}\dots a_1}\Big)
\end{align}
with
\begin{align}
\label{eq:Fqrn}
&F_{a_k \dots a_1}^{qr,n} \notag \\
&= 2 \Braket{\Phi_{\mathbf{s}}(\mathbf{z}) ( \Ja{k} \dots \Ja{r+1} J^{a_q}_{-n-2} \Ja{r-1} \dots \Ja{q+1} \notag \\
&\quad\quad J^{a_r}_{n} \Ja{q-1} \dots \Ja{1})(0)} \notag \\
&\quad - 2 \delta_{a_r a_q} \Braket{\Phi_{\mathbf{s}}(\mathbf{z}) (\Ja{k} \dots \Ja{r+1} J^{c}_{-n-2} \Ja{r-1} \dots \Ja{q+1} \notag \\
&\quad\quad J^{c}_{n} \Ja{q-1} \dots \Ja{1})(0)} \notag \\
&\quad + 2 N \tilde{\delta}_{n+2} i \varepsilon_{a_r a_q c} \Braket{\Phi_{\mathbf{s}}(\mathbf{z}) (\Ja{k} \dots \Ja{r+1} \Ja{r-1} \dots \Ja{q+1} \notag\\
&\quad\quad J^{c}_{n} \Ja{q-1} \dots \Ja{1}(0))(0)} .
\end{align}
In the last term, $\tilde{\delta}_{n+2} = 1$ if $(n+2)$ mod $N = 0$
and $\tilde{\delta}_{n+2} = 0$ otherwise.
Let us now argue that this expression contains $k$ current operators
of order $-1$.
There are two terms that are not yet explicitly written in the desired
form, namely the term proportional to $\delta_{N 2}$ and the term
abbreviated by $F_{a_k \dots a_1}^{qr,n}$.
The operators $J^{a_q}_{-n-2}$ (and $J^{c}_{-n-2}$, respectively), can
be written as a linear combination of $n+2$ current operators of order
$-1$ by repeated application of Eq. \eqref{eq:elim-higher-order-J}.
On the other hand, the operators $J^{a_r}_{n}$ (and $J^{c}_{n}$,
respectively), can be commuted to the right, using the current algebra
(cf. Eq. \eqref{eq:Kac-Moody-current-algebra})
\begin{align}
\label{eq:current-algebra-move-right}
\Comm{J^a_{n}}{J^b_{-1}} &= i \varepsilon_{abc} J^{c}_{n-1} + \frac{1}{2} \delta_{ab} \delta_{n-1,0}.
\end{align}
The resulting terms either have $n$ current operators less or vanish
because $J^a_m \Ket{0} = 0$ for $m \ge 0$.
The total number of current operators in the first two terms of $F_{a_k \dots a_1}^{qr,n}$
is therefore
\begin{align}
\label{eq:total-number-of-current-operators}
k - 2 + \underbrace{n + 2}_{J^{a_q}_{-n-2}} \underbrace{-n}_{\vphantom{J^{a_q}_{-n-2}} J^{a_r}_{n}} = k.
\end{align}
If $k \ge N$, there can be a contribution from the third term in $F_{a_k \dots a_1}^{qr,n}$,
namely if $n+2 = m N$ for $m \in \{1, 2, \dots\}$.
This term can be written in terms of $k - m N$ current operators.
Note, however, that the space of states with $k$ current operators
contains the space of states with $k - m N$ current operators. The reason
for this is that a current operator $J^a_{-1}$ corresponds to the
application of a Fourier transformed spin operator $u^a_{-1}$,
for which $u^a_{-1} = u^a_{-1 - m N}$.
This argument also applies to the term that is proportional to $\delta_{N 2}$.
Thus, the Hamiltonian is block diagonal in the states
with a fixed number of current operators.
Note that this observation does not follow from translational invariance
only. Translational invariance implies that the Hamiltonian does
not mix states with different lattice momenta. Since each
current operator in $\psi_{a_k \dots a_1}$ contributes a unit of $2 \pi / N$
to the momentum, it follows from translational
invariance that $\mathcal{H}\psi_{a_k \dots a_1}$ is a linear combination of states with $k
\text{ mod } N$ current operators.
The above considerations moreover show that it is possible to write $\mathcal{H}\psi_{a_k \dots a_1}$
as a linear combination of states with strictly $k$ modes $J^a_{-1}$. In
particular, it is not necessary, to include terms with a higher number
of current operators.
This allows us to block-diagonalize the Hamiltonian starting
with the smaller blocks, i.e. those with a small number of current operators.
\subsection{Eigenstates from Current Operators}
\label{SEC:Eigenstates-Analytically}
We have solved the eigenvalue equation of
Eq. \eqref{eq:derivation-H-Jpsi} for up to eight current operator
modes analytically. Since the Hamiltonian is $SU(2)$ invariant, we
have decomposed the eigenstates into different spin sectors.
The momenta of the states are directly related to the number of
current operators, with each current operator changing the
momentum by $2 \pi / N$.
We summarize our results for up to four current operators in Table \ref{table:computed-eigenstates}.
\begin{table*}[htb]
\caption{\label{table:computed-eigenstates}
Eigenstates of $\mathcal{H}$ in terms of states obtained by insertion of $k$ current operator modes for $k \le 4$. The momentum of the ground state is $p_0 = \pi$ if $N/2$ is odd
and $p_0=0$ if $N/2$ is even.}
\centering
\begin{tabular}{rllllll}
$k$ & State & Null for & Energy & Spin & Momentum & Number of states \\
\hline
$0$ & $\varphi^{(0)} = \psi_0$ & & $0$ & $0$& $p_0$ & $1$ \\
$1$ & $\varphi^{(1)}_a = \psi_a$ & & $N - 1$ & $1$ & $p_0 - 2 \pi/N$ & $3$ \\
$2$ & $\varphi^{(2)} = \sum_c \psi_{c c} - 3 \delta_{N 2} \psi_0$& $N \le 2$ & $2(N-3)$ & $0$ & $p_0 - 4 \pi/N$ & $1$ \\
$2$ & $\varphi^{(3)}_a = \sum_{c d} \varepsilon_{acd} \psi_{c d}$ & $N \le 2$ & $2(N-3)$ & $1$& $p_0 - 4 \pi/N$ & $3$ \\
$3$ & $\varphi^{(4)} = \sum_{cde} \varepsilon_{cde} \psi_{cde}$ & $N \le 4$ & $3(N-5)$ & $0$ & $p_0 - 6 \pi/N$ & $1$ \\
$3$ & $\varphi^{(5)}_a = \sum_c (2 \psi_{acc} - 3 \psi_{cac} + \psi_{cca}) - 4 \delta_{N2} \psi_a$ & $N \le 4$ & $3(N-5)$ & $1$ & $p_0 - 6 \pi/N$ & $3$ \\
$3$ & $\varphi^{(6)}_a = \sum_c (\psi_{cca} - \psi_{acc}) + 2 \delta_{N2} \psi_a$ & $N \le 2$ & $3(N-3)$ & $1$ & $p_0 - 6 \pi/N$ & $3$ \\
$4$ & $\varphi^{(7)} = \sum_{c d} (5 \psi_{c d c d} - 3 \psi_{c d d c} - 2 \psi_{c c d d}) + 16 \delta_{N 4} \psi_0 + 12 \delta_{N2} \psi_0$ & $N \le 6$ & $4 (N - 7)$ & $0$ & $p_0 - 8 \pi / N$ & $1$\\
$4$ & $\varphi^{(8)}_{a} = \sum_{c d e} (4 \varepsilon_{a c d} \psi_{c d e e} - 3 \varepsilon_{a c d} \psi_{c e d e})$ & $N \le 6$ & $4 (N-7)$ & 1 & $p_0 -8 \pi / N$ & $3$ \\
$4$ & $\varphi^{(9)} = \sum_{c d} (\psi_{c d d c} - \psi_{c c d d}) - 12 \delta_{N 4} \psi_0 + 6 \delta_{N2} \psi_0$ & $N \le 4$ & $4 N - 18$ & $0$ & $p_0 - 8 \pi / N$ & $1$ \\
$4$ & $\varphi^{(10)}_{a} = \sum_{c d e} (\varepsilon_{a c d} \psi_{c d e e} + 3 \varepsilon_{a c d} \psi_{c e d e})$ & $N \le 4$ & $4 N - 18$ & $1$ & $p_0 - 8 \pi / N$ & $3$ \\
$4$ & $\varphi^{(11)}_{ab} = \sum_{c} \left(\frac{1}{2}(\psi_{a b c c } + \psi_{b a c c }) - \frac{1}{3} \delta_{ab} \sum_d \psi_{ddcc} \right)$ & $N \le 2$ & $4 N - 10$ & $2$ & $p_0 - 8 \pi / N$ & $5$
\end{tabular}
\end{table*}
At level one we find a triplet (spin one), at level two a
singlet and a triplet, at level three we find a singlet and two
triplets with different energies. A spin two state appears at
level 4 as the symmetric traceless part of a state with two non-contracted
$SU(2)$-indices.
Note that the number of eigenstates is smaller than the number of
possible combinations we can build with $k$ current operators. The
reason is that some CFT states $J^{a_k}_{-1} \dots J^{a_1}_{-1}
\Ket{0}$ are null, such that the norm of the corresponding spin state
vanishes. At level three, for example, there is the null state
$\sum_b (3 \psi_0^{bab} + 3 \psi_0^{bba} - 2 \psi_0^{abb})$.
The number of states that we find with a certain number of
current operators is in agreement with the characters of the $SU(2)_1$
algebra\cite{Nielsen2011}: At level $0$, $1$, $2$, $3$, and 4 we find
$1$, $3$, $4$, $7$, and $13$ states, respectively.
The size of the matrices that need to be diagonalized at a given
level in current operators thus corresponds to the characters of
$SU(2)_1$.
When considering the action of the Hamiltonian on states with $k$
current operators, there are two types of terms that depend on the
number of spins $N$, cf Eq. \eqref{eq:derivation-H-Jpsi}. The first
one, $k (N-1) \psi_{a_k \dots a_1}$, is already diagonal. All other terms only appear
if $k \ge N$. These terms are strictly upper triangular in the sense that
they can be written in terms of $k - m N$ current operators with $m >
0$. This upper triangular structure is preserved by a diagonalization
of all other terms. Therefore, only the diagonal term $k (N-1) \psi_{a_k \dots a_1}$
contributes to the $N$-dependence of the energies.
We thus arrive at an $N$-independent representation of the energies by
subtracting the contribution $k (N-1)$ from the energies,
\begin{align}
\label{eq:shifted-and-rescaled-energies} \tilde{E} &= \begin{cases}
\frac{E - (N-1) k}{k}, &k > 0,\\ 0, &k = 0.
\end{cases}
\end{align}
(We have also rescaled the energies by $1/k$ for
convenience.)
The shifted and rescaled energies $\tilde{E}$ as well as the spin
content of the corresponding eigenspaces are plotted for up to
eight current operators in Fig. \ref{fig:analytical_spectrum}.
\begin{figure*}[htb]
\centering
\includegraphics[width=.9\linewidth]{analytical_spectrum.pdf}
\caption{\label{fig:analytical_spectrum} (Color online) Analytically
calculated energy levels and their spin content as a function of the
number of current operators $k$. $\tilde{E}$ is defined as $\tilde{E}
= (E - (N-1) k)/k$ for $k > 0$ and $\tilde{E} = E = 0$ for $k=0$. The
three rows in the boxes next to the energy level correspond to the
calculated values of (1) the energy $\tilde{E}$ (2) the spin content
(3) a value for the number of spins $N'$ so that the corresponding
state is null for all systems with $N \le N'$ spins. The inequalities
indicate for which $N$ the condition $E \ge 0$ is satisfied. For a
given number of spins $N$, all eigenvalues that lie in bands where the
inequality is not satisfied correspond to null states. By applying
additional current operators to these null states, we could
identify further null states that do not violate the energy condition.}
\end{figure*}
Note that for a given $N$, some of these energies do not
occur because the corresponding states have zero norm. These null
states appear dependent on $N$ and in addition to null states
identified at the CFT level. The occurrence of additional null states
reflects the fact that the CFT Hilbert space is infinite, while the
spin system's Hilbert space is finite.
A necessary condition for a particular state to be non-null is given
by $E \ge 0$. The region of energies, where $E \ge 0$ is satisfied is
indicated by the bands in
Fig. \ref{fig:analytical_spectrum}. Depending on $N$, we find certain
states that violate this condition and are therefore null. These
states then lead to further null states at higher levels through the
application of additional current operators. We find that if a state
corresponding to a certain energy is null for a number of spins $N'$,
it is also null for $N$ spins with $N \le N'$. For each energy level,
the highest $N'$ that we found exploiting the energy condition $E \ge
0$ is given in Fig.~\ref{fig:analytical_spectrum} below the spin
content.
Let us give an example for the notation used in
Fig. \ref{fig:analytical_spectrum}. At $k=7$ and $\tilde{E}=-6$ we
find two spin $1$ states and one spin $2$ state. One of the spin $1$
states is null for all $N \le 6$, the other spin $1$ state and the
spin $2$ state are null for all $N \le 8$. This multiplet is shown in
Fig. \ref{fig:analytical_spectrum} as
\begin{equation}
\label{eq:example-of-multiplet}
\boxed{
\begin{matrix}-6 & & & &\\1 &\oplus &1 &\oplus &2 \\ 6 && 8 && 8
\end{matrix}
}.
\end{equation}
Note that the violation of the energy condition $E \ge 0$ is
sufficient for a state to be null. A complete separation of null
states requires the computation of inner products between our ansatz
states at the level of the spin system. As shown in
Ref. \onlinecite{Nielsen2011}, the spin correlation functions in $\psi_0$
can be computed by solving linear algebraic equations. Using
these correlation functions, one could compute the norms of the states
$\psi_{a_k \dots a_1}$ numerically, even for large system sizes.
\subsection{Construction of Yangian Highest Weight States and Comparison to Spinon Basis}
The spectrum of the Haldane-Shastry model has previously been
constructed by exploiting a hidden symmetry of the Hamiltonian, which
is generated by the rapidity operator $\Lambda^a$\cite{Haldane1992},
\begin{align}
\label{eq:definition-of-Yangian}
\Lambda^a &= \frac{i}{2} \sum_{i \neq j} w_{ij} \varepsilon_{abc} t^b_i t^c_j = \frac{i}{2} \sum_{i \neq j} w_{ij} (\vec{t}_i \times \vec{t}_j)^a.
\end{align}
Using $\sum_{j(\neq i)} w_{ij} = 0$, it follows that
\begin{align}
\label{eq:rapidity-in-terms-of-c}
\Lambda^a = \frac{3}{4} \sum_{i=1}^N \Co{a}{i}.
\end{align}
The rapidity $\Lambda^a$ and the total spin $T^a$ form a basis of the
Yangian algebra. They both commute with the Haldane-Shastry
Hamiltonian, but the rapidity $\Lambda^a$ does not commute with the
total spin operator $T^a T^a$. This is the reason for the degenerate
energy levels formed by multiplets with different total
spin\cite{Haldane1992}.
The key to the construction of eigenstates in this approach is the
notion of a Yangian highest weight state $h$, which is annihilated by
$\Lambda^{+} = \Lambda^{x} + i \Lambda^{y}$ and $T^{+} = T^{x} + i T^{y}$.
A multiplet of states with the same energy is then given by application
of powers of $\Lambda^{-} = \Lambda^{x} - i \Lambda^{y}$ to $h$.
In order to relate our method to this approach, we have computed
the action of $\Lambda^a$ on $\psi_{a_k \dots a_1}$ using the decoupling equation derived
in section \ref{SEC:decoupling-equation-for-Jpsi} of the Appendix.
We find
\begin{align}
\label{eq:action-of-Yangian}
2 \Lambda^a \psi_{a_k \dots a_1} &= \sum_{q=1}^k (N-1) i \varepsilon_{a a_q c} \psi_{a_k \dots a_{q+1} c a_{q-1} \dots a_1} \notag \\
&\quad + \sum_{q=1}^k i \varepsilon_{a_q a c} \psi_{c a_{k} \dots a_{q+1} a_{q-1} \dots a_{1}} \notag \\
&\quad + \sum_{q=2}^k \sum_{n=0}^{q-1} G^{q, n}_{a_k \dots a_q},
\end{align}
with
\begin{align}
\label{eq:G-symbol-defintion}
&G^{q, n}_{a_k \dots a_q} = \notag \\
&2 \langle \Phi_{\mathbf{s}}(\mathbf{z}) (J^{a_q}_{-n-1} \Ja{k} \dots \Ja{q+1} J^a_n \Ja{q-1} \dots \Ja{1})(0) \rangle \notag \\
&- 2 \delta_{a_q a} \langle \Phi_{\mathbf{s}}(\mathbf{z}) (J^c_{-n-1} \Ja{k} \dots \Ja{q+1} J^{c}_n \Ja{q-1} \dots \Ja{1})(0) \rangle\notag\\
&+2 N \tilde{\delta}_{n+1} i \varepsilon_{a a_q c} \langle \Phi_{\mathbf{s}}(\mathbf{z}) (\Ja{k} \dots \Ja{q+1} \notag\\
&\quad\quad J^{c}_n \Ja{q-1} \dots \Ja{1})(0) \rangle.
\end{align}
Furthermore, we have (cf. Eq. \eqref{eq:T-Jpsi-in-epsilon} in Appendix \ref{SEC:decoupling-equation-for-Jpsi})
\begin{align}
\label{eq:action-of-T-on-Jstate}
T^a \psi_{a_k \dots a_1} &= i \sum_{q=1}^N \varepsilon_{a a_qc} \psi_{a_k \dots a_{q+1} c a_{q-1}\dots q_1}.
\end{align}
This shows that the Yangian, like the Hamiltonian, leaves the
subspaces of states with a fixed number of current operators
invariant. It is thus possible to write the highest weight states of
the Yangian algebra in terms of the states $\psi_{a_k \dots a_1}$. We have computed
the highest weight states for up to four current operators and
expanded the result in the eigenstates listed in Table
\ref{table:computed-eigenstates}. These states thus correspond to the
eigenstates constructed by Haldane in Ref.~\onlinecite{Haldane1991},
which have a polynomial form in the particle basis and were identified
as the highest weight states of the Yangian algebra in
Ref.~\onlinecite{Haldane1992}. We summarize our results in Table
\ref{table:highest-weight-states}.
\begin{table}[htb]
\caption{\label{table:highest-weight-states}
Highest weight states of the Yangian algebra in terms of the eigenstates of the Hamiltonian
given in Table \ref{table:computed-eigenstates}.}
\centering
\begin{tabular}{rlllll}
$k$ & State & Energy & Spin \\
\hline
$0$ & $\varphi^{0}$ & $0$ & $0$ \\
$1$ & $\varphi^{(1)}_x + i \varphi^{(1)}_y$ & $N-1$ & $1$ \\
$2$ & $\varphi^{(3)}_x + i \varphi^{(3)}_y$ & $2(N-3)$ & $1$ \\
$3$ & $\varphi^{(5)}_x + i \varphi^{(5)}_y $ & $3(N-5)$ & $1$ \\
$3$ & $\varphi^{(6)}_x + i \varphi^{(6)}_y$ & $3(N-3)$ & $1$ \\
$4$ & $\varphi^{(8)}_x + i \varphi^{(8)}_y$ & $4 (N-7)$ & $1$ \\
$4$ & $\varphi^{(10)}_x + i \varphi^{(10)}_y$ & $4N - 18$ & $1$ \\
$4$ & $\varphi^{(11)}_{zz} + 2 \varphi^{(11)}_{xx} + 2i \varphi^{(11)}_{xy}$ & $4 N - 10$ & $2$
\end{tabular}
\end{table}
Note that our ansatz is manifestly $SU(2)$ invariant, whereas
the construction in terms of highest weight states is not.
The states $\psi_{a_k \dots a_1}$ have a simple form in the sense that
they are created from the ground state by successive application
of Fourier transformed spin operators (cf. Eq. \eqref{eq:def-I-spin-ops}).
We have compared the highest weight states of the spin system to the highest
weight states of the Yangian algebra in the CFT, which is spanned
by\cite{Haldane1992} $J^a_{0}$ and
\begin{align}
\label{eq:CFT-lambda}
Q^a = \frac{i}{2} \sum_{m=1}^{\infty} \varepsilon_{abc} J^b_{-m} J^c_{m}.
\end{align}
We find that the highest weight states that we computed in the spin system
are precisely the highest weight states of $J^a_0$ and $Q^a$, when
seen as states of the CFT under the correspondence of
Table \ref{table:structure-of-excited-states}.
Finally, let us relate our ansatz to the spinon basis of the
CFT Hilbert space \cite{Bouwknegt1994}. In this approach,
states are not constructed by application of modes of the affine current $J^a_{-n}$
to the CFT vacuum, but by application of modes $\phi_{s, -m}$ of the primary field $\phi_s(z)$.
These modes are defined with respect to the Laurent expansion \cite{Bouwknegt1994}
\begin{align}
\label{eq:Laurent-expansion-of-spinon-modes}
\phi_s(z) &= \sum_{m} z^{m + \frac{q}{2}} \phi_{s, -m-\frac{1}{4}-\frac{q}{2}}.
\end{align}
Here $q \in \{0, 1\}$ corresponds to the sector of the CFT Hilbert space on which $\phi_s(z)$
is acting: $q=0$ for a state built from an even number of spinon modes and $q=1$ for a
state with an odd number of spinon modes. The authors
of Ref.~\onlinecite{Bouwknegt1994} derived generalized commutation relations
for the modes of $J^a(z)$ and $\phi_s(z)$, which can be used
to rewrite states built from current operator modes in terms of
spinon modes $\phi_{s, -m}$. At level $k=1$, for example,
the state $\Braket{\Vi{1} \dots \Vi{N} J^a_{-1}(0)}$ corresponds to
\begin{align}
\label{eq:J-lowest-level-to-spinon-modes}
\sum_{s s'} (t^a)_{s s'} \Braket{\Vi{1} \dots \Vi{N} \left(\phi_{-s, -\frac{3}{4}} \phi_{s', -\frac{1}{4}}\right)(0)}.
\end{align}
A general state with $k$ current operators of order $-1$ will be a linear
combination of terms with spinon modes $\phi_{\alpha_l, -m_l} \dots \phi_{\alpha_1, -m_1}$ with $k = \sum_{i=1}^{l} m_i$.
\subsection{Complete Spectrum}
\label{SEC:Numerics}
In the previous subsections, we have analytically diagonalized the
Hamiltonian in the states $\psi_{a_k \dots a_1}$ for $k \le 8$ and $N$
even. For a given $k$, the size of the matrices that need to be
diagonalized and therefore the complexity of the analytical
calculation does not depend on the number of spins $N$. However, it
becomes increasingly difficult for larger $k$. In order to test if our
method yields all excited states or just a subset thereof, we have
thus constructed the states $\psi_{a_k \dots a_1}$ numerically for small $N$ and
performed a numerical diagonalization of the Hamiltonian in that
basis. Our numerical calculations confirm that the complete
spectrum is indeed obtained from our ansatz states.
In an analogous way to $\psi_0$, we studied the tower of states
obtained from $\psi^{s_0, s_\infty}_0$ by insertion of current
operators (cf. Section \ref{SEC:definition-psi-zero-inf}). We find
numerically that the complete Hilbert space is generated, starting
from $\psi^{s_0s_\infty}_0$ with $s_0, s_\infty \in \{-1, 1\}$ and
successively inserting current operator modes. As for $\psi_0$, the
Hamiltonian is block diagonal in the states with a fixed number of
current operators.
Furthermore, we constructed the spectrum numerically for the case of
an odd number of spins (cf. Section \ref{SEC:odd-number-of-spins}).
We find that the Hamiltonian is block diagonal in the states
$\psi_{a_k \dots a_1}^{s_0}$ and $\psi_{a_k \dots a_1}^{s_\infty}$ and that the complete Hilbert
space can be constructed from states of this form.
The numerically
obtained spectra are shown for $N=7$ and $N=8$ spins in
Fig. \ref{fig:numerical-spectra}.
\begin{figure}[htb]
\centering
\includegraphics[width=.9\linewidth]{spectrum_N_7_8.pdf}
\caption{\label{fig:numerical-spectra}
Numerically calculated spectra for the odd-$N$ spin chain ($N=7$, left
panels) and the even-$N$ spin chain ($N=8$, right panels). The upper
panels show data obtained by applying current operators to the states
$\psi^{s_\infty}_0$ and $\psi_0$, respectively. In the lower panels,
the current operators were applied to the states $\psi^{s_0}_0$ and
$\psi^{s_0 s_\infty}_0$, respectively. The horizontal axis shows the
number of current operators $k$ that are present in the corresponding
states. The numbers above the levels indicate their degeneracy.
Notice that the spectrum is the same in the upper and lower plots, but
a given state does not necessarily appear at the same number of
current operators.}
\end{figure}
We observe a tendency that states with a higher
number of current operators have a larger energy.
This means that by inserting few current operators, we get access to
low-lying excited states. Note, however, that this
relation does not hold in a strict sense; it may happen that, by
inserting additional current operators, we obtain states with a lower
energy.
We observe that the spectra shown in the upper and in the lower panels
of Fig. \ref{fig:numerical-spectra} are the same. This
means that for both $N$ even and $N$ odd, the complete spectrum
is obtained for either of the two classes of states.
The top right panel of Fig. \ref{fig:numerical-spectra} corresponds to
our analytical results shown in Fig. \ref{fig:analytical_spectrum}.
Note that not all states of Fig. \ref{fig:analytical_spectrum} appear
in the numerically calculated spectrum, because certain states of
Fig. \ref{fig:analytical_spectrum} are null for $N=8$. This is the
case for the all states with $E < 0$, which appear in the bands
below $\tilde{E} = -7$ in Fig. \ref{fig:analytical_spectrum}. The
remaining null states have the values
\begin{center}
\begin{tabular}{llll}
$k$ & $\tilde{E}$ & $E$ & Spin\\
\hline
$6$ & $-7$ & $0$ & $0 \oplus 1$\\
$7$ & $-\frac{48}{7}$ & $1$ & $0 \oplus 1$\\
$7$ & $-6$ & $7$ & $1 \oplus 2$\\
$8$ & $-6$ & $8$ & $0 \oplus 1^2 \oplus 2$\\
\end{tabular}
\end{center}
and can be obtained from null states violating the energy
condition by the insertion of additional current operators.
We have numerically computed the maximal number of current operator
insertions of order $-1$ needed to generate the complete spectrum as a
function of $N$. The results are shown for the tower of states built
on $\psi_0$ ($N$ even) in
Fig. \ref{fig:kmax-vs-N-for-N-even-and-psi0}. In this case, our
numerical data suggests that $(N/2)^2$ current operators are
sufficient to obtain the complete spectrum.
\begin{figure}[htb]
\centering
\includegraphics[width=.9\linewidth]{kmax_vs_N_for_N_even_and_psi0.pdf}
\caption{\label{fig:kmax-vs-N-for-N-even-and-psi0}
Maximal number of current operator insertions $k_{\mathrm{max}}$
of order $-1$ needed to obtain the spectrum of the Hamiltonian
dependent on the number of spins $N$. The data is consistent with
$k_{max} = (N/2)^2$.}
\end{figure}
Finally, let us comment on how the first
excited states of the Haldane-Shastry model for $N$ even
are related to the states $\psi_{a_k \dots a_1}$.
Except for $N=2$, these cannot be given by the states at
level one, $\psi_{a_1}$, which have $E=N-1$. The reason is that the states
$\psi^{s_0, s_\infty}_0$ are eigenstates with a lower energy of $E=N/2$. This follows
from
\begin{align}
\label{eq:decoupling-psi-0inf}
\mathcal{C}^a_i \psi^{s_0, s_\infty}_0 &= (K^a_b)_i (t^b_{\infty} - t^b_{0}) \psi^{s_0, s_\infty}_0
\end{align}
and
\begin{align}
\label{eq:cdagger-on-circle}
(\mathcal{C}^a_i)^{\dagger} = \mathcal{C}^a_i - \frac{4}{3} \sum_{j (\neq i)} w_{ij} t^a_j.
\end{align}
Note that the latter equality is only valid on the circle with $z_j = e^{2 \pi i j / N}$. Our
numerical calculations indicate that the states $\psi^{s_0,
s_{\infty}}_0$ are indeed the first excited states. Furthermore, we find a multiplet
with spin content $0 \oplus 1$ and energy $N/2$ at $k=N/2$ among the states constructed
from $\psi_{a_k \dots a_1}$, both in the numerical and in the analytical calculations. These multiplets
appear at $\tilde{E} = -2 k + 2$ in the analytically computed
spectrum shown in Fig. \ref{fig:analytical_spectrum}.
Noting that the states $\psi^{s_0, s_\infty}_0$ can be decomposed into a singlet
and a triplet, this suggests that the first excited states
are given by linear combinations of states $\psi_{a_k \dots a_1}$ with $k=N/2$
current operator modes.
\section{Conclusion}
\label{SEC:Conclusion}
We studied a model of a quantum spin chain that is constructed
from the $SU(2)_1$ WZW model and is equivalent to the Haldane-Shastry model.
We have shown that it is possible to construct excited states
of the spin system from CFT current operators in a
way, which directly reflects the structure of
the CFT spectrum.
In the approach pursued in this work,
correlation functions of operators in the CFT
are interpreted as wave functions of spin states.
Based on earlier work,
where the correlation function of \emph{primary} operators
were interpreted as the ground state wave function, we provided a method
of constructing excited states by inserting \emph{descendant} fields.
In the case of an even number of spins $N$, we have diagonalized the
Hamiltonian analytically for $k \le 8$ inserted current operators.
Depending on $N$, we identified certain states that are null but
correspond to non-null CFT states. These additional null states occur
due to the finite size of the spin system's Hilbert space. Our method
of detecting them is based on a sufficient criterion for a state to be
null. In order to test if a given state is not null, one could use
the known algebraic equations for spin correlation
functions\cite{Nielsen2011} in $\psi_0$ to compute the needed inner
products numerically, even for large system sizes.
We have given numerical evidence that this method yields the complete
spectrum for a given number of spins $N$, independent of which primary
state we build the tower states from. Furthermore, we have shown
numerically that a similar construction can be made for an odd number
of spins.
Our manifestly $SU(2)$ invariant ansatz is
compatible with the construction of eigenstates
of the Haldane-Shastry model as multiplets of the Yangian algebra
in the sense that the Yangian operator does not change the number
of current operators. This allowed us to explicitly relate
the excited states constructed from current operators to
the highest weight states of the Yangian operator. Furthermore, we
have argued that our ansatz wave functions with $k$ current operator
modes of order $-1$ can be rewritten in terms of a wave function
with spinon modes whose mode numbers sum up to $-k$.
In the case of the $SU(2)_1$ WZW model, which we studied here, the
resulting spin system is equivalent to the Haldane-Shastry model.
Thus our method provides an alternative way of constructing the
excited states of the Haldane-Shastry model, which emphasizes its
close relation to the underlying CFT. We expect that this method could
be generalized to the $SU(n)_1$ WZW model, which is related to the
$SU(n)$ Haldane-Shastry spin chain \cite{Schoutens1994}. Another
generalization of our construction could be possible within the
Laughlin lattice models with filling fraction $1/q$
\cite{Tu2014b}. Even though these systems do not have a Yangian
symmetry for $q > 2$, part of the spectrum is described by integer
eigenvalues. We have carried out exemplary numerical calculations for
$q=3$ which indicate that at least some of the excited states can be
obtained by insertion of current operators.
It is crucial for the construction used above that the system
is one-dimensional with periodic boundary conditions. This can be
seen, for example, noting that the ansatz $\psi_{a_k \dots a_1}$ is a decomposition
into momentum space eigenstates. However, the states obtained by
insertion of current operators might still describe low energy
eigenstates of an interesting, two-dimensional system with a different
Hamiltonian. We plan to investigate the properties of these states in
two dimensions in future work.
\begin{acknowledgments}
We are grateful to J. Ignacio Cirac for many fruitful discussions.
This work has been supported by the EU project SIQS, FIS2012-33642,
QUITEMAD (CAM), and the Severo Ochoa Program. Figures were created
using \emph{matplotlib} \cite{Hunter2007} and \emph{MayaVi}
\cite{Ramachandran2011}.
\end{acknowledgments}
|
\section{Introduction}
\IEEEPARstart{A}{therosclerosis} is a systemic/chronic disease characterized by the accumulation of inflammatory cells and lipids in the inner lining of the arteries. It is a leading cardiovascular disease causing deaths worldwide~\cite{Farouc2006,Christopher1997}. Four million Americans have survived a stroke and lead disabled lives. More than $1$ out of $3$ ($83$ million) U.S. adults currently live with one or more types of cardiovascular diseases~\cite{CDC2011}. In $2010$, the total amount spent on cardiovascular diseases in the United States was estimated to be $\$444$ billion~\cite{CDC2011}. The number of deaths due to coronary artery disease in India were projected to increase from $1.591$ million in the year $2000$ to $2.034$ million by the year $2010$~\cite{Bhanot1999}.
Atherosclerosis has a long asymptomatic phase with a sub-clinical incubation period ranging from $30$ to $50$ years. The physicians would like to assess the risk of the patients having a severe clinical event such as stroke or heart attack and predict the same if possible. Some of the major known risk factors that eventually lead to the development of atherosclerosis are as follows: (i) family history of premature coronary heart disease or stoke in a first degree relative under the age of $60$, (ii) tobacco abuse, (iii) type II diabetes, (iv) high blood pressure, (v) left ventricular hypertrophy, (vi) high triglycerides, (vii) high low-density lipoprotein (LDL) cholesterol (viii) low high-density lipoprotein (HDL) cholesterol, (ix) high total cholesterol. Large number of factors influence the onset of atherosclerosis making it a difficult task for the physicians to diagnose in its early stages. Though main risk factors are identified, development of automated effective risk prediction models using data mining techniques becomes essential for better health monitoring and prevention of deaths due to cardiovascular diseases~\cite{Kukar1999,luckas2002}. Establishing a diagnostic procedure for early detection of atherosclerosis disease is very important as any delay would increase the risk of serious complications or even disability. Determining the conditions (risk factors) predisposing the development of atherosclerosis can lead to tests for identifying the disease in its early stages.
Though the clinical risk factors of CHD are identified there is a need for additional understanding on the disease progression for effective management~\cite{Natasha2011}. Researchers have been focusing on techniques for quantifying atherosclerosis plaque morphology, composition, mechanical forces etc., hoping for better patient screening procedures. Imaging of atherosclerotic plaques helps in both diagnosis and monitoring of the progression for future management~\cite{Davies2004,Tardif2011}. There are two modes of imaging atherosclerosis (i) invasive and (ii) non-invasive. Among the invasive methods X-ray angiography is the gold standard imaging technique even though it has certain limitations in providing information on plaque composition~\cite{Liang2011}. The invasive procedures such as intravascular ultrasound (IVUS)~\cite{Garcia-Garcia2010} and angioscopy~\cite{Ueda2003} help in understanding the plaque size and to a limited extent its composition. The intravascular thermography~\cite{Belardi2005} aids in monitoring the changes in plaque composition and metabolism. The non-invasive procedures such as B-mode ultrasound~\cite{Magnussen2011}, computerized tomography (CT)~\cite{Wintermark2008} and magnetic resonance imaging (MRI)~\cite{Corti19042011}, can provide information on plaque composition on vascular beds but they fail to throw much light on the metabolic activity of the plaque inflammatory cells. Though nuclear imaging techniques such as single photon emission computed tomography (SPECT) and positron emission tomography (PET)~\cite{Laufer2009} have the potential for 2D and 3D surface reconstruction of thrombus using radio labels to provide information on molecular, cellular and metabolic activity of plaques~\cite{Choudhury2009}, they lack the required resolution, sensitivity for detection and functional assessment in medium to small size arteries found in coronary circulation.
Keeping in view the limitations of the imaging techniques there is still a greater need for developing automated methods for predicting the risk factors of atherosclerosis disease in individuals which would be of great help in reducing the disease related deaths. Machine learning approaches have been employed in a variety of real world problems to extract knowledge from data for predictive tasks. The presence of large number of attributes in medical databases affects the decision making process as some of the factors may be redundant or irrelevant. Also, the presence of missing values and highly skewed value distributions in the attributes of medical datasets require development of new preprocessing strategies. Feature selection methods are aimed at identifying feature subsets to construct models that can best describe the dataset. The other advantages in using feature selection methods include: identifying and removing of redundant/irrelevant features~\cite{liu2005}, reducing the dimensionality of the dataset, and improving the predictive capability of the classifier. The present study attempts to identify risk factors causing atherosclerosis and the possible risk that the individuals are running at.
In view of the above challenges, we present the following novel features of our work:
\begin{itemize}
\item identifying the missing values (MV) in the dataset and imputing them by using a newly developed non-parametric imputation procedure;
\item determining factors that would help in predicting the risk of developing atherosclerosis among the different groups in the community;
\item building a predictive model that has the capability of rendering effective prediction of risk factors in realtime.
\item comparing the performance of our methodology with other state-of-the-art methods used in the identification of risk factors of atherosclerosis;
\item estimating the time complexity and scalability of our new methodology.
\end{itemize}
This paper is organized as follows: A brief survey of the state-of-the-art techniques employed in predicting the risk factors and understanding the biology of atherosclerosis is discussed in Section~\ref{review}, while in Section~\ref{chap8:sec1} we discuss a novel methodology for predicting the clinical risk factors of atherosclerosis. The description of the datasets, experiments and results are presented in Section~\ref{chap8:sec3}. The conclusions are placed in Section~\ref{chap8:sec6} and the discussion is deferred to Section~\ref{chap8:sec7}.
\section{A brief Survey of state-of-the-art techniques}
\label{review}
In~\cite{Luckas2003} a test for predicting atherosclerosis is proposed using genetic algorithm and a fitness function that depends on area under the curve (AUC) of receiver operator characteristics (ROC). In~\cite{Couturier2004} a three step approach based on clustering, supervised classification and frequent itemsets search is adopted to predict if a patient can develop atherosclerosis according to the correlation between his or her habits and the social environment. Support vector machines were employed in~\cite{Hongzong2007} for discriminating patients between coronary and non-coronary heart disease. Supervised classifiers such as Naive Bayes (NB), Multi Layer Perceptron (MLP), Decision Trees (DT) utilize the associations among the attributes for predicting future cardiovascular disorders in the individuals~\cite{Jose2006}. A correlation based feature selection with C4.5 decision tree is applied~\cite{Tsang-Hsiang2006} for risk prediction of cardiovascular disease. Recent developments in imaging methods for diagnosis have given new insights on the molecular and metabolic activity of atherosclerotic plaques~\cite{Crouse2006,Corti19042011,Ibanez2009}. There are many other studies wherein machine learning techniques have also been employed for predicting the risk of CHD due to atherosclerosis using ultrasound and other imaging methods~\cite{Mougiakakou2007,Christodoulou2003}. Community based studies help in understanding the risk factors of atherosclerosis in different social strata~\cite{Chambless2003,Selvin2005}.
In the above studies the missing values (MV) were either deleted~\cite{Tsang-Hsiang2006}, or filled with approximate values~\cite{Luckas2003}. A windowing method was employed~\cite{Jiri2008} for obtaining the aggregates of the attributes for imputation of MVs. These approaches would lead to biased estimates and may either reduce or exaggerate the statistical power. Methods such as logistic regression, maximum likelihood and expectation maximization have been employed for imputation of MV, but they can be applied only on data sets that are either nominal or numeric. There are other imputation methods such as k-nearest neighbor imputation (KNNI)~\cite{BM2003}; k-means clustering imputation (KMI)~\cite{DSSL2004}; weighted k-nearest neighbor imputation (WKNNI)~\cite{TCSBHTBA2001} and fuzzy k-means clustering imputation (FKMI)~\cite{DSSL2004}.
\section{Novel Methodology for Predicting Risk factors of Atherosclerosis}
\label{chap8:sec1}
The mean value imputation proposed by Sree Hari Rao and Naresh Kumar~\cite{Rao2011} can be employed only when the attribute values are normally distributed. In case of highly skewed attribute values the above method may result in biased estimates as mean value is not a true representative of a non-normal distribution. Motivated by the above issues we propose a methodology comprising of a novel nonparametric missing value imputation method that can be applied on (i) data sets consisting of attributes that are of the type categorical (nominal) and/or numeric (integer or real), and (ii) attribute values that belong to highly skewed distribution. The methodology proposed by Freud and Mason~\cite{Freund1999} ignores missing values while generating the decision tree, which renders lower prediction accuracies.
In this paper we propose a new feature subset selection methodology where in, a particle optimization search (PSO) is wrapped around an alternating decision tree (ADT) embedded with new imputation strategy discussed in~Section~\ref{chap9:RNI}, for generation of effective decision rules. This methodology can predict the diagnosis of CHD in real time. In fact the decision rules obtained by employing this novel methodology will be useful to diagnose other individuals based on their risk factors. We designate the present machine learning approach as predictive risk assessment of atherosclerosis ($PRAA$) methodology throughout this work.
We adopt the procedure suggested by Sree Hari Rao and Naresh Kumar~\cite{Rao2011} for building and evaluating the classification model using an ADT.
\subsection{Data Representation}
A medical dataset can be represented as a set \emph{S} having row vectors $(R_{1},R_{2},\hdots,R_{m})$ and column vectors $(C_{1},C_{2},\hdots,C_{n})$. Each record can be represented as an ordered n-tuple of clinical and laboratory attributes $(A_{i1}, A_{i2},\hdots, A_{i(n-1)}, A_{in})$ for each $\emph{i}=1,2,\hdots,m$ where the last attribute $(A_{in})$ for each \emph{i}, represents the physician's diagnosis to which the record $(A_{i1}, A_{i2},\hdots, A_{i(n-1)})$ belongs and without loss of generality we assume that there are no missing elements in this set. Each attribute of an element in \emph{S} that is $A_{ij}$ for $\emph{i}=1,2,\hdots,m$ and $\emph{j}=1,2,\hdots,n-1$ can either be a categorical (nominal) or numeric (real or integer) type. Clearly all the sets considered are finite sets.
\subsection{New Imputation Strategy}
\label{chap9:RNI}
The first step in any imputation algorithm is to compute the proximity measure in the feature space among the clinical records to identify the nearest neighbors from where the values can be imputed. The most popular metric for quantifying the similarity between any two records is the Euclidean distance. Though this metric is simpler to compute, it is sensitive to the scales of the features involved. Further it does not account for correlation among the features. Also, the categorical variables can only be quantified by counting measures which calls for the development of effective strategies for computing the similarity~\cite{Tadashi2009}. Considering these factors we first propose a new indexing measure $I_{C_{l}}(R_{i},R_{k})$ between two typical elements $R_{i}$, $R_{k}$ for $i,k=1,2,\ldots,m,$ $l=1,2,\ldots,n-1$ belonging to the column $C_{l}$ of \emph{S} which can be applied on any type of data, be it categorical (nominal) and/or numeric (real). We consider the following cases:
\begin{description}
\item[Case I:]~ $A_{in}=A_{kn}$ \\
Let \emph{A} denote the collection of all members of \emph{S} that belong to the same decision class to which $R_{i}$ and $R_{k}$ belong. Based on the type of the attribute to which the column $C_{l}$ belongs, the following situations arise:
\begin{itemize}
\item [(i)]
Members of the column $C_{l}$ of S i.e
$(A_{1l},A_{2l},\ldots,A_{ml})^{T}$ are of nominal or categorical or integer type:\\
\noindent We now express \emph{A} as a disjoint union of non-empty subsets of \emph{A}, say $B_{\gamma_{p_{1l}}},B_{\gamma_{p_{2l}}},\ldots,B_{\gamma_{p_{sl}}}$ obtained in such a manner that every element of \emph{A} belongs to one of these subsets and no element of \emph{A} is a member of more than one subset of \emph{A}. That is $\emph{A}=B_{\gamma_{p_{1l}}}\bigcup B_{\gamma_{p_{2l}}}\bigcup, \ldots,\bigcup B_{\gamma_{p_{sl}}}$, in which $\gamma_{p_{1l}}, \gamma_{p_{2l}}, \ldots, \gamma_{p_{sl}}$ denote the cardinalities of the respective subsets $B_{\gamma_{p_{1l}}},B_{\gamma_{p_{2l}}},\ldots,B_{\gamma_{p_{sl}}}$ formed out of the set \emph{A}, with the property that each member of the same subset has the same first co-ordinate and members of no two different subsets have the same first co-ordinate. We define an index $I_{C_{l}}(R_{i},R_{k})$ for each $l=1,2,\ldots,n-1$
\begin{eqnarray*}
I_{C_{l}}(R_{i},R_{k})=\left\{
\begin{array}{ll}
\min\{\frac{\gamma_{p_{il}}}{\gamma},\frac{\gamma_{q_{kl}}}{\gamma}\}, & \hbox{for $i \neq k$;} \\
0, & \hbox{otherwise.}
\end{array}
\right.
\end{eqnarray*}
\noindent where $\gamma_{p_{il}}$ represents the cardinality of the subset $B_{\gamma_{p_{il}}}$, all of whose elements have first co-ordinates $A_{il}$, $\gamma_{q_{kl}}$ represents the cardinality of that subset $B_{\gamma_{q_{kl}}}$, all of whose elements have first co-ordinates $A_{kl}$ and $\gamma=\gamma_{p_{1l}}+ \gamma_{p_{2l}}+ \ldots,+\gamma_{p_{sl}}$ represents the cardinality of the set $A$.
\item [(ii)] Members of the column $C_{l}$ of S i.e $(A_{1l},A_{2l},\ldots,A_{ml})^{T}$ are of real (fractional or non-integer numbers):\\
\noindent We consider the set $P_{l}$ for $l=1,2,\dots,n-1$ which is a collection of all the members of the column $C_{l}$. We then compute the skewness measure $sk(P_{l})=\frac{\frac{1}{m}\sum_{i=1}^{m}(A_{il}-\overline{A_{il}})^3}{(\sqrt \frac{1}{m} \sum_{i=1}^{m}(A_{il}-\overline{A_{il}})^2)^3}$ where $ \overline{A_{il}}$ denote the mean of $A_{il}$ for each $l=1,2,\ldots,n-1$.
Define the sets $\begin{array}{@{\hspace{0mm}}r@{\;}l@{\hspace{0mm}}} M_{l} & = \displaystyle \left\{a\in P_{l}|a \leq A_{il}, \hbox{for $sk(P_{l}) < 0$} \right\}\end{array}$ or, $\begin{array}{@{\hspace{0mm}}r@{\;}l@{\hspace{0mm}}}
M_{l} & = \displaystyle \left\{b\in P_{l}|b > A_{il}, \hbox{for $sk(P_{l}) \geq 0$}\right\}\end{array}$ and similarly $\begin{array}{@{\hspace{0mm}}r@{\;}l@{\hspace{0mm}}}
N_{l} & = \displaystyle \left\{a\in P_{l}|a \leq A_{kl},\hbox{for $sk(P_{l}) < 0$} \right\}\end{array}$ or, $\begin{array}{@{\hspace{0mm}}r@{\;}l@{\hspace{0mm}}}
N_{l} & = \displaystyle \left\{b\in P_{l}|b > A_{kl},\hbox{for $sk(P_{l}) \geq 0$} \right\}\end{array}$. Let $\tau_{l}$ and $\rho_{l}$ be the cardinalities of the sets $M_{l}$ and $N_{l}$ respectively. Construct the index $I_{C_{l}}(R_{i},R_{k})$,
\begin{eqnarray*}
I_{C_{l}}(R_{i},R_{k})=\left\{
\begin{array}{ll}
\min\{\frac{\tau_{l}}{\gamma},\frac{\rho_{l}}{\gamma}\}, & \hbox{for $i \neq k$;} \\
0, & \hbox{otherwise.}
\end{array}
\right.
\end{eqnarray*}
\noindent In the above definition $\gamma$ represents the cardinality of the set $P_{l}$.
\end{itemize}
\item [Case II:]~ $A_{in}\neq A_{kn}$ \\
\label{case2}
Clearly $R_{i}$ and $R_{k}$ belong to two different decision classes. Consider the subsets $P_{i}$ and $Q_{k}$ consisting of members of $\emph{S}$ that share the same decision with $R_{i}$ and $R_{k}$ respectively. Clearly $P_{i}\bigcap Q_{k} = \emptyset$. Based on the type of the attribute of the members, the following situations arise:
\begin{itemize}
\item[(i)] Members of the column $C_{l}$ of S i.e $(A_{1l},A_{2l},\ldots,A_{ml})^{T}$ are of nominal or categorical type:\\
\noindent Following the procedure discussed in Case I item (i) we write \emph{$P_{l}$} and \emph{$Q_{l}$} for each $l=1,2,\ldots,n-1$ as a disjoint union of non-empty subsets of $P_{\beta_{1l}},P_{\beta_{2l}},\ldots,P_{\beta_{rl}}$ and $Q_{\delta_{1l}},Q_{\delta_{2l}},\ldots,Q_{\delta_{sl}}$ respectively in which $\beta_{1l}, \beta_{2l}, \ldots, \beta_{rl}$ and $\delta_{1l}, \delta_{2l}, \ldots, \delta_{sl}$ indicate the cardinalities of the respective subsets. We define the indexing measure between the two records $R_{i}$ and $R_{k}$ as
\begin{eqnarray*}
I_{C_{l}}(R_{i},R_{k})=\left\{
\begin{array}{ll}
\max\{\frac{\beta_{rl}}{\delta_{sl}+\beta_{rl}},\frac{\delta_{sl}}{\delta_{sl}+\beta_{rl}}\}, & \hbox{for $i \neq k$;} \\
0, & \hbox{otherwise.}
\end{array}
\right.
\end{eqnarray*}
where $\beta_{rl}$ represents the cardinality of the subset $P_{\beta_{rl}}$ all of whose elements have first co-ordinates $A_{il}$ in set $P_{l}$ and $\delta_{sl}$ represents the cardinality of that subset $Q_{\delta_{sl}}$, all of whose elements have first co-ordinates $A_{kl}$ in set $Q_{l}$. \label{case2:item1}
\item[(ii)] Members of the column $C_{l}$ of S i.e $(A_{1l},A_{2l},\ldots,A_{ml})^{T}$ are of numeric type:\\
\noindent If the type of the attribute is an integer we follow the procedure discussed in Case II item (i) . For fractional numbers we follow the procedure discussed in Case I item (ii) and we define the set $P_{l}$ as the members of column $C_{l}$ that belong to decision class $A_{in}$ and $Q_{l}$ as the members of column $C_{l}$ that belong to decision class $A_{kn}$. We then compute the skewness measure $sk(P_{l})=\frac{\frac{1}{m}\sum_{i=1}^{m}(A_{il}-\overline{A_{il}})^3}{(\sqrt \frac{1}{m} \sum_{i=1}^{m}(A_{il}-\overline{A_{il}})^2)^3}$ where $ \overline{A_{il}}$ denote the mean of $A_{il}$ for each $l=1,2,\ldots,n-1$ and $sk(Q_{l})=\frac{\frac{1}{m}\sum_{i=1}^{m}(A_{kl}-\overline{A_{kl}})^3}{(\sqrt \frac{1}{m} \sum_{i=1}^{m}(A_{kl}-\overline{A_{kl}})^2)^3}$ where $ \overline{A_{kl}}$ denote the mean of $A_{kl}$ for each $l=1,2,\ldots,n-1$.
We then construct the sets $T_{l}$ and $S_{l}$ as follows:
$\begin{array}{@{\hspace{0mm}}r@{\;}l@{\hspace{0mm}}} T_{l} & = \displaystyle \left\{a\in P_{l}|a \leq A_{il}, \hbox{for $sk(P_{l}) < 0$} \right\}\end{array}$ or, $\begin{array}{@{\hspace{0mm}}r@{\;}l@{\hspace{0mm}}}
T_{l} & = \displaystyle \left\{b\in P_{l}|b > A_{il}, \hbox{for $sk(P_{l}) \geq 0$}\right\}\end{array}$ and similarly $\begin{array}{@{\hspace{0mm}}r@{\;}l@{\hspace{0mm}}}
S_{l} & = \displaystyle \left\{a\in P_{l}|a \leq A_{kl},\hbox{for $sk(P_{l}) < 0$} \right\}\end{array}$ or, $\begin{array}{@{\hspace{0mm}}r@{\;}l@{\hspace{0mm}}}
S_{l} & = \displaystyle \left\{b\in P_{l}|b > A_{kl},\hbox{for $sk(P_{l}) \geq 0$} \right\}\end{array}$. We now define the index $I_{C_{l}}(R_{i},R_{k})$ between the two records $R_{i},R_{k}$ as
\begin{eqnarray*}
I_{C_{l}}(R_{i},R_{k})=\left\{
\begin{array}{ll}
\min\{\frac{\beta_{l}}{\lambda},\frac{\delta_{l}}{\lambda}\}, & \hbox{for $i \neq k$;} \\
0, & \hbox{otherwise.}
\end{array}
\right.
\end{eqnarray*}
\noindent In the above definition $\beta_{l}$ and $\delta_{l}$ represents the cardinalities of the sets $T_{l}$ and $S_{l}$ respectively. The sum of the cardinalities of the sets $P_{l}$ and $Q_{l}$ is represented by $\lambda$
\end{itemize}
\end{description}
The proximity or distance scores between the clinical records in the data set $S$ can be represented as $D=\{\{0,d_{12},\ldots,d_{1m}\};\{d_{21},0,\ldots,d_{2m}\};\ldots;\{d_{m1},d_{m2},\ldots,0\}\}$ where $d_{ik}=\sqrt{\sum_{l=1}^{n-1}I_{C_{l}}^{2}(R_{i},R_{k})}$. For each of the missing value instances in a record $R_{i}$ our imputation procedure first computes the score $\alpha(x_{k})=\frac{(x_{k}-median(x))}{median{|x_{i}-median(x)|}}$ where $\{x_{1},x_{2},\ldots,x_{n} \}$ denote the distances of $R$ from $R_{k}$. We then pick up only those records (nearest neighbors) which satisfy the condition $\alpha(x_{k}) \leq 0$ where $\{d_{i1},d_{i2},\ldots,d_{im} \}$ denote the distances of the current record $R_{i}$ to all other records in the data set $S$. If the type of attribute is categorical or integer, then the data value that has the highest frequency (mode) of occurrence in the corresponding columns of the nearest records is imputed. For the data values of type real we first collect all non zero elements in the set $D$ and denote this set by $B$. For each element in set $B$ we compute the quantity $\beta(j)$=$\frac{1}{B(j)}$ $\forall j=1,\ldots,\gamma$ where $\gamma$ denote the cardinality of the set $B$. We compute the weight matrix as $W(j)=\frac{\beta_{j}}{\sum_{i=1}^{\gamma} \beta(i)}$ $\forall j=1,\ldots,\gamma$. The value to be imputed may be taken as $\sum_{i=1}^{\gamma} P(i)*W(i)$.
\subsection{Particle swarm optimization search for feature subset selection (Risk factors)}
\label{FS}
A PSO search consists of a set of particles initialized with a candidate solution to a problem. Each particle is associated with a position vector and a velocity vector. The particles evaluate the fitness of the solutions iteratively and store the location where they had their best fit known as the local best ($L$). The particles change their position and velocity iteratively in a suitable manner with respect to the best fit solution to reach a global optimal solution. The best fit solution among the particles is called the global best ($G$). We represent the position vector of the particle as a binary string and accuracy of the learning algorithm as the fitness function for evaluation. The velocity and position vectors of the particles are modified using the procedure suggested in~\cite{Kennedy2001}.
\begin{algorithm}[!t]\scriptsize
\caption{The $PRAA$ Methodology}
\label{chap9:PRA}
\begin{algorithmic}
\Require
\begin{enumerate}[label=(\alph{*})]
\item Data sets for the purpose of decision making $S(m,n)$ where $m$ and $n$ are number of records and attributes respectively and the members of $S$ may have MV in any of the attributes except in the decision attribute, which is the last attribute in the record.
\item The type of attribute $C$ of the columns in the data set.
\end{enumerate}
\Ensure
\begin{enumerate}[label=(\alph{*})]
\item Classification accuracy for a given data set $S$.
\item Performance metrics AUC, SE, SP.
\end{enumerate}
\textbf{Algorithm}
\begin{enumerate}[label=(\arabic{*})]
\item Identify and collect all records in a data set $S$
\item Impute the MV in the data set $S$ using the procedure discussed in Section~\ref{chap9:RNI}.
\item Extract the influential features using a wrapper based approach with particle swarm optimization search for identifying feature subsets and ADT for its evaluation as discussed in Section~\ref{FS}.
\item Split the dataset in to training and testing sets using a stratified $k$ fold cross validation procedure. Denote each training and testing data set by $T_{k}$ and $R_{k}$ respectively.
\item For each $k$ compute the following
\begin{enumerate}[label=(\roman{*})]\label{NRSproc}
\item Build the ADT using the records obtained from $T_{k}$. \label{buildadt}
\item Compute the predicted probabilities (scores) for both positive and negative diagnosis of CHD from the ADT built in Step (5)-(i) using the test data set $R_{k}$. Designate the set consisting of all these scores by $P$.
\item Identify and collect the actual diagnosis from the test data set $R_{k}$ in to set denoted by $L$.\label{step3}
\end{enumerate}
\item Repeat the Steps (5)-(i) to Step (5)-(iii) for each fold.
\item Obtain the performance metrics AUC, SE and SP utilizing the sets $L$ and $P$.
\item RETURN AUC, SE, SP.
\item END.
\end{enumerate}
\end{algorithmic}
\end{algorithm}
\section{Experiments and Results}
\label{chap8:sec3}
\subsection{Dataset}
In the present study we use the STULONG dataset~\cite{pkdd2004} which is a longitudinal primary preventive study of middle-aged men lasting twenty years for accessing the risk of atherosclerosis and cardiovascular health depending on personal and family history collected at Institute of Clinical and Experimental Medicine (IKEM) in Praha and the Medicine Faculty at Charles University in Plzen (Pilsen). The STULONG dataset is divided into four sub-groups namely Entry, Letter, Control and Death. The Entry dataset consists of $1417$ patient records with $64$ attributes having either codes or results of size measurements of different variables or results of transformations of the rest of the attributes during the first level examination. We utilize the Entry, Control and Death datasets for our predictive modeling. The Entry level dataset is divided into three groups (a) normal group (NG), (b) pathological group (PG), (c) risk group (RG), and (d) not allotted (NA) group, based on the studied group of patients (KONSKUP) in ($1$, $2$), ($5$), ($3$, $4$), ($6$) respectively. We form a new dataset by joining Entry, Control and Death datasets as follows: (i) we write the identification number of a patient (ICO) based on the selection criteria suggested in~\cite{Hoa2003} and determine the susceptibility of a patient to atherosclerosis based on the attributes recorded in Control and Death tables. An individual is considered to have cardiovascular disease if he or she has history of heart disease (i.e., he or she has at least one positive value on attributes such as myocardial infarction (HODN2), cerebrovascular accident (HODN3), myocardial ischaemia (HODN13), silent myocardial infarction (HODN14)), or died of heart disease (i.e., the record appears in the Death table with PRICUMR attribute equal to $05$ (myocardial infarction), $06$ (coronary heart disease), $07$ (stroke), or $17$ (general atherosclerosis)). Based on the above definition we divide the Entry dataset in to two datasets DS1 and DS2 depending on whether the patients are in NG or RG group respectively.
\subsection{Description of Experiments}
In our methodology we have employed a stratified ten-fold cross validation ( $k=10$) procedure. We have applied a standard implementation of SVM with radial basis function kernel using LibSVM package~\cite{chang2001}. We have taken the following standard parameter values for PSO (i) number of particles $Z=50$, (ii) number of iterations $G=100$, (iii) cognitive factor $c_1=2$, and (iv) social factor $c_2=2$. The standard implementation of C4.5, Naive Bayes (NB), Multi Layer Perceptron (MLP) algorithms in Weka$^\copyright$~\cite{Witten2005} are considered for evaluating the performance of our algorithm. An implementation of correlation based feature selection (CFS)~\cite{Liu1996} algorithm with genetic search has been considered for comparing with our methodology. Also, we have implemented the $PRAA$ algorithm and the performance evaluation methods in Matlab$^\copyright$. A non-parametric statistical test proposed by Wilcoxon~\cite{Wilcoxon1945} is used to compare the performance of the algorithms.
\subsection{Performance Measures and Results}
The PRAA methodology has outperformed (see Table~\ref{chap8:tab1}) the classifiers C4.5, SVM, MLP and NB in terms of sensitivity (SE), specificity (SP) and AUC performance metrics. In risk group dataset DS2 our methodology could identify the patients with an accuracy of $99.73\%$ who are affected by atherosclerosis using only $13$ out of $51$ attributes. The wrapper based feature selection using PSO and ADT could identify the influential factors such as alcohol (ALKOHOL), daily consumption of tea (CAJ), hypertension or ictus (ICT), hyperlipoproteinemia (HYPLIP), since how long hyper tension (HT) has appeared (HTTRV), before how many years hyperlipidemia had appeared (HYPLTRV), blood pressure II systolic (DIAST2), cholesterol in mg \% (CHLST), Glucose in urine (MOC), obesity (OBEZRISK), hypertension (HTRISK) which are in conformity with other studies related to cardiovascular diseases~\cite{McGill2002,Drechsler2010,Lydia2003}. We have identified an important fact that even in normal group DS1 individuals who mostly confine to sitting positions without any physical activity (AKTPOZAM=1) may lead to atherosclerosis as observed in~\cite{Thompson2003}.
\begin{table}[!t]\scriptsize
\centering
\caption{Performance comparison of the $PRAA$ with other methodologies (C4.5, SVM, NB, MLP) on the data sets used in the present study}
\label{chap8:tab1}
\begin{tabular}{c||c||c||c||c||c}
\hline
\bf{Dataset}&\bf{Method}&\bf{Accuracy}&\bf{SE}&\bf{SP}&\bf{AUC} \\
&&(\%)&&&\\
\hline \hline
\multirow{6}{*}{DS1}&\bf $PRAA$&\bf 98.04& \bf 93.75& \bf 100.00& \bf 0.94\\
&C4.5& 70.59& 6.25&100.00& 1.00\\
&NB& 52.94& 56.25& 51.43& 0.53\\
&SVM& 35.29&100.00& 5.71& 0.53\\
&MLP& 66.67& 0.00& 97.14& 0.97\\
\hline
\multirow{6}{*}{DS2}&\bf $PRAA$&\bf 99.73& \bf 99.35&\bf 100.00&\bf 1.00\\
&C4.5& 50.00& 55.48& 45.97& 0.52\\
&NB& 61.20& 48.39& 70.62& 0.58\\
&SVM& 45.63& 92.26& 11.37& 0.56\\
&MLP& 57.92& 0.65&100.00& 1.00\\
\hline
\end{tabular}
\end{table}
The ADT generated for the risk group (DS2) is shown in Fig.~\ref{chap9:adt}.
\begin{figure}[!t]
\centering
\includegraphics[width=0.5\textwidth]{DS2-ADT.jpg}\\
\caption{Alternating Decision tree generated for dataset DS2}\label{chap9:adt}
\end{figure}
The following decision rules are extracted from the decision tree shown in Fig.~\ref{chap9:adt}:
\begin{enumerate}
\item The risk of atherosclerosis increases by a factor of $-2.098$ if an individual is suffering from hyperlipidemia since $1.5$ years
\item The presence of hypertension ($< 3.5$ years) would increase the risk of atherosclerosis by a factor of $-1.348$;
\item The risk is estimated as $-2.098 -1.348 - 2.849 -0.966-0.43-0.249=-7.94$ if an individual is suffering from hyperlipidemia from anywhere between $1.5$ years to $4$ years ($HYPLTRV \ge 1.5$ and $<4.0$), with hypertension since $3.5$ years and cholesterol levels less than $166$. It is observed that the presence of hyperlipoproteinemia and Glucose in urine would increase the risk of atherosclerosis.
\end{enumerate}
\section{Conclusions}
\label{chap8:sec6}
A new methodology ($PRAA$) with built in features for imputation of missing values that can be applied on datasets wherein the attribute values are either normal and/or highly skewed having either categorical and/or numeric attributes and identification of risk factors using wrapper based feature selection is discussed. The $PRAA$ methodology has outperformed over the state-of-the-art methodologies in determining the risk factors associated with the onset of atherosclerosis disease. The $PRAA$ methodology has generated a decision tree with an accuracy of $99.7\%$ for dataset DS2. Based on the performance measures we conclude that the use of PSO search for feature subset selection wrapped around ADT embedded with new imputation strategy for fitness evaluation have improved the prediction accuracies.
\section{Discussion}
\label{chap8:sec7}
In this section we present a discussion on the performance of $PRAA$ methodology on benchmark datasets, its computational complexity, scalability and comparative studies with other methodologies.
\subsection{Performance Comparison of new imputation procedure on Benchmark Datasets}
\label{perMV}
Since no specific studies on imputation of missing values in cardiovascular disease data sets are available in the literature we have utilized some bench mark data sets obtained from Keel and University of California Irvin (UCI) machine learning data repositories~\cite{Alcala-Fdez2010,Frank2010} to test the performance of the new imputation algorithm. The Wilcoxon statistics in Table~\ref{chap8:tab5} is computed based on the accuracies obtained by the new imputation algorithm with the accuracies of those obtained by using embedded algorithms for handling missing values such as C4.5 decision tree. The results in Table~\ref{chap8:tab5} below clearly demonstrate that the imputation procedure presented in Section~\ref{chap9:RNI} has superior performance when compared to other imputation algorithms as the test statistics are well below or equal to the critical values with $p<0.05$ in all cases.
\begin{table}[H]\scriptsize
\caption{Wilcoxon sign rank statistics for matched pairs comparing the new imputation algorithm with other imputation methods using C4.5 decision tree}
\label {chap8:tab5}
\centering
\begin{tabular}{l||c||c||c||c}
\hline
\bfseries Method & \bfseries Rank Sums &\bfseries Test& \bfseries Critical& \bfseries p-value \\
& \bfseries (+, -)&\bfseries Statistics&\bfseries Value&\\
\hline \hline
FKMI & 28.0, 0.0& 0.0& 3 &0.02\\ \hline
KMI &28.0, 0.0 &0.0 &3 &0.02\\ \hline
KNNI & 15.0, 0.0& 0.0&0 &0.06\\ \hline
WKNNI & 21.0, 0.0 &1.0 &18 &0.03\\ \hline
\end{tabular}
\end{table}
\subsection{Comparison with other Related Methodologies on Atherosclerosis}
\label{chap8:sec5}
In this section we compare the results obtained in~\cite{Kukar1999, Tsang-Hsiang2006,Hongzong2007} with the results of our new methodology on the risk group dataset DS2. As compared to~\cite{Tsang-Hsiang2006} where in CFS with genetic search and C4.5 is employed, an SE of $39.4\%$ and SP of $82.8\%$ is observed. In~\cite{Hongzong2007} SVM was used to classify the patients with an SE of $95\%$ and SP of $90\%$. In~\cite{Kukar1999} both NB and MLP were used for classification with an accuracy of $80\%$ and the could obtain an SE of $92\%$ and $82\%$, an SP of $53\%$ and $76\%$ respectively. The new methodology when applied on the dataset DS2 resulted in an accuracy of $99.73\%$, an SE of $99.35\%$ and an SP of $100\%$ which is regarded as a good classification model since both SE and SP are higher than $80\%$.
\subsection{Computational Complexity and Scalability of $PRAA$ Algorithm}
\label{chap8:sec4}
The computational complexity is a measure of the performance of the algorithm which can be measured in terms of the number of CPU clock cycles elapsed in seconds for performing the methodology on a dataset. For each data set having $n$ attributes and $m$ records, we select only those subset of records $m_{1} \le m$, in which missing values are present. The distances are computed for all attributes $n$ excluding the decision attribute. So, the time complexity for computing the distance would be $O(m_{1}*(n-1))$. The time complexity for computing skewness is $O(m_{1})$. The time complexity for selecting the nearest records is of order $O(m_{1})$. For computing the frequency of occurrences for nominal attributes and weighted average for numeric attributes the time taken would be of the order $O(m_{1})$. Therefore, for a given data set with $k$-fold cross validation having $n$ attributes and $m$ records, the time complexity of our new imputation algorithm would be $k*(O(m_{1}*(n-1)*m)+3*O(m_{1}))$ which is asymptotically linear. Our experiments were conducted on a personal computer having an Intel(R) core (TM) 2 Duo, CPU @$2.93$ GHZ processor with $4$ GB RAM and the time taken by $PRAA$ for varying database sizes is shown Fig.~\ref{chap8:fig6}. We have employed a linear regression on our results and obtained the relation between the time taken (T) and the data size (D) as $T=14.909D-104.655$, $\alpha=0.05$, $p=0.0003$, $r^2=0.994$.
\begin{figure}[!t]
\centering
\includegraphics[width=0.5\textwidth]{timecomplexity-new.jpg}\\
\caption{Computational complexity of the $PRAA$}\label{chap8:fig6}
\end{figure}
The presence of the linear trend between the time taken and the varying database sizes ensure the numerical scalability of the performance of $PRAA$ methodology. A comparison with other related methodologies used in the study of atherosclerosis yields a conclusion that the new $PRAA$ methodology presented in this paper has a superior performance over other methods studied in~\cite{Kukar1999, Tsang-Hsiang2006,Hongzong2007}. We hold the view that more intensive and introspective studies of this kind will pave way for effective risk prediction and diagnosis of atherosclerosis.
|
\section{Introduction} \label{sec:Intro}
We begin by briefly summarising some historical details regarding fractional level $\AKMA{sl}{2}$ models and the application of Jack symmetric functions to conformal field theory{}. The work presented here forms a part of of an ambitious project aimed at elucidating the properties of general fractional level models as fundamental examples of logarithmic conformal field theories{}. A short digression on the notion of admissibility follows as we use the term in a non-standard manner as compared to much of the literature. This also provides us with an opportunity to fix some notation. Finally, we outline the main results of the research reported here.
\subsection{Fractional level models and Jack symmetric functions}
The fractional level $\AKMA{sl}{2}$ models are conformal field theories{} with a long and notorious history, originally proposed by Kent \cite{KenInf86} as non-unitary models whose existence would lead to a uniform coset construction for all the Virasoro minimal models. This proposal received a significant boost from the subsequent announcement of Kac and Wakimoto \cite{KacMod88} that $\AKMA{sl}{2}$ has, for precisely the levels required for this coset construction, a finite set of simple \hw{} modules{} whose characters close under modular transformations. Moreover, they observed that this property persisted for higher rank affine Kac-Moody algebras. However, computations \cite{KohFus88,BerFoc90,AwaFus92} of the fusion rules of the purported theories gave confusing and conflicting results. In particular, substituting the modular S-matrix entries into the Verlinde formula resulted in negative fusion multiplicities. Despite a flurry of subsequent work, no resolution was agreed upon; in their discussion, the authors of the textbook \cite{DiFCon97} suggested that the fractional level models may be ``intrinsically sick''.
The first steps towards curing this sickness were made by Gaberdiel \cite{GabFus01}, whose explicit fusion computations for the $\AKMA{sl}{2}$ model of level $k=-\tfrac{4}{3}$ demonstrated that one was forced to consider modules that were not highest weight{}. Indeed, he found that fusing the \hw{} modules{} of Kac and Wakimoto resulted in infinite numbers of modules whose conformal weights were not bounded below, modules that were not highest weight{} with respect to any Borel subalgebra (relaxed modules), and modules upon which the Virasoro mode $L_0$ acted non-semisimply. This showed that the $k=-\tfrac{4}{3}$ model was not a rational conformal field theory{}, as had been implicitly assumed in earlier studies, but was, in fact, logarithmic. Subsequent works \cite{LesSU202,LesLog04,RidSL210,RidFus10} extended these results to $k=-\tfrac{1}{2}$ and thence to the closely related $\beta \gamma$ ghost systems \cite{RidSL208,RidFus10,RidBos14}. Moreover, it was also shown \cite{RidSL208} that a misunderstanding concerning the role of convergence regions in the modular transformations of Kac and Wakimoto was to blame for the failure of the Verlinde formula. To get a genuine action of the modular group and a working Verlinde formula, one needs to include all the simple relaxed modules and all their twists under spectral flow \cite{CreMod12,CreMod13}.
We regard these recent advances as demonstrating that the sickness of the $\AKMA{sl}{2}$ models has been cured and we expect that this cure will be just as effective for higher rank Kac-Moody algebras. Given the logarithmic nature of the theories, it is now reasonable to expect that the fractional level models will play an important role in understanding logarithmic conformal field theory{}, much as the non-negative integer level Wess-Zumino-Witten{} models did for rational conformal field theory{}. We therefore view the fractional level models as objects that deserve intense study. In particular, one should at least determine the spectrum of simple modules, compute character formulae and verify that the modular properties of these characters lead to non-negative integers upon applying the Verlinde formula. More ambitiously, one would like to understand the fusion ring generated by the simple modules, the structure of the indecomposable modules generated by fusion, and the corresponding three- and four-point correlation functions. From a mathematical perspective, one can ask about homological properties of the spectrum (a category of modules over the corresponding vertex operator algebra{}) including rigidity and the identification of the projective and injective modules. We think that it is not unreasonable to expect that there are beautiful answers to all these questions, given that they concern structures built from affine Kac-Moody algebras.
While some, though not all, of these questions have been answered for $\AKMA{sl}{2}$ at levels $k=-\tfrac{4}{3}$ and $-\tfrac{1}{2}$, some of the machinery employed is clearly unfeasible for more general levels and ranks. In our opinion, it is likely that the modular story will remain under control within the so-called standard module framework \cite{CreLog13,RidVer14}. On the other hand, brute force methods such as the Nahm-Gaberdiel-Kausch algorithm \cite{NahQua94,GabInd96} for fusion products are already computationally-prohibitive in all but the simplest cases. Further progress will instead require the development of more general alternatives and free field realisations seem to be the obvious candidates in this respect. Here, we use a free field realisation to address one of the most basic questions of all, that of determining the simple modules of a fractional level theory. We restrict ourselves to the $\AKMA{sl}{2}$ models as a testing ground, leaving the challenge of higher ranks for future publications.
The standard free field realisation of an affine Kac-Moody algebra is due to Wakimoto \cite{WakFoc86}, for $\AKMA{sl}{2}$, and Feigin and Frenkel \cite{FeiFam88} in general. For $\AKMA{sl}{2}$, the Wakimoto realisation combines a free boson with a pair of bosonic ghosts of central charge $c=2$. The virtue of free field theories such as these is that their underlying Lie algebras are almost abelian. More precisely, the negative modes (creation operators) all commute among themselves, hence one can invoke symmetric group theory, in particular symmetric polynomials and functions, to analyse certain representation-theoretic questions. For the question of classifying the simple modules of the fractional level $\AKMA{sl}{2}$ models, it turns out that the key lies in the Jack symmetric functions \cite{JacCla70}.
The relevance of Jack symmetric functions to conformal field theory{} goes back to the work of Mimachi and Yamada \cite{MimSin95} who realised that they provide elegant expressions for the singular vectors{} of Verma modules over the Virasoro algebra. Explicit singular vector{} formulae are useful for many field-theoretic investigations including those of the spectrum of primary fields, the fusion rules and the correlation functions. However, the general singular vector{} formulae that were then known, for example those of \cite{BenDeg88,BauCov91}, are not particularly tractable for these purposes. On the other hand, the Jack function formulae were derived directly from the Feigin-Fuchs free field realisation of the Virasoro algebra (also known as the Coulomb gas), as developed by Tsuchiya and Kanie \cite{TsuFoc86} and Felder \cite{FelBRST89}, which is far better suited to explicit computation.
Unfortunately, it appears that the power of symmetric function theory was not immediately exploited in conformal field theory{} studies, perhaps because of an unfamiliarity with Jack symmetric functions (Macdonald's influential textbook \cite{MacSym95} did not appear until several years later). However, in more recent times, symmetric function theory has been embraced by the community, particularly as a means to prove the AGT conjecture \cite{AldLio10} which relates Liouville conformal field theories{} to the instanton calculus of Yang-Mills theories (although it now appears that a generalisation of the Jack symmetric functions will be required \cite{MorTow14} to prove this conjecture).
The work reported here, using Jack symmetric functions to classify simple modules of the fractional level $\AKMA{sl}{2}$ models, has its genesis in \cite{TsuExt13} where this formalism was used to classify, among other things, the simple modules of a family \cite{FeiLog06} of (logarithmic) conformal field theories{} called the $(p_+,p_-)$ triplet models. The methods developed for this purpose were also applied to give a far more elegant proof of the singular vector{} formulae of Mimachi and Yamada. More recently \cite{RidJac14}, it was shown that these methods lead to an elegant new proof of the classification of the Virasoro minimal model modules. We recall that the original classification proof of Wang \cite{WanRat93} combined a projection formula for the singular vector{} of the vacuum module, stated by Feigin and Fuchs and eventually proven (fifteen years later) by Astashkevich and Fuchs \cite{AstAsy97}, with some intricate cohomological arguments.
The content of this paper is then that these simplified methods generalise to the fractional level $\AKMA{sl}{2}$ models. Specifically, we deduce explicit formulae for singular vectors{} in Wakimoto modules and classify the simple modules. We take this as strong evidence that the symmetric function techniques developed here will further generalise to models over higher rank Kac-Moody algebras and superalgebras. Moreover, it is clear that these techniques can be profitably exploited to investigate other important representation-theoretic questions including the structure theory of relaxed Verma and Wakimoto modules. We hope to report on this in the future.
It should be emphasised that, as with the Virasoro case reported in \cite{RidJac14}, the $\AKMA{sl}{2}$ classification result is not new. The highest weight{} classification is due to Adamovi\'{c} and Milas \cite{AdaVer95} and, independently, Dong, Li and Mason \cite{DonVer97}. However, their proofs mimic that of Wang, relying upon a projection formula for the singular vector{} of the vacuum module stated (without proof) by Fuchs \cite{FucTwo89}. From this projection formula, Adamovi\'{c} and Milas also derive a classification result for what we call, following \cite{FeiEqu98}, the \emph{relaxed} \hw{} modules{}.\footnote{The existence of even more general weight modules follows directly from twisting by the spectral flow automorphisms of $\AKMA{sl}{2}$. Well known in the physics literature, this twisting is an important ingredient in \cite{FeiEqu98,MalStr01} and first seems to have been explicitly noted for fractional level $\AKMA{sl}{2}$ models in \cite{GabFus01}.} It is not clear to us if these projection formulae will generalise easily to higher ranks; the tedium of the proof in the Virasoro case alone warrants, in our opinion, the development of the rather more elegant symmetric function methods. With this in mind, we remark that, to the best of our knowledge, there are no general classification results known for fractional level models of rank greater than $1$.
\subsection{Basic concepts and notation}
Suppose that we have a conformal field theory{} whose chiral algebra is identified as a Lie (super)algebra $\alg{g}$, for example, the Virasoro algebra or an affine Kac-Moody (super)algebra. More general chiral algebras can be accommodated within the formalism to follow, but this level of generality will suffice for the purposes of the article. There is always a module of the chiral algebra, called the vacuum module, that carries the structure of a vertex operator algebra{} $\VOA{V}$. The elements of the chiral algebra are organised into fields that generate $\VOA{V}$ and the (anti)commutation relations of the chiral algebra are equivalent to the operator product expansions{} of these generating fields. We make the following definition:
\begin{defn}
Consider a vertex operator algebra{} $\VOA{V}$ corresponding to a Lie (super)algebra $\alg{g}$ as above. If the operator product expansions{} of the generating fields constitute a complete set of algebraic relations, then $\VOA{V}$ is said to be \emph{universal}.
\end{defn}
\noindent The terminology comes from noting that a vertex operator algebra{} is a quotient of $\VOA{V}$ if, and only if, it has a set of generating fields that satisfy the same operator product expansions{} as those of $\VOA{V}$.
One way to understand universality is to consider what it means for the vacuum module. When $\alg{g}$ is the Virasoro algebra, the vacuum module of the universal vertex operator algebra{} is the quotient of the Verma module whose (generating) \hw{} vector{} has conformal weight $0$ by the Verma submodule generated by the singular vector{} of conformal weight $1$ (the vacuum must be translation-invariant). For $\alg{g} = \AKMA{sl}{2}$, one quotients the Verma module generated by the \hw{} vector{} of $\SLA{sl}{2}$-weight $0$ by the Verma submodule generated by the singular vector{} of $\SLA{sl}{2}$-weight $-2$ (as required by the state-field correspondence). In both cases, it may happen that this universal vacuum module is not simple. But, quotienting by a non-trivial proper submodule amounts to imposing additional relations upon the vertex operator algebra{}, so the result would no longer qualify as universal.
Suppose now that we have a parametrised family of universal vertex operator algebras{}. For example, the Virasoro algebra and $\AKMA{sl}{2}$ each define a one-parameter family of vertex operator algebras{}, parametrised by the central charge $c \in \mathbb{C}$ and the level $k \in \mathbb{C} \setminus \set{-2}$, respectively. We characterise the non-simple members of this family.
\begin{defn}
Suppose that one has a family of universal vertex operator algebras{} $\set{\VOA{V}_i}_{i \in I}$, parametrised by some index set $I$. A given value $i$ of the parameter is said to be \emph{admissible} if the corresponding universal vertex operator algebra{} $\VOA{V}_i$ is not simple.
\end{defn}
\noindent For the Virasoro algebra, the admissible central charges $c$ are then precisely those for which there exist coprime integers $p,p' \in \mathbb{Z}_{\ge 2}$ satisfying $c = 13 - 6(t + t^{-1})$, where $t = \tfrac{p}{p'}$. These central charges correspond, of course, to the Virasoro minimal models which are commonly denoted by $\MinMod{p}{p'}$. We will also denote the simple Virasoro vertex operator algebra{} of this (admissible) central charge by $\MinMod{p}{p'}$, for convenience.
For $\AKMA{sl}{2}$, the structure theory of its Verma modules \cite{KacStr79} leads to the following characterisation:
\begin{prop}
The admissible levels \(k\) of the universal vertex operator algebras{} of $\AKMA{sl}{2}$ are precisely those for which there exist coprime integers \(u\in\mathbb{Z}_{\ge2}\) and \(v\in\mathbb{Z}_{\ge1}\) satisfying $k=-2+t$, where \(t=\frac{u}{v}\).
\end{prop}
\noindent Needless to say, the admissible levels of $\AKMA{sl}{2}$ are precisely those of the fractional level $\AKMA{sl}{2}$ models (including those of non-negative integer level, for convenience). To emphasise the analogy with the Virasoro minimal models $\MinMod{p}{p'}$, we shall denote by $\AdmMod{u}{v}$ both the fractional level $\AKMA{sl}{2}$ model with $k = -2 + \tfrac{u}{v}$ and the corresponding simple vertex operator algebra{}. We regard the $\AdmMod{u}{v}$ as the minimal models of $\AKMA{sl}{2}$. Note that $\AdmMod{k+2}{1}$ requires \(k\in\mathbb{Z}_{\ge 0}\) and is therefore just the level $k$ Wess-Zumino-Witten{} model on $\SLG{SU}{2}$.
It should be clear now that the focus of our interest is not so much on the universal vertex operator algebras{} themselves, but rather on their admissible level simple quotients. The point is that these simple quotients will have constrained representation theories, due to their additional defining relations, about which we expect to be able to prove classification theorems. The representation theory of the universal vertex operator algebras{} is, on the other hand, unconstrained by additional relations so that (almost) every $\alg{g}$-module is allowed.\footnote{One should only exclude modules, such as the adjoint module, that lead to operator product expansions{} with essential singularities.}
To complete the analogy between Virasoro minimal models and fractional level $\AKMA{sl}{2}$ models, we consider the modules of the simple quotient vertex operator algebras{}.
\begin{defn}
Suppose that $i \in I$ is admissible for a given family $\set{\VOA{V}_i}$ of universal vertex operator algebras{}. Then, any module of the simple quotient of $\VOA{V}_i$ is said to be an \emph{admissible module} of $\VOA{V}_i$.
\end{defn}
\noindent Note that the admissible modules of the universal Virasoro vertex operator algebra{} are precisely the central charge $c = 13 - 6(\tfrac{p}{p'} + \tfrac{p'}{p})$ modules of the Virasoro minimal model $\MinMod{p}{p'}$. These are the highest weight{} Virasoro modules corresponding to the entries in the Kac table. The admissible modules of the universal $\AKMA{sl}{2}$ vertex operator algebras{} are the subject of this article. As noted above, they were first classified in \cite{AdaVer95,DonVer97}. We remark that they can also be associated with the entries of a table with similar properties to the Virasoro Kac table, see \cite{CreMod13}.
Finally, note that the definition of admissible module given above is not the original definition of Kac and Wakimoto \cite{KacMod88}, who originally defined admissible \hw{} modules{}, for arbitrary affine Kac-Moody algebras, in terms of criteria that guaranteed a character formula, generalising that of Weyl-Kac, and good modular properties. The vertex operator algebra{} definition given above is certainly more general and, in our opinion, more fundamental. However, the two definitions of admissibility coincide for highest weight{} $\AKMA{sl}{2}$-modules. A generalised version of this coincidence for higher rank affine Kac-Moody algebras has recently appeared in \cite{AraRat15}.
\subsection{Outline}
We close with a brief outline of the contents of this paper. \secref{sec:GenHWTh} is a pedagogical introduction to the notion of relaxed \hw{}{} theory, crucial for studying the $\AdmMod{u}{v}$ models. The idea actually reduces to a special case of parabolic (also called parahoric or generalised) highest weight{} theory, but the connection to vertex operator algebras{} via Zhu's algebra is so important that we feel it warrants separate consideration. This is then followed by a detailed discussion of the \rhw{} modules{} of the three Lie algebras used in the remainder of the paper: The Heisenberg algebra, the (bosonic) $\beta \gamma$ ghost algebra and, of course, $\AKMA{sl}{2}$.
\secref{sec:Wak} opens with a brief review of the Wakimoto free field realisation of $\AKMA{sl}{2}$ and the corresponding Wakimoto modules. We pay particular attention to the screening fields and operators as a means to motivate the usage of symmetric function theory. We then discuss the construction of certain singular vectors{} in highest weight{} and relaxed \hw{}{} $\AKMA{sl}{2}$-modules using Jack symmetric function technology to deduce explicit formulae for these (\thmref{thm:ConstructSVs}). We believe that these formulae are new: Even in the standard highest weight{} case, the only similar result we are aware of is an old paper of Kato and Yamada \cite{KatMis92} where an integral formula is derived and evaluated in special cases using Schur polynomials. For singular vectors{} of \rhw{} modules{}, the only other formulae we know are of the Malikov-Feigin-Fuchs (complex power) form \cite{FeiEqu98}.
One consequence of this singular vector{} study is that at admissible levels, Wakimoto's construction yields a free field realisation of the universal vertex operator algebra{} of $\AKMA{sl}{2}$ rather than of its simple quotient $\AdmMod{u}{v}$. This is important as it means that the singular vector{} of the universal vacuum module is accessible in the Wakimoto realisation, hence it may be exploited to such ends as determining the spectrum of $\AdmMod{u}{v}$-modules. We first obtain an upper bound on the highest weight{} spectrum (\corref{cor:ClassHWMods}) using a surprisingly effortless calculation that combines the form of the vacuum singular vector{} with the specialisation formula for Jack symmetric functions. This bound turns out to be saturated, but to prove this we must address the more involved relaxed \hw{}{} spectrum. In this case, a few more symmetric function manipulations allow us to identify an explicit presentation for Zhu's algebra (\thmref{thm:ComputeZhu}) in terms of generators and relations.
It is now easy to classify the simple \rhw{} modules{} of $\AdmMod{u}{v}$, reproducing in an elegant fashion the results of \cite{AdaVer95,DonVer97}. We emphasise that our proofs are, to the best of our knowledge, independent of those which have appeared in the literature. We also deduce that highest weight{} $\AdmMod{u}{v}$-modules are semisimple, giving a new proof of an old result that (essentially) appeared in \cite{KacMod88}. In contrast, $\AdmMod{u}{v}$ is shown to admit non-semisimple \rhw{} modules{} with two composition factors and these are characterised using short exact sequences. We moreover conjecture that these modules, together with their simple analogues, exhaust the indecomposable relaxed \hw{}{} $\AdmMod{u}{v}$-modules, remarking that proving this would require, among other things, a more detailed knowledge of the submodule structure of relaxed \hw{}{} $\AKMA{sl}{2}$-modules. Finally, we prove that the Virasoro zero mode $L_0$ acts semisimply on all relaxed \hw{}{} $\AdmMod{u}{v}$-modules, independent of our conjecture on the indecomposable spectrum, and discuss briefly how this is consistent with the expectation that, for $v>1$, the $\AdmMod{u}{v}$ model is a logarithmic conformal field theory{}.
\appref{sec:SymmPoly} gives a brief, but thorough, introduction to the aspects of symmetric function theory, in particular, those relating to Jack symmetric functions, that are used in the text. This material is all standard and may be found in \cite{MacSym95}. We have also included a brief introduction to Zhu's algebra in \appref{sec:zhu}, concentrating on motivating it as an abstract version of the algebra of zero modes acting on the space of ``ground states'' of a \rhw{} module{}. This appendix is aimed at physicists, in particular, it uses physics conventions for Fourier expansions, but we hope that it will also prove useful to mathematicians.
\section{Generalising highest weight theory} \label{sec:GenHWTh}
In this section, we consider a generalisation of highest weight{} theory that we will qualify as \emph{relaxed}. Originally introduced for $\AKMA{sl}{2}$ in order to study a correspondence relating $\AKMA{sl}{2}$-modules to those over the $N=2$ superconformal algebra \cite{FeiEqu98}, \rhw{} modules{} have since appeared as necessary constituents of the $\SLG{SL}{2;\mathbb{R}}$ Wess-Zumino-Witten{} model \cite{MalStr01}, in admissible level fusion rules \cite{GabFus01,RidFus10}, in relations to logarithmic minimal models \cite{AdaCon05,RidSL210,CreCos13}, in demonstrating the modular invariance of admissible level theories \cite{CreMod12,CreMod13}, and in the full description of bosonic $\beta \gamma$ ghosts \cite{RidFus10,RidBos14}. Necessity aside, we feel that from some points of view, particularly that of Zhu's algebra \cite{ZhuMod96}, discussed in \appref{sec:zhu}, it is more natural to consider these relaxed modules instead of the standard \hw{} modules{} that one typically encounters in rational conformal field theory{}.
\subsection{Relaxed highest weight theory} \label{sec:Relaxed}
We recall that the formalism of highest weight{} theory for a Lie algebra $\alg{g}$ is built from a triangular decomposition
\begin{equation} \label{eq:TriDec}
\alg{g} = \alg{g}_- \oplus \alg{h} \oplus \alg{g}_+.
\end{equation}
This is a vector space direct sum of subalgebras of $\alg{g}$ in which the Cartan subalgebra $\alg{h}$ is abelian and acts semisimply, through the adjoint action, on both $\alg{g}_-$ and $\alg{g}_+$. In particular, $\comm{\alg{h}}{\alg{g}_{\pm}} \subseteq \alg{g}_{\pm}$. Moreover, the subalgebras $\alg{g}_-$ and $\alg{g}_+$ are assumed to be antiequivalent in that there exists a (linear) order two antiautomorphism, the adjoint $\dag$, satisfying $\alg{g}_{\pm}^{\dag} = \alg{g}_{\mp}$. The elements of $\alg{h}$ are supposed to be fixed by $\dag$, meaning that each $x \in \alg{h}$ is self-adjoint \cite{MooLie95}. This is, however, a little too restrictive; we instead only demand that the adjoint preserves the Cartan subalgebra. Since $\alg{h}$ is abelian, this implies that each $x \in \alg{h}$ is normal: $\comm{x}{x^{\dag}} = 0$. Note that being a linear antiautomorphism means that $\brac{ax}^{\dag} = a x^{\dag}$ and $\comm{x}{y}^{\dag} = \comm{y^{\dag}}{x^{\dag}}$, for all $a \in \mathbb{C}$ and $x,y \in \alg{g}$.
Given such a triangular decomposition, one defines a \hw{} vector{} in a given $\alg{g}$-module to be a simultaneous eigenvector of the elements of $\alg{h}$ which is annihilated by all the elements of $\alg{g}_+$. Any module generated by a single \hw{} vector{} is called a \hw{} module{}. Conspicuous examples include the Verma modules $\Ver{\lambda} = \UEA \alg{g} \otimes_{\UEA \alg{b}} \mathbb{C}_{\lambda}$, where $\UEA \alg{g}$ denotes the universal enveloping algebra{} of $\alg{g}$, $\UEA \alg{b}$ that of the Borel subalgebra $\alg{b} = \alg{h} \oplus \alg{g}_+$, $\lambda \in \alg{h}^*$ is the highest weight{}, and $\mathbb{C}_{\lambda}$ is the one-dimensional $\alg{b}$-module upon which $\alg{g}_+$ acts as $0$ and each $x \in \alg{h}$ acts as $\func{\lambda}{x} \in \mathbb{C}$. The \hw{} vector{} generating $\Ver{\lambda}$ is $\wun \otimes_{\UEA \alg{b}} 1_{\lambda}$, where $\wun$ denotes the unit of $\UEA \alg{g}$ (and $1_{\lambda}$ that of $\mathbb{C}_{\lambda} \cong \mathbb{C}$).
The Lie algebras $\alg{g}$ that one typically encounters in conformal field theory{} have triangular decompositions. However, they also have more structure in that they are graded by the semisimple action of the Virasoro zero mode $L_0$ so that, if $\alg{g}_n$ denotes the $\func{\ad}{L_0}$-eigenspace of eigenvalue $-n$, then $\comm{\alg{g}_m}{\alg{g}_n} \subseteq \alg{g}_{m+n}$. Moreover, the Cartan subalgebra $\alg{h}$ is generally chosen to include $L_0$, which we will always assume is self-adjoint, hence it follows that $\alg{g}_m^{\dag} = \alg{g}_{-m}$. Given this structure, the following definition is natural:
\begin{defn}
Let $\alg{g}$ be a Lie algebra with triangular decomposition \eqref{eq:TriDec} (where the Cartan subalgebra may include elements that are not self-adjoint). If there exists $L_0 \in \alg{h}$ such that $\alg{g} = \bigoplus_n \alg{g}_n$, where $\alg{g}_n$ is the eigenspace of $\func{\ad}{L_0}$ of eigenvalue $-n$, and $L_0$ is the zero mode of a subalgebra of $\alg{g}$ isomorphic to the Virasoro algebra, then we will say that $\alg{g}$ is \emph{conformally graded}.
\end{defn}
\noindent The most obvious example is the Virasoro algebra itself which is clearly conformally graded with $\alg{g}_n = \vspn \set{L_n}$, for $n \in \mathbb{Z} \setminus \set{0}$, $\alg{g}_0 = \vspn \set{L_0, C}$ and $\alg{g}_n = \set{0}$ otherwise. This shows that the usual, but somewhat confusing, convention that $\comm{L_0}{L_n} = -L_n$ is responsible for $\alg{g}_n$ having $\func{\ad}{L_0}$-eigenvalue $-n$ in the above definition.
\begin{defn}
Given a conformally graded Lie algebra $\alg{g}$, its \emph{relaxed triangular decomposition} is
\begin{equation} \label{eq:RelTriDec}
\alg{g} = \alg{g}_< \oplus \alg{g}_0 \oplus \alg{g}_>,
\end{equation}
where $\alg{g}_< = \bigoplus_{n<0} \alg{g}_n$ and $\alg{g}_> = \bigoplus_{n>0} \alg{g}_n$. A \emph{relaxed \hwv{}{}} is then a simultaneous eigenvector of $\alg{h} \subseteq \alg{g}_0$ that is annihilated by $\alg{g}_>$. A \emph{\rhw{} module{}} is a module that is generated by a single relaxed \hwv{}{}. The \emph{relaxed Borel subalgebra} is $\alg{g}_{\ge} = \alg{g}_0 \oplus \alg{g}_>$ and a \emph{relaxed Verma module} is a $\alg{g}$-module isomorphic to $\RelVer{\Mod{M}} = \UEA \alg{g} \otimes_{\UEA \alg{g}_{\ge}} \Mod{M}$, where $\Mod{M}$ is a simple weight module of $\alg{g}_0$ upon which $\alg{g}_>$ acts as $0$.
\end{defn}
\noindent These definitions have obvious analogues for Lie superalgebras and other more general structures, but we will not need this level of generality in what follows.
In this article, we will only consider triangular decompositions of a conformally graded Lie algebra $\alg{g}$ that satisfy $\alg{g}_< \subseteq \alg{g}_-$ and $\alg{g}_> \subseteq \alg{g}_+$. Thus, positive modes are always in $\alg{g}_+$ and negative modes are always in $\alg{g}_-$. When $\alg{g}_0 = \alg{h}$, the relaxed triangular decomposition then coincides with the (unique) triangular decomposition of this type. We will shortly see examples of \rhw{} modules{} which are not \hw{} modules{} in the usual sense. First, however, we mention that \rhw{} modules{} may be identified as generalised \hw{} modules{} with respect to the parabolic subalgebra $\alg{g}_{\ge} = \alg{g}_0 \oplus \alg{g}_>$. We recall that a parabolic subalgebra is any subalgebra that contains a Borel subalgebra (see \cite{MooLie95,HumRep08} for a quick overview of parabolic subalgebras).
It will be occasionally convenient to take this a step further and introduce a category of modules, generalising the well known category $\categ{O}$, that contains the \rhw{} modules{} of $\alg{g}$. This is essentially (a variant of) the parabolic category $\categ{O}$ discussed, for example, in \cite{HumRep08}. The explicit details of this category are not essential for understanding the results to follow, but we shall devote a few paragraphs to explaining their physical motivation.
\begin{defn}
Given a conformally graded Lie algebra $\alg{g} = \alg{g}_< \oplus \alg{g}_0 \oplus \alg{g}_>$, define the \emph{relaxed category $\categ{R}$} to consist of the $\alg{g}$-modules $\Mod{M}$ satisfying the following axioms:
\begin{itemize}
\item $\Mod{M}$ is finitely generated.
\item $\Mod{M}$ is a weight module (the action of the Cartan subalgebra $\alg{h}$ is semisimple).
\item The action of $\alg{g}_>$ is locally nilpotent: For each $v \in \Mod{M}$, the space $\UEA \alg{g}_> \cdot v$ is finite-dimensional.
\end{itemize}
The morphisms are the $\alg{g}$-module homomorphisms between these modules, as usual.
\end{defn}
\noindent All highest weight{} and \rhw{} modules{} belong to category $\categ{R}$. Moreover, if $\alg{g}$ has finite-dimensional root spaces, then it follows that each module in $\Mod{M}$ will have finite-dimensional weight spaces. Another important consequence of these axioms is that every (non-zero) module in category $\categ{R}$ possesses a relaxed \hwv{}{}, hence that every simple category $\categ{R}$ module is a \rhw{} module{}.
One can, and should, ask whether the mathematical axioms that we impose on category $\categ{R}$ will end up excluding modules relevant for applications. This is an important question and the answer is that they do exclude relevant modules, but in a well-controlled manner. Our motivation for introducing this category is that we want to classify the modules of the admissible level $k$ vertex operator algebra{} $\AdmMod{u}{v}$ by identifying these modules as $\AKMA{sl}{2}$-modules. For the physical application of investigating the corresponding conformal field theories{}, we must insist that the category of $\AdmMod{u}{v}$-modules be closed under the conjugation operation of $\AKMA{sl}{2}$ (see \secref{sec:RelEx}) and fusion. Moreover, we want the characters to behave well under modular transformations so that one can identify modular invariant partition functions and compute (Grothendieck) fusion rules from a Verlinde-type formula.
Category $\categ{O}$ is not sufficient for these purposes, in particular, the conjugate of an $\AdmMod{u}{v}$-module from category $\categ{O}$ need not lie in category $\categ{O}$. Relaxing to category $\categ{R}$ alleviates this problem and has recently been shown \cite{CreMod13} to lead to characters with excellent modular behaviour, provided that one extends the category again to take into account twists by the so-called spectral flow automorphisms. The upshot is that these spectrally-flowed modules are not in category $\categ{R}$, but the twisting is very well understood, justifying the above statement that the exclusion of these physically relevant modules is under control.
Axiomatically, accounting for the spectrally-flowed modules would require weakening the local nilpotency axiom above. However, this axiom has the advantage that category $\categ{R}$ provides a very natural setting for the important, and very useful, technology of Zhu's algebra, discussed in \appref{sec:zhu}. We restrict to weight modules because the fusion coproduct formulae \cite{GabFus94} for $\AKMA{sl}{2}$ show that the fusion product of two weight modules will again be weight. Similarly, the conjugate of a weight module is weight and omitting non-weight modules does not restrict the characters in any way. Moreover, being weight does not preclude the Virasoro zero mode $L_0$ from acting non-semisimply as is required in logarithmic conformal field theories{}.
To summarise, the relaxed category $\categ{R}$ is a rich source of modules for affine Kac-Moody (super)algebras that appears to be even more relevant to conformal field theory{} than the much more familiar category $\categ{O}$.
\subsection{Examples} \label{sec:RelEx}
In this paper, we will use the Wakimoto free field realisation to study the \rhw{} modules{} of $\AKMA{sl}{2}$ and determine which of these are modules for the admissible level vertex operator algebras{} $\AdmMod{u}{v}$, where $u \in \mathbb{Z}_{\ge 2}$ and $v \in \mathbb{Z}_{\ge 1}$ are coprime and the level $k$ is determined by $\frac{u}{v} = t = k+2$. We will therefore need to investigate the (relaxed) highest weight{} theory of the Heisenberg algebra $\HA$, the $c=2$ bosonic \(\beta\gamma\) ghost system $\GA$ and $\AKMA{sl}{2}$ itself. This investigation constitutes the rest of the section.
\subsection*{The Heisenberg algebra $\HA$}
We will use the same notations and conventions for the Heisenberg algebra as in \cite{RidJac14}. The free boson vertex operator algebra{} is generated by a single bosonic field \(a(z)\), defined by the operator product expansion{}
\begin{equation} \label{eq:HeisOPE}
a(z)a(w)\sim\frac{\wun}{(z-w)^2}.
\end{equation}
With the standard Fourier expansion, $a(z)=\sum_{n\in\mathbb{Z}}a_n z^{-n-1}$, the operator product expansion{} implies the following commutation relations:
\begin{equation} \label{eq:BosComm}
\comm{a_m}{a_n}=m\delta_{m+n,0}\wun, \qquad m,n\in\mathbb{Z}.
\end{equation}
The Heisenberg algebra $\HA$ is then the infinite-dimensional Lie algebra spanned by the $a_n$ and $\wun$, the latter being identified with the unit of the universal enveloping algebra{} of \(\HA\), as usual. We will assume that \(\wun\) acts as the identity operator on any \(\HA\)--module. This is only a minor restriction since a simple rescaling of the generators lets this operator act as multiplication by any non-zero number.
As is well known, the free boson vertex operator algebra{} admits a one-parameter family of conformal structures. We will write the energy-momentum tensor and central charge in the form
\begin{equation} \label{eq:HeisT}
T^{\textup{bos.}}(z)=\frac{1}{2}\normord{a(z)a(z)}-\frac{1}{\alpha}\partial a(z),\qquad c^{\textup{bos.}}=1-\frac{12}{\alpha^2},
\end{equation}
where $\alpha$ parametrises the conformal structure. We note that $\alpha \to \infty$ reproduces the standard free boson central charge $c^{\textup{bos.}} = 1$. In Wakimoto's construction, this would correspond to $k \to \infty$, so it is permissible to ignore this case. It is worth recalling that $a(z)$ is not a Virasoro primary for $\alpha$ finite; instead we have
\begin{equation}
T^{\textup{bos.}}(z)a(w)\sim\frac{2/\alpha\:\wun}{(z-w)^3}+\frac{a(w)}{(z-w)^2}+\frac{\pd a(w)}{z-w}.
\end{equation}
The Fourier expansion $T^{\textup{bos.}}(z) = \sum_{n \in \mathbb{Z}} L_n^{\textup{bos.}} z^{-n-2}$ defines the Virasoro modes and it is easy to check that the Lie algebra $\alg{g}$ spanned by the $a_n$, $L_n^{\text{bos.}}$ and $\wun$ is conformally graded. It is likewise easy to check that (for finite $\alpha$) the only adjoint on the Heisenberg algebra consistent with the standard Virasoro adjoint $(L_n^{\textup{bos.}})^{\dag} = L_{-n}^{\textup{bos.}}$ is
\begin{equation} \label{eq:HeisAdj}
a_n^{\dag} = -a_{-n} - \frac{2}{\alpha} \delta_{n,0} \wun.
\end{equation}
Since $\alg{g}_0 = \vspn \set{a_0, L_0, \wun}$ is abelian, it coincides with the Cartan subalgebra $\alg{h}$, hence relaxed \hw{}{} theory reduces to ordinary highest weight{} theory for the Heisenberg algebra. We remark that this is one example where we cannot insist that the Cartan subalgebra consist of self-adjoint elements.
The highest weight{} theory of the Heisenberg Lie algebra is well known. The Verma modules are known as Fock spaces and are parametrised by the $a_0$-eigenvalue $p$ of the \hw{} vector{}. They are simple for all \(p\in\mathbb{C}\). We will denote the Fock spaces by \(\FF{p}\) and their corresponding \hw{} vectors{} by \(\hv{p}\). In accordance with \eqref{eq:HeisAdj}, the module conjugate to the Fock space $\FF{p}$ is $\FF{-p-2/\alpha}$.
\subsection*{The ghost algebra $\GA$}
For the $\beta \gamma$ ghost system, we follow the notations and conventions of \cite{RidBos14}. The ghost vertex operator algebra{} is generated by two bosonic fields, \(\beta(z)\) and \(\gamma(z)\), whose operator product expansions{} are
\begin{equation} \label{eq:GhOPE}
\beta(z)\beta(w)\sim 0, \qquad
\gamma(z)\beta(w)\sim\frac{\wun}{z-w}, \qquad
\gamma(z)\gamma(w)\sim 0.
\end{equation}
From these, one constructs a Heisenberg field $J(z)$ and an energy-momentum tensor $T^{\textup{gh.}}(z)$ by
\begin{equation}
J(z) = \normord{\beta(z)\gamma(z)}, \qquad
T^{\textup{gh.}}(z) = -\normord{\beta(z)\pd\gamma(z)}.
\end{equation}
These give $\beta(z)$ and $\gamma(z)$ Heisenberg weights $+1$ and $-1$, and conformal weights $1$ and $0$, respectively. We remark that $J(z)$ is not normalised as in \eqref{eq:HeisOPE}, nor is it primary with respect to $T^{\textup{gh.}}(z)$:
\begin{equation}
J(z)J(w)\sim\frac{-\wun}{(z-w)^2}, \qquad
T^{\textup{gh.}}(z)J(w)\sim\frac{-\wun}{(z-w)^3}+\frac{J(w)}{(z-w)^2}+\frac{\pd J(w)}{z-w}.
\end{equation}
As with the free boson, there is actually a one-parameter family of conformal structures; $T^{\textup{gh.}}(z)$ has been chosen so that the ghost fields $\beta(z)$ and $\gamma(z)$ have the required conformal weights. This choice also fixes the central charge of the ghost system to be $c^{\textup{gh.}}=2$.
The Fourier expansions $\beta(z)=\sum_{n\in\mathbb{Z}}\beta_n z^{-n-1}$ and $\gamma(z)=\sum_{n\in\mathbb{Z}}\gamma_nz^{-n}$ now yield the commutation relations
\begin{equation} \label{eq:BosGhComm}
\comm{\beta_m}{\beta_n}=0,\quad
\comm{\gamma_m}{\beta_n}=\delta_{m+n,0}\wun,\quad
\comm{\gamma_m}{\gamma_n}=0;\qquad
m,n\in\mathbb{Z}.
\end{equation}
The infinite-dimensional Lie algebra $\GA$, spanned by the $\beta_n$, $\gamma_n$ and $\wun$, is called the ghost Lie algebra. Again, \(\wun\) is identified with the unit of $\UEA \GA$ and we assume that it acts as the identity on all \(\GA\)-modules. As with the Heisenberg algebra, we will extend this algebra by the modes $J_n, L_n^{\textup{gh.}} \in \UEA \GA$, defined by $J(z) = \sum_{n \in \mathbb{Z}} J_n z^{-n-1}$ and $T^{\textup{gh.}}(z) = \sum_{n \in \mathbb{Z}} L_n^{\textup{gh.}} z^{-n-2}$. The Lie algebra $\alg{g}$ spanned by the $\beta_n$, $\gamma_n$, $J_n$, $L_n^{\textup{gh.}}$ and $\wun$ is then conformally graded with relations including
\begin{subequations}
\begin{gather}
\comm{J_m}{\beta_n} = \beta_{m+n}, \qquad
\comm{J_m}{\gamma_n} = -\gamma_{m+n}, \qquad
\comm{L_m^{\textup{gh.}}}{\beta_n} = -n \beta_{m+n}, \qquad
\comm{L_m^{\textup{gh.}}}{\gamma_n} = -(m+n) \gamma_{m+n}, \\
\comm{J_m}{J_n} = -m \delta_{m+n,0} \wun, \qquad
\comm{L_m^{\textup{gh.}}}{J_n} = - n J_{m+n} - \frac{1}{2} m(m+1) \delta_{m+n,0} \wun.
\end{gather}
\end{subequations}
The Cartan subalgebra $\alg{h} = \vspn \tset{J_0, L_0^{\textup{gh.}}, \wun}$ is a proper subalgebra of $\alg{g}_0 = \vspn \tset{\beta_0, \gamma_0, J_0, L_0^{\textup{gh.}}, \wun}$, so the relaxed and ordinary highest weight{} theories of the $\beta \gamma$ ghost system do not coincide. In this case, the ghost adjoint
\begin{equation} \label{eq:GhAdj}
\beta_n^{\dag} = \gamma_{-n}
\end{equation}
implies that all the elements of $\alg{h}$ are self-adjoint.
We start with the ordinary highest weight{} theory corresponding to the triangular decomposition in which $\beta_0 \in \alg{g}_+$ annihilates the \hw{} vector{} $\gvac$ and $\gamma_0 \in \alg{g}_-$ need not. It turns out that this yields a unique Verma module because these conditions imply that $J_0 \gvac = L_0^{\textup{gh.}} \gvac = 0$ (and $\wun$ acts as the identity, as always). This module is simple; in fact, we may take it to be the ghost vacuum module $\GhostVac$ because these conditions also imply that $L_{-1}^{\textup{gh.}} \gvac = 0$. Of course, one can also take the triangular decomposition in which $\gamma_0$ annihilates the \hw{} vector{} and $\beta_0$ need not.\footnote{There are other triangular decompositions, but they will not concern us here. Indeed, they may be obtained from those already mentioned by twisting with a so-called spectral flow automorphism, see \cite{RidBos14}.} The resulting Verma module is the simple module $\ghconjmod{\GhostVac}$ conjugate to $\GhostVac$, obtained by twisting the action of $\GA$ by $\ghconjaut$, the (order $4$) conjugation automorphism that sends $\beta_n$ to $\gamma_n$ and $\gamma_n$ to $-\beta_n$.
The relaxed \hw{}{} theory of the ghost system is significantly more interesting as we no longer require that $\beta_0$ (or $\gamma_0$) annihilates the \hw{} vector{}. In fact, we may induce from a fairly arbitrary simple weight module of $\alg{g}_0$. Given that $J_0$ and $L_0^{\textup{gh.}}$, as abstract elements of $\alg{g}_0$, are going to be identified with elements of $\UEA \GA$ and that the $\alg{g}_0$-module weight vectors are going to be identified with the relaxed \hwvs{}{} of the induced module, we may restrict to $\alg{g}_0$-modules on which the action of $J_0$ is identified with that of $\gamma_0 \beta_0$ and the action of $L_0^{\textup{gh.}}$ is always $0$.
\begin{cprop}[\protect{\cite[Prop.~1]{RidBos14}}] \label{prop:FinGhReps}
A simple weight module over $\alg{g}_0$ upon which $J_0 = \gamma_0 \beta_0$
and $L_0^{\textup{gh.}} = 0$ is isomorphic to one of the following:
\begin{itemize}
\item The module $\FinGhostVac$ generated by a vector \(\gv{}\) which is annihilated by \(\beta_0\) and thus also \(J_0\). This module has a basis of weight vectors $\gv{j}$, $j \in \mathbb{Z}_{\le 0}$, where $J_0 \gv{j} = j \gv{j}$.
\item The module $\finghconjmod{\FinGhostVac}$, conjugate to $\FinGhostVac$, generated by a vector \(\gv{}\) which is annihilated by \(\gamma_0\) and thus \(J_0\gv{}=\gv{}\). This module has a basis of weight vectors $\gv{j}$, $j \in \mathbb{Z}_{\ge 1}$, where $J_0 \gv{j} = j \gv{j}$.
\item The modules $\FinGhostRel{q}$, where $q\in\mathbb{C}\setminus\mathbb{Z}$, each of which is generated by a vector \(\gv{}\) satisfying \(J_0\gv{}=q\gv{}\); it follows that no non-zero vector is annihilated by $\beta_0$ or $\gamma_0$. The eigenvalues of $J_0 = \gamma_0 \beta_0$ all lie in $q + \mathbb{Z}$ and these modules have a basis of weight vectors $\gv{j}$, $j \in q + \mathbb{Z}$, satisfying $J_0 \gv{j} = j \gv{j}$. The modules $\FinGhostRel{q}$ and $\FinGhostRel{q+1}$ are isomorphic.
\end{itemize}
\end{cprop}
Additionally, one can consider the indecomposable $\alg{g}_0$-modules $\FinGhostRel{0}^+$ and $\FinGhostRel{0}^-$ that likewise have a basis of weight vectors $\gv{j}$, $j \in \mathbb{Z}$, satisfying $J_0 \gv{j} = j \gv{j}$, but they are not simple. They are determined (up to isomorphism) by the following non-split short exact sequences:
\begin{equation}
\dses{\FinGhostVac}{}{\FinGhostRel{0}^+}{}{\finghconjmod{\FinGhostVac}}, \qquad
\dses{\finghconjmod{\FinGhostVac}}{}{\FinGhostRel{0}^-}{}{\FinGhostVac}.
\end{equation}
Inducing the $\alg{g}_0$-modules $\FinGhostVac$, $\finghconjmod{\FinGhostVac}$ and $\FinGhostRel{q}$ ($q \notin \mathbb{Z}$) therefore results in relaxed Verma modules for $\GA$. The first two give the simple ghost vacuum module $\GhostVac$ and its conjugate $\ghconjmod{\GhostVac}$, respectively; the last gives new modules $\GhostRel{q}$ which are also simple and satisfy $\GhostRel{q} \cong \GhostRel{q+1}$. We may similarly induce the non-simple $\alg{g}_0$-modules $\FinGhostRel{0}^+$ and $\FinGhostRel{0}^-$ to obtain non-simple $\GA$-modules $\FinGhostRel{0}^+$ and $\FinGhostRel{0}^-$ which are likewise determined (up to isomorphism) by the following non-split short exact sequences:
\begin{equation}
\dses{\GhostVac}{}{\GhostRel{0}^+}{}{\ghconjmod{\GhostVac}}, \qquad
\dses{\ghconjmod{\GhostVac}}{}{\GhostRel{0}^-}{}{\GhostVac}.
\end{equation}
\subsection*{The affine Kac-Moody algebra $\AKMA{sl}{2}$}
We consider the universal vertex operator algebra{} of non-critical level $k \neq -2$ generated by three fields $e(z)$, $h(z)$ and $f(z)$ satisfying the operator product expansions{}
\begin{equation} \label{eq:SL2OPE}
\begin{aligned}
\func{h}{z} \func{e}{w} &\sim \frac{+2 \func{e}{w}}{z-w} \vphantom{\frac{2k}{\brac{z-w}^2}}, \\
\func{h}{z} \func{f}{w} &\sim \frac{-2 \func{f}{w}}{z-w} \vphantom{\frac{-k}{\brac{z-w}^2}},
\end{aligned}
\qquad
\begin{aligned}
\func{h}{z} \func{h}{w} &\sim \frac{2k\:\wun}{\brac{z-w}^2}, \\
\func{e}{z} \func{f}{w} &\sim \frac{-k\:\wun}{\brac{z-w}^2} - \frac{\func{h}{w}}{z-w},
\end{aligned}
\qquad
\begin{aligned}
\func{e}{z} \func{e}{w} &\sim 0 \vphantom{\frac{2k}{\brac{z-w}^2}}, \\
\func{f}{z} \func{f}{w} &\sim 0 \vphantom{\frac{-k}{\brac{z-w}^2}},
\end{aligned}
\end{equation}
and no other (independent) relations. We denote this vertex operator algebra{} by \(\UVOA{k}\). The maximal proper ideal of \(\UVOA{k}\) is non-trivial if and only if \(k\) is admissible. Moreover, this ideal is generated by a single primary field (singular vector) \cite{KacStr79}; we do not set this primary field to zero. Note that we have chosen the $\SLA{sl}{2}$ basis $\set{e,h,f}$ to be consistent with previous work, \cite{RidSL208} in particular, where it was necessary to tailor the basis to the $\SLA{sl}{2;\mathbb{R}}$ adjoint rather than the (more traditional) $\SLA{su}{2}$ adjoint. This is reflected in the signs appearing in the formulae for $e(z)f(w)$ above.
With the usual Fourier expansions $g(z) = \sum_{n \in \mathbb{Z}} g_n z^{-n-1}$,
where $g=e,h,f$, the commutation relations are
\begin{equation} \label{eq:CommSL2}
\begin{aligned}
\comm{h_m}{e_n} &= +2 e_{m+n}, \\
\comm{h_m}{f_n} &= -2 f_{m+n},
\end{aligned}
\quad
\begin{aligned}
\comm{h_m}{h_n} &= 2m \delta_{m+n,0} k \: \wun, \\
\comm{e_m}{f_n} &= -h_{m+n} - m \delta_{m+n,0} k \: \wun,
\end{aligned}
\quad
\begin{aligned}
\comm{e_m}{e_n} &= 0, \\
\comm{f_m}{f_n} &= 0,
\end{aligned}
\qquad m,n \in \mathbb{Z}
\end{equation}
and these make $\vspn \set{e_n, h_n, f_n, \wun}$ into a Lie algebra which we denote by $\AKMA{sl}{2}$. Once again, we assume that the unit $\wun \in \UEA \tbrac{\AKMA{sl}{2}}$ acts as the identity on each $\AKMA{sl}{2}$-module.
The standard conformal structure of \(\UVOA{k}\)
is uniquely determined by requiring that $e(z)$, $h(z)$ and $f(z)$ are Virasoro primaries of conformal weight $1$ (this structure exists for all $k \neq -2$). The Sugawara construction then gives the explicit form of the energy-momentum tensor as
\begin{equation} \label{eq:DefT}
\func{T}{z} = \frac{1}{2t} \brac{\frac{1}{2} \normord{\func{h}{z} \func{h}{z}} - \normord{\func{e}{z} \func{f}{z}} - \normord{\func{f}{z} \func{e}{z}}},
\end{equation}
where $t = k+2$. With $\tfunc{T}{z} = \sum_{n \in \mathbb{Z}} L_n z^{-n-2}$, one finds that the modes $L_n$ generate a copy of the Virasoro algebra with central charge $c = 3 - 6/t$. The Lie algebra $\alg{g}$ spanned by the $e_n$, $h_n$, $f_n$, $L_n$ and $\wun$ is then conformally graded with Cartan subalgebra $\alg{h} = \vspn \set{h_0, L_0, \wun}$. We have chosen the $\SLA{sl}{2}$ basis so that the $\SLA{sl}{2;\mathbb{R}}$ adjoint becomes
\begin{equation}
e_n^{\dag} = f_{-n}, \qquad h_n^{\dag} = h_{-n}.
\end{equation}
Again, the Cartan subalgebra consists of self-adjoint elements.
The highest weight{} theory of $\AKMA{sl}{2}$ is well known. The standard triangular decomposition splits the zero modes so that $e_0 \in \alg{g}_+$ annihilates \hw{} vectors{} but $f_0 \in \alg{g}_-$ need not. Then, the Verma modules $\Ver{\lambda}$ are parametrised by the $\SLA{sl}{2}$-weight ($h_0$-eigenvalue) $\lambda \in \mathbb{C}$ of the \hw{} vector{} because it follows from \eqref{eq:DefT} that its conformal weight is then given by
\begin{equation} \label{eq:ConfDimHWV}
\Delta_{\lambda} = \frac{\lambda \brac{\lambda + 2}}{4t}.
\end{equation}
These Verma modules need not be simple. The quotient module $\Ver{0} / \Ver{-2}$, where the submodule $\Ver{-2}$ is generated by acting with $f_0$ on the (generating) \hw{} vector{} of $\Ver{0}$, is the vacuum module; it carries the structure of the universal vertex operator algebra{} \(\UVOA{k}\) defined by \eqref{eq:SL2OPE}. As with the ghosts, one can also consider the triangular decomposition in which $f_0$ annihilates \hw{} vectors{} but $e_0$ need not.\footnote{And as with the ghosts, there are again other triangular decompositions that will not concern us here, being related to those discussed here by spectral flow automorphisms, see \cite{RidSL208,CreMod13}.} The Verma modules with respect to this decomposition are then the conjugates $\conjmod{\Ver{\lambda}}$ of the Verma modules $\Ver{\lambda}$, where the (order $2$) conjugation automorphism $\conjaut$ sends $e_n$ to $-f_n$ and $h_n$ to $-h_n$.
Because $\alg{g}_0 = \vspn \set{e_0, h_0, f_0, L_0, \wun}$ is non-abelian, it strictly contains $\alg{h}$, so the relaxed \hw{}{} theory of $\AKMA{sl}{2}$ is strictly more general. We note that inducing from a $\alg{g}_0$-module reduces to choosing an $\SLA{sl}{2}$-module because we require that $4t L_0$ acts as $h_0^2 - 2 e_0 f_0 - 2 f_0 e_0$ (this is how $4t L_0$ acts on relaxed \hwvs{}{}). The analogue of \propref{prop:FinGhReps} is then the classification of weight modules for $\SLA{sl}{2}$ (see \cite{MazLec10}, for example):
\begin{prop} \label{prop:FinSL2WtMods}
The simple weight modules of $\SLA{sl}{2}$ are exhausted by the following:
\begin{itemize}
\item The $\brac{\lambda + 1}$-dimensional modules $\FinIrr{\lambda}$, with $\lambda \in \mathbb{Z}_{\ge 0}$. The module $\FinIrr{\lambda}$ has a basis of weight vectors $\av{m}$, where $m = \lambda, \lambda -2 , \ldots, -\lambda$ and $h_0 \av{m} = m \av{m}$. It is both highest and lowest weight.
\item The infinite-dimensional \hw{} modules{} $\FinDisc{\lambda}$, with $\lambda \in\mathbb{C}\setminus \mathbb{Z}_{\ge 0}$. The module $\FinDisc{\lambda}$ has a basis of weight vectors $\av{m}$, where $m = \lambda, \lambda -2, \lambda - 4, \ldots$ and $h_0 \av{m} = m \av{m}$.
\item The infinite-dimensional lowest weight modules $\finconjmod{\FinDisc{-\lambda}}$, with $\lambda \in\mathbb{C}\setminus \mathbb{Z}_{\le 0}$. Here, $\finconjaut$ is the Weyl reflection of $\SLA{sl}{2}$ that sends $e_0$ to $-f_0$ and $h_0$ to $-h_0$. The module $\finconjmod{\FinDisc{-\lambda}}$ has a basis of weight vectors $\av{m}$, where $m = \lambda, \lambda + 2, \lambda + 4, \ldots$ and $h_0 \av{m} = m \av{m}$.
\item The infinite-dimensional weight modules $\FinRel{\lambda; \Delta}$, with $\lambda, \Delta \in \mathbb{C}$ satisfying $4t \Delta \neq \mu \brac{\mu+2}$ for any $\mu \in \lambda + 2 \mathbb{Z}$. The module $\FinRel{\lambda; \Delta}$ has a basis of weight vectors $\av{\mu}$, with $\mu \in \lambda + 2 \mathbb{Z}$, satisfying $h_0 \av{\mu} = \mu \av{\mu}$. It is neither highest nor lowest weight. Moreover, there are isomorphisms $\FinRel{\lambda; \Delta} \cong \FinRel{\lambda + 2; \Delta}$.
\end{itemize}
\end{prop}
\noindent As with \propref{prop:FinGhReps}, there are non-simple analogues of the $\FinRel{\lambda; \Delta}$ when $4t \Delta = \mu \brac{\mu+2}$ for some $\mu \in \lambda + 2 \mathbb{Z}$. The structures of these indecomposables depend upon precisely how many $\mu \in \lambda + 2 \mathbb{Z}$ satisfy this constraint \cite{MazLec10} and we shall defer their consideration until they are needed (\secref{sec:HWRelAdmMod}).
Inducing each of the simple $\SLA{sl}{2}$-modules now yields relaxed Verma modules for $\AKMA{sl}{2}$. More precisely:
\begin{itemize}
\item Inducing $\FinIrr{\lambda}$, where $\lambda \in \mathbb{Z}_{\ge 0}$, results in the \hw{} module{} $\Ver{\lambda} / \Ver{-\lambda - 2}$, where the \hw{} vectors{} of $\Ver{\lambda}$ and $\Ver{-\lambda - 2}$ are related by $\av{-\lambda - 2} = f_0^{\lambda + 1} \av{\lambda}$. This induced module need not be simple; its simple quotient will be denoted by $\Irr{\lambda}$. These simple modules are self-conjugate: $\conjmod{\Irr{\lambda}} = \Irr{\lambda}$.
\item Inducing $\FinDisc{\lambda}$, where $\lambda \in \mathbb{C} \setminus \mathbb{Z}_{\ge 0}$, results in the Verma module $\Ver{\lambda}$; the simple quotient will be denoted by $\Disc{\lambda}$.
\item Inducing $\finconjmod{\FinDisc{-\lambda}}$, where $\lambda \in \mathbb{C} \setminus \mathbb{Z}_{\le 0}$, results in $\conjmod{\Ver{-\lambda}}$; the simple quotient is $\conjmod{\Disc{-\lambda}}$.
\item Inducing $\FinRel{\lambda; \Delta}$, with $\lambda, \Delta \in \mathbb{C}$ satisfying $4t \Delta \neq \mu \brac{\mu+2}$ for any $\mu \in \lambda + 2 \mathbb{Z}$, results in a new relaxed Verma module that we shall denote by $\RelVer{\lambda; \Delta}$; its simple quotient will be denoted by $\Rel{\lambda; \Delta}$. As above, there are isomorphisms $\RelVer{\lambda; \Delta} \cong \RelVer{\lambda + 2; \Delta}$ and $\Rel{\lambda; \Delta} \cong \Rel{\lambda + 2; \Delta}$. Finally, the module conjugate to $\Rel{\lambda; \Delta}$ is $\Rel{-\lambda; \Delta}$.
\end{itemize}
We emphasise that whereas the simple \hw{} modules{} (and their conjugates) are characterised by a single parameter, the highest weight{}, the $\Rel{\lambda; \Delta}$ require two parameters in general. The three classes of simple $\AKMA{sl}{2}$-modules $\Irr{\lambda}$, $\Disc{\lambda}$ and $\conjmod{\Disc{-\lambda}}$, and $\Rel{\lambda; \Delta}$ are distinguished by their relaxed \hwvs{}{}: $\Irr{\lambda}$ has finitely many ($\lambda + 1$ in fact); $\Disc{\lambda}$ and $\conjmod{\Disc{-\lambda}}$ have infinitely many, but their $\SLA{sl}{2}$-weights are bounded above and below, respectively; $\Rel{\lambda; \Delta}$ has infinitely many with no bound on the $\SLA{sl}{2}$-weights.
\section{The Wakimoto Free Field Realisation} \label{sec:Wak}
A free field realisation of $\AKMA{sl}{2}$ for any level was constructed by Wakimoto in \cite{WakFoc86}. It shows that fields $\tfunc{e}{z}$, $\tfunc{h}{z}$ and $\tfunc{f}{z}$, satisfying the operator product expansions{} \eqref{eq:SL2OPE}, may be constructed in terms of a free boson and a pair of bosonic ghosts. We review this and the screening operator formalism of free field theories, concluding by deriving explicit formulae, in terms of Jack symmetric polynomials, for certain (relaxed) singular vectors in (relaxed) Wakimoto modules over $\AKMA{sl}{2}$. A corollary of this analysis is that for admissible levels, Wakimoto's construction describes the universal vertex operator algebra{} $\UVOA{k}$ of $\AKMA{sl}{2}$, rather than its simple quotient $\AdmMod{u}{v}$.
\subsection{The Wakimoto construction}
There is a one-parameter family of realisations of \(\AKMA{sl}{2}\) in terms of tensor products of the free boson and ghost fields. Given the operator product expansions{} \eqref{eq:HeisOPE} and \eqref{eq:GhOPE}, it is straightforward to verify that defining (we omit the tensor products for notational simplicity)
\begin{equation}\label{eq:ffrealisation}
\begin{gathered}
e(z)=\beta(z),\qquad
h(z)=2\normord{\beta(z)\gamma(z)}+\alpha\:a(z), \\
f(z)=\normord{\beta(z)\gamma(z)\gamma(z)}+\alpha\:a(z)\gamma(z)+\Bigl(\frac{\alpha^2}{2}-2\Bigr)\pd\gamma(z)
\end{gathered}
\end{equation}
reproduces the $\AKMA{sl}{2}$ operator product expansions{} \eqref{eq:SL2OPE} with the level $k = t-2$ being related to the parameter \(\alpha\) by $\alpha^2 = 2t$. Moreover, the $\AKMA{sl}{2}$ energy-momentum tensor \eqref{eq:DefT} and central charge then decompose as
\begin{equation} \label{eq:WakT}
\begin{split}
T(z)&=T^{\text{bos.}}(z) + T^{\text{gh.}}(z) = \frac{1}{2}\normord{a(z)a(z)}-\frac{1}{\alpha}\pd
a(z)-\normord{\beta(z)\pd\gamma(z)},\\
c&=c^{\text{bos.}} + c^{\text{gh.}} = 1-\frac{12}{\alpha^2} + 2=3-\frac{6}{t},
\end{split}
\end{equation}
identifying $\alpha$ with the deformation parameter in the free boson conformal structure \eqref{eq:HeisT}.
\begin{defn}
We define the \emph{Wakimoto vertex operator algebra{}} to be the tensor product of the Heisenberg and ghost vertex operator algebras{}, equipped with the conformal structure given in \eqref{eq:WakT}.
\end{defn}
\noindent Recall that when $k$ is admissible, the universal $\AKMA{sl}{2}$ vertex operator algebra{} is not simple, but has a unique maximal ideal that is generated by a single primary field (singular vector). One of the results of this section (\corref{cor:WakVacIsUniversal}) is that this field is non-zero in the Wakimoto vertex operator algebra{}.
The Wakimoto free field realisation endows the tensor product of a Heisenberg Fock space and a ghost module with the structure of an \(\AKMA{sl}{2}\)-module, by restriction. We refer to such modules as Wakimoto modules, distinguishing at least four types:
\begin{itemize}
\item The highest weight{} Wakimoto modules $\Wak{p} = \FF{p} \otimes \GhostVac$, recalling that $\GhostVac$ denotes the ghost vacuum module.
\item The conjugate highest weight{} Wakimoto modules $\ghconjmod{\Wak{p}} = \FF{p} \otimes \ghconjmod{\GhostVac}$, obtained by conjugating the ghost module.
\item The relaxed \hw{}{} Wakimoto modules $\WakRel{p; q} = \FF{p} \otimes \GhostRel{q}$, for $q \notin \mathbb{Z}$.
\item The relaxed \hw{}{} Wakimoto modules $\WakRel{p; 0}^{\pm} = \FF{p} \otimes \GhostRel{0}^{\pm}$.
\end{itemize}
The structure of the highest weight{} Wakimoto modules $\Wak{p}$ was determined in \cite{BerFoc90}. To the best of our knowledge, the relaxed modules $\WakRel{p;q}$ have not previously been considered in the literature.
Consider now the tensor product of a Heisenberg \hw{} vector{} $\hv{p}$ and the ghost vacuum $\gvac$, denoting it by $\wv{p} = \hv{p} \otimes \gvac$ for convenience. The free field realisations \eqref{eq:ffrealisation} of the $\AKMA{sl}{2}$ fields imply that $\wv{p}$ is an $\AKMA{sl}{2}$ \hw{} vector{} of $\SLA{sl}{2}$- and conformal weight
\begin{equation}
\lambda_p = \alpha p, \qquad \Delta_p = \frac{1}{2} p \brac{p + \frac{2}{\alpha}} = \frac{\lambda_p \brac{\lambda_p + 2}}{4t},
\end{equation}
where we recall that $\alpha^2 = 2t$. Similarly, the tensor product $\wv{p;q} = \hv{p} \otimes \gv{q}$ of $\hv{p}$ with a relaxed \hwv{}{} $\gv{q}$ for the ghosts is an $\AKMA{sl}{2}$ relaxed \hwv{}{} of $\SLA{sl}{2}$- and conformal weight
\begin{equation} \label{eq:RelaxedWts}
\lambda_{p;q} = \alpha p + 2q, \qquad \Delta_p = \frac{1}{2} p \brac{p + \frac{2}{\alpha}} = \frac{\brac{\lambda_{p;q} - 2q} \brac{\lambda_{p;q} - 2q + 2}}{4t}.
\end{equation}
This shows that all the simple $\AKMA{sl}{2}$-modules, highest weight{} and relaxed, may be realised as subquotients of Wakimoto modules.
\subsection{Screening fields and operators}
We begin by recalling the construction of vertex operators for the free boson. For this, one extends the Heisenberg algebra by introducing a generator \(\hat a\) satisfying
\begin{equation}
\comm{a_m}{\hat a}=\delta_{m,0}\wun.
\end{equation}
The vertex operators $\vop{p}(z)$, parametrised by $p \in \mathbb{C}$, are then defined by
\begin{equation}\label{eq:vertop}
\vop{p}(z)=\ee^{p\hat a}z^{p a_0}
\prod_{m\ge 1}\exp\brac{p\frac{\alpha_{-m}}{m}z^m}
\exp\brac{-p\frac{\alpha_{m}}{m}z^{-m}}.
\end{equation}
A standard computation shows that the vertex operators are free boson primaries of Heisenberg weight $p$ and conformal weight $\frac{1}{2} p \brac{p + \frac{2}{\alpha}}$, by virtue of the operator product expansions{}
\begin{equation}
a(z)\vop{p}(w)\sim\frac{p \vop{p}(w)}{z-w}, \qquad
T^{\text{bos.}}(z)\vop{p}(w)\sim \frac{\frac{1}{2} p \big( p + \frac{2}{\alpha} \bigr) \vop{p}(w)}{(z-w)^2} + \frac{\pd\vop{p}(w)}{z-w}.
\end{equation}
For later use, we record that the composition of \(k\) vertex operators is given by
\begin{equation} \label{eq:CompVerOps}
\vop{p_1}(z_1)\cdots \vop{p_k}(z_k)=
\prod_{i<j}(z_i-z_j)^{p_ip_j}\cdot
\ee^{\sum_{i=1}^kp_i\hat a}
\prod_{i=1}^k z_i^{p_i a_0}\cdot
\prod_{m\ge1}\exp\left(\frac{a_{-m}}{m}\sum_{i=1}^k p_i z_i^m\right)
\exp\left(-\frac{a_{m}}{m}\sum_{i=1}^kp_i z_{i}^{-m}\right).
\end{equation}
We now define the notion of a screening field.
\begin{defn}
A \emph{screening field} $\scr(w)$ for a vertex operator algebra{} $\VOA{V}$ is a field, generally not belonging to $\VOA{V}$ itself, which has the property that the singular terms of the operator product expansions{} of each of the generating fields of $\VOA{V}$ with $\scr(w)$ are total derivatives in $w$.
\end{defn}
\noindent Our definition of a \emph{screening operator}, in a moral sense at least, is then a vertex operator algebra{} module homomorphism that can be constructed from screening fields. This means that a screening operator commutes with the fields, and hence the mode algebra, of the vertex operator algebra{}. The standard way of constructing screening operators is as the residues (zero modes) of screening fields; these are guaranteed to commute with all the vertex operator algebra{} fields, provided that the residues are well defined. Their chief application stems from the fact that they map (relaxed) \hw{} vectors{} to (relaxed) \hw{} vectors{} and thereby explicitly construct (relaxed) singular vectors. We remark that if the screening field is a Virasoro primary (excluding the identity field), then its operator product expansion{} with the energy-momentum tensor forces its conformal weight to be $1$.
The operator product expansions{} of the free field $\AKMA{sl}{2}$ fields \eqref{eq:ffrealisation} with a free boson vertex operator are easily computed to be
\begin{equation}
e(z)\vop{p}(w)\sim 0, \qquad
h(z)\vop{p}(w)\sim\frac{\alpha p\vop{p}(w)}{z-w},\qquad
f(z)\vop{p}(w)\sim\frac{\alpha p\vop{p}(w)\gamma(w)}{z-w}.
\end{equation}
For a vertex operator to be a non-trivial screening field, its Heisenberg weight would have to satisfy $\frac{1}{2} p \tbrac{p + \frac{2}{\alpha}} = 1$. However, the above operator product expansions{} show that these vertex operators are not screening fields for $\AKMA{sl}{2}$. However, the field \(\scr(z)=\vop{-2/\alpha}(z)\beta(z)\) is a screening field \cite{BerFoc90}:
\begin{equation}\label{eq:scrope}
e(z)\scr(w)\sim 0, \qquad
h(z)\scr(w)\sim 0, \qquad
f(z)\scr(w)\sim-t\pd_w\frac{V_{-2/\alpha}(w)}{z-w}.
\end{equation}
It follows that the zero mode
\begin{equation}
\scrs{1} = \oint_0 \scr(z) \: \frac{\dd z}{2 \pi \ii} = \oint_0 \vop{-2/\alpha}(z) \beta(z) \: \frac{\dd z}{2 \pi \ii}
\end{equation}
is a screening operator, whenever the contour around $0$ actually closes. This will be the case when $\scrs{1}$ acts on a state for which the relevant Fourier expansion of $\scr(z)$ has only integer powers of $z$. Equivalently, $\scrs{1}$ has a well defined action on a given state if and only if the operator product expansion{} of $\scr(z)$ with the corresponding field is a Laurent series. For example, $\scrs{1}$ only acts on $\wv{p} = \hv{p} \otimes \gvac$, the tensor product of a Heisenberg vacuum $\hv{p}$ with the ghost vacuum $\gvac$, when $p=\frac{1}{2} m \alpha$, $m \in \mathbb{Z}$, because
\begin{equation}
\scr(z) \vop{p}(w) = \vop{-2/\alpha}(z) \vop{p}(w) \beta(z) = \frac{\vop{p-2/\alpha}(w) \beta(w)}{(z-w)^{2p/\alpha}} + \cdots.
\end{equation}
This shows that the screening field $\scr(z)$ only defines module homomorphisms, hence constructs singular vectors, for certain vertex operator algebra{} modules.
To construct more module homomorphisms, it is natural to consider products of screening fields. To check that such products also yield screening operators, it is convenient to use the language of differential forms. Suppose then that $j(z)$ is a vertex operator algebra{} field and that $\scr_i(w_i)$ is a screening field, so that
\begin{equation}
\comm{j_n}{\scr_i(w_i) \: \dd w_i} = \pd_{w_i} \antiscr_i(w_i) \: \dd w_i = \dd \antiscr_i(w_i),
\end{equation}
for some ($n$-dependent) $\antiscr_i(w_i)$. Then,
\begin{align}
\comm{j_n}{\scr_1(w_1) \: \dd w_1 \wedge \scr_2(w_2) \: \dd w_2} &= \dd \antiscr_1(w_1) \wedge \scr_2(w_2) \: \dd w_2 + \scr_1(w_1) \: \dd w_1 \wedge \dd \antiscr_2(w_2) \notag \\
&= \dd \tbrac{\antiscr_1(w_1) \: \scr_2(w_2) \: \dd w_2 - \scr_1(w_1) \: \dd w_1 \: \antiscr_2(w_2)}
\end{align}
is exact, hence the commutator vanishes upon integrating over a closed cycle. The generalisation to more than two screening fields is immediate.
We therefore ask the question of when there exists a closed cycle over which some given product of the screening fields $\scr(z) = \vop{-2/\alpha}(z) \beta(z)$ can be integrated to define \(\AKMA{sl}{2}\)-homomorphisms. We introduce the shorthand
\begin{equation}
\scrs{r}(z)=\scrs{r}(z_1,\dots,z_k)=\scr(z_1)\cdots\scr(z_r)
\end{equation}
and consider the action of this product of screening fields on the Wakimoto module \(\Wak{p}\) whose Heisenberg highest weight{} is $p$ . Using \eqref{eq:CompVerOps}, this action can be written in the form
\begin{multline}\label{eq:scrprod}
\Bigl.\scrs{r}(z)\Bigr\rvert_{\Wak{p}}=\Bigl.\prod_{1\le i< j\le r}(z_i-z_j)^{4/\alpha^2}\cdot
\prod_{i=1}^r z_i^{-2p/\alpha}\\
\cdot\prod_{m\ge 1}\exp\brac{-\frac{2}{\alpha}\fpowsum{m}{z}\frac{a_{-m}}{m}}
\exp\brac{\frac{2}{\alpha}\overline{\fpowsum{m}{z}}\frac{a_{m}}{m}}\cdot
\prod_{i=1}^r\beta(z_i)\Bigr\rvert_{\Wak{p-2r/\alpha}},
\end{multline}
where \(\fpowsum{m}{z}\) denotes a power sum and the overline \(\overline{\fpowsum{m}{z}}=\fpowsum{m}{z_1^{-1},\dots,z_r^{-1}}\) denotes variable inversion (see \appref{sec:SymmPoly} for our conventions for power sums and other symmetric polynomials). The action of $\scrs{r}(z)$ on the other Wakimoto modules with Heisenberg highest weight{} $p$, for example the \rhw{} module{} $\WakRel{p;q}$, is identical --- the ghost weight $q$ is not changed.
Up to an unimportant phase factor, which we suppress, the first two factors on the right-hand side{} of \eqref{eq:scrprod} are
\begin{equation} \label{eq:scrprod'}
\prod_{1\le i< j\le r}(z_i-z_j)^{4/\alpha^2}\cdot\prod_{i=1}^r z_i^{-2p/\alpha}
=\prod_{1\le i\neq j\le r}\Bigl(1-\frac{z_i}{z_j}\Bigr)^{1/t}\cdot
\prod_{i=1}^rz_i^{(r-1)2/\alpha^2-2p/\alpha},
\end{equation}
where we recall that $\alpha^2=2t$. The second factor on the right-hand side{} of this expression therefore isolates all the (potential) non-integer powers of the $z_i$ in \eqref{eq:scrprod}, hence a closed cycle over which $\scrs{r}(z)$ may be integrated will exist precisely when the common exponent of the \(z_i\) in this factor is an integer, \(s \in \mathbb{Z}\) say. This requires the Heisenberg weight \(p\) of the Wakimoto module $\Wak{}$ to have the form
\begin{equation} \label{eq:p_rs}
p_{r,s}=\frac{r-1}{\alpha}-\frac{s\alpha}{2}, \qquad r\in\mathbb{Z}_{\ge 1},\quad s\in\mathbb{Z}.
\end{equation}
We remark that the multivalued function
\begin{equation} \label{eq:DefG}
G_r(z;t)=\prod_{1\le i\neq j\le r}\Bigl(1-\frac{z_i}{z_j}\Bigr)^{1/t},
\end{equation}
appearing on the right-hand side{} of \eqref{eq:scrprod'}, is just the integration kernel of the inner product \eqref{eq:innerprod} with respect to which the Jack symmetric polynomials are orthogonal. Setting $p=p_{r,s}$, \eqref{eq:scrprod} now takes the form
\begin{equation} \label{eq:ScrProd}
\Bigl.\scrs{r}(z)\Bigr\rvert_{\Wak{p_{r,s}}}
=\Bigl.G_r(z;t)\prod_{i=1}^r z_i^s\cdot
\prod_{m\ge 1}\exp\brac{-\frac{2}{\alpha}\fpowsum{m}{z}\frac{a_{-m}}{m}}
\exp\brac{\frac{2}{\alpha}\overline{\fpowsum{m}{z}}\frac{a_{m}}{m}}\cdot
\prod_{i=1}^r\beta(z_i)\Bigr\rvert_{\Wak{p_{-r,s}}}.
\end{equation}
Again, the action on $\WakRel{p_{r,s};q}$ is identical except that the right-hand side{} now acts on $\WakRel{p_{-r,s};q}$.
\subsection{Singular vectors}
The existence of cycles over which the product $\scrs{r}(z)$ of screening fields may be integrated follows from the same arguments used in the analogous question for the free field realisation of the universal Virasoro vertex operator algebra{}, because the multivalued function \(G_r(z;t)\) is the same in both cases. This question was answered for the Virasoro case (see \thmref{thm:tkthm}) by Tsuchiya and Kanie \cite{TsuFoc86} to whom we refer for further details. We will use the cycles $[\Delta_r]$ that they construct in what follows, but normalised so that
\begin{equation}
\int_{[\Delta_r]}G_r(z;t)\:\frac{\dd z_1\cdots\dd z_r}{z_1\cdots z_r}=1.
\end{equation}
We mention that there are various explicit constructions of cycles, over which screening operators can be integrated, in the conformal field theory{} literature, but that the symmetric function literature uses different constructions again. However, Cohen and Varchenko \cite{CohCyc03} showed that, up to normalisation, there is only one non-trivial homology class of cycles when the integrand is $G_r(z;t)$ times a symmetric function, bar some restrictions on \(t\), and thus the various constructions in the literature are all essentially equivalent.\footnote{It is because of these minor restrictions on \(t\) that we state our choice of class of cycle explicitly. These restrictions are only relevant for the precise statement of \thmref{thm:ConstructSVs}. A more detailed understanding of these cycles is not required for reading the remainder of this article.}
The normalised cycles \([\Delta_r]\) let us construct $\AKMA{sl}{2}$-homomorphisms (screening operators) from the products $\scrs{r}(z)$ of screening fields. Our aim is to use these homomorphisms to explicitly construct singular vectors in (relaxed) \hw{} modules{} over $\AKMA{sl}{2}$ using symmetric function technology (we refer to \appref{sec:SymmPoly} for a primer on what is needed). Of course, we need to verify that the screening operators are not just zero maps.
To this purpose, we introduce an algebra isomorphism \(\rho_\delta\), for each \(\delta\in\mathbb{C}^\ast\), from the algebra $\symfunc$ of symmetric polynomials in infinitely many variables to the universal enveloping algebra $\UEA \HA_-$ of the negative subalgebra of the Heisenberg algebra. This isomorphism is given, on the power sum generators, by
\begin{equation} \label{eq:DefRho}
\rho_{\delta}(\fpowsum{m}{y})=\delta a_{-m}.
\end{equation}
For convenience, we will always take $\rho_{\delta}$ to act only on symmetric polynomials of the variables $y_i$. For the ghosts, we analogously define injective linear maps $\sigma_r$ and $\widetilde{\sigma}_r$ from the algebra $\sympol{r}$ of symmetric polynomials in $r$ variables to the ghost universal enveloping algebra{} $\UEA \GA$. These maps are defined on the basis $\set{\dmonsym{\nu}{t} \st \ell(\nu)\le r}$ dual to the symmetric monomials $\monsym{\nu}$ by
\begin{equation} \label{eq:DefSigmas}
\sigma_r(\fdmonsym{\nu}{t}{x})=\beta_{-\nu_1-1}\cdots \beta_{-\nu_r-1},\qquad
\widetilde{\sigma}_r(\fdmonsym{\nu}{t}{x})=\beta_{-\nu_1}\cdots\beta_{-\nu_r}.
\end{equation}
We will only take these maps to act on symmetric polynomials in the $x_i$. Here, the partition $\nu$ may have length $\ell(\nu)$ strictly less than $r$; in this case, we pad the partition with zeroes so that the images $\sigma_r(\dmonsym{\nu}{t})$ and $\widetilde{\sigma}_r(\dmonsym{\nu}{t})$ are padded by $\beta_{-1}$ or $\beta_0$, respectively, so that the right hand sides of \eqref{eq:DefSigmas} each consist of \(r\) factors.
With these maps, we can prove that the $\scrs{r}(z)$ yield non-trivial screening operators and derive explicit formulae for certain (relaxed) singular vectors of the (relaxed) Verma modules over $\AKMA{sl}{2}$.
\begin{thm} \label{thm:ConstructSVs}
Let \(r\in\mathbb{Z}_{\ge 1}\), \(s\in\mathbb{Z}\) and \(t\in\mathbb{C}^\ast\) and suppose that \(d(d+1)/t\notin\mathbb{Z}\) and \(d(r-d)/t\notin\mathbb{Z}\), for all integers \(d\) satisfying \(1\le d\le r-1\).
Then,
\begin{equation}
\scrs{r}=\int_{[\Delta_r]}\scrs{r}(z_1,\dots,z_r) \: \dd z_1 \cdots \dd z_r
\end{equation}
defines non-trivial \(\AKMA{sl}{2}\)-module homomorphisms between (relaxed) Wakimoto modules:
\begin{equation}
\scrs{r}\colon\Wak{p_{r,s}}\rightarrow\Wak{p_{-r,s}},\qquad
\scrs{r}\colon\WakRel{p_{r,s};q}\rightarrow\WakRel{p_{-r,s};q}.
\end{equation}
In particular, if $\wv{p_{r,s}} = \hv{p_{r,s}} \otimes \gvac$ and $\wv{p_{r,s}; q} = \hv{p_{r,s}} \otimes \gv{q}$, where $\hv{p}$ denotes the Heisenberg \hw{} vector{} of weight $p$, $\gvac$ denotes the ghost vacuum and $\gv{q}$ denotes a ghost relaxed \hwv{}{} of weight $q \notin \mathbb{Z}$, then
\begin{subequations}
\begin{align}
\scrs{r}\wv{p_{r,s}}&=
\begin{cases}
(\rho_{-\alpha}\circ\sigma_r)
\left(\fdjack{[(-s-1)^r]}{t}{x,y}\right)\wv{p_{-r,s}} \neq 0 & \text{if \(s\le-1\),} \\
0 & \text{if \(s\ge 0\),}
\end{cases}\label{eq:SV}\\
\scrs{r}\wv{p_{r,s};q}&=
\begin{cases}
(\rho_{-\alpha}\circ\widetilde{\sigma}_r)
\left(\fdjack{[-s^r]}{t}{x,y}\right)\wv{p_{-r,s};q} \neq 0 & \text{if \(s\le0\),} \\
0 \hphantom{(\rho_{-\alpha}\circ\sigma_r)
\left(\fdjack{[(-s-1)^r]}{t}{x,y}\right)\wv{p_{-r,s}} \neq {}} & \text{if \(s\ge 1\),}
\end{cases}\label{eq:RSV}
\end{align}
\end{subequations}
where $\djack{\nu}{t}$ denotes the symmetric polynomials dual to the Jack polynomials (see \appref{sec:SymmPoly}). The $\scrs{r}$-images of $\wv{p_{r,s}}$ ($\wv{p_{r,s};q}$) in $\Wak{p_{-r,s}}$ ($\WakRel{p_{-r,s};q}$) are therefore (relaxed) singular vectors, for $s \le -1$ ($s \le 0$).
\end{thm}
\begin{proof}
We first show the non-triviality of $\scrs{r}$. Let \(\dwv{p_{-r,s}}\) and \(\dwv{p_{-r,s};q}\) denote the functionals dual to \(\wv{p_{-r,s}}\) and \(\wv{p_{-r,s};q}\), respectively. Then, it follows from the ghost commutation relations \eqref{eq:BosGhComm} and adjoint \eqref{eq:GhAdj} that
\begin{subequations}
\begin{equation}
\braket{\gv{q}}{\prod_{i=1}^r \beta(z_i) \gamma_{-s}^r\gv{q}} = (-1)^r r! \braket{\gv{q}}{\gv{q}} \prod_{i=1}^r z_i^{-s-1},\qquad \text{\(s\ge1\), \(q\in\mathbb{C}\).}
\end{equation}
Replacing the relaxed \hwv{}{} $\gv{q}$ by the ghost vacuum $\gvac$ gives the same result, but for $s \ge 0$ (because $\beta_0$ annihilates $\gvac$):
\begin{equation}
\braket{\gvac}{\prod_{i=1}^r \beta(z_i) \gamma_{-s}^r\gvac} = (-1)^r r! \braket{\gvac}{\gvac} \prod_{i=1}^r z_i^{-s-1},\qquad \text{\(s\ge0\).}
\end{equation}
\end{subequations}
The non-triviality of \(\scrs{r}\) on $\Wak{p_{r,s}}$, for \(s\ge 0\), now follows from \eqref{eq:ScrProd} and the normalisation of \([\Delta_r]\):
\begin{subequations}
\begin{align}
\dwv{p_{-r,s}}\scrs{r}\gamma_{-s}^r\wv{p_{r,s}} &=\int_{[\Delta_r]}G_r(z;t)\prod_{i=1}^r z_i^s\cdot
\dwv{p_{-r,s}}\prod_{i=1}^r \beta(z_i) \gamma_{-s}^r\wv{p_{-r,s}}\:\dd z_1\cdots\dd z_r\notag\\
&=(-1)^rr!\braket{p_{-r,s}}{p_{-r,s}}\int_{[\Delta_r]}G_r(z;t)\:\frac{\dd z_1\cdots z_r}{z_1\cdots z_r}\notag\\
&=(-1)^r r!\braket{p_{-r,s}}{p_{-r,s}}\neq0.
\end{align}
Here, we note that the $a_{-m}$ and $a_{m}$, with $m \ge 1$, appearing in \eqref{eq:ScrProd} annihilate $\dwv{p_{-r,s}}$ and $\wv{p_{r,s}}$, respectively. A similar calculation gives the relaxed version for $s\ge 1$:
\begin{equation}
\dwv{p_{-r,s};q}\scrs{r}\gamma_{-s}^r\wv{p_{r,s};q} = (-1)^r r!\braket{p_{-r,s};q}{p_{-r,s};q} \neq 0.
\end{equation}
\end{subequations}
We note that the conformal weight of $\wv{p_{-r,s}}$ ($\wv{p_{-r,s}; q}$) is greater than that of $\wv{p_{r,s}}$ ($\wv{p_{r,s}; q}$) by $rs$. It follows that $\scrs{r}\wv{p_{r,s}}=0$ ($\scrs{r}\wv{p_{r,s};q}=0$) for all $s\ge 1$ because $\scrs{r}$ is an $\AKMA{sl}{2}$-module homomorphism and hence it preserves conformal weights.
We settle the non-triviality of $\scrs{r}$ for the remaining values of \(s\) by explicitly computing the image of the (relaxed) \hw{} vectors{} \(\wv{p_{r,s}}\) and \(\wv{p_{r,s};q}\). In the former case, we obtain
\begin{subequations}
\begin{align}
\scrs{r}\wv{p_{r,s}}
&= \int_{[\Delta_r]}G_r(z;t)\prod_{i=1}^r z_i^s\cdot
\prod_{m\ge1}\exp\left(\frac{-2}{\alpha}\fpowsum{m}{z}\frac{a_{-m}}{m}\right)\cdot
\prod_{i=1}^r\beta(z_i)\wv{p_{-r,s}}\:\dd z_1\cdots \dd z_r\notag\\
&=\int_{[\Delta_r]}G_r(z;t)\prod_{i=1}^r z_i^s\cdot
\prod_{m\ge1}\exp\left(\frac{-2}{\alpha}\frac{\fpowsum{m}{z}a_{-m}}{m}\right)
\sum_{\nu\st\ell(\nu)\le r}\beta_{-\nu_1-1}\cdots\beta_{-\nu_r-1}
\fmonsym{\nu}{z}\wv{p_{-r,s}}\:\dd z_1\cdots \dd z_r\notag\\
&=\int_{[\Delta_r]}G_r(z;t)\prod_{i=1}^r z_i^s\cdot
\rho_{-\alpha}\biggl(\prod_{m\ge1}\exp\left(\frac{1}{t}\frac{\fpowsum{m}{y}\fpowsum{m}{z}}{m}\right)\biggr)\cdot
\sigma_r\biggl(\sum_{\nu\st\ell(\nu)\le r}\fdmonsym{\nu}{t}{x}\fmonsym{\nu}{z}\biggr)\wv{p_{-r,s}}\:\dd
z_1\cdots \dd z_r\notag\\
&=\int_{[\Delta_r]}G_r(z;t)\prod_{i=1}^r z_i^{s+1}\cdot
(\rho_{-\alpha}\circ\sigma_r)\left(\prod_{m\ge1}\exp\left(\frac{1}{t}\frac{\left(\fpowsum{m}{x}+\fpowsum{m}{y}\right)\fpowsum{m}{z}}{m}\right)\right)
\wv{p_{-r,s}}\:\frac{\dd z_1\cdots \dd z_r}{z_1\cdots z_r} \notag
\intertext{(if $s=0$, this vanishes, in agreement with \eqref{eq:SV}, as the powers of the $z_i$ in the integrand are all positive)}
&=\sum_{\nu\st\ell(\nu)\le r} \intprod{\fjack{[(-1-s)^r]}{t}{z}}{\fjack{\nu}{t}{z}}_t^r
\:(\rho_{-\alpha}\circ\sigma_r)\left(\fdjack{\nu}{t}{x,y}\right) \wv{p_{-r,s}}\notag\\
&=(\rho_{-\alpha}\circ\sigma_r)\left(\fdjack{[(-1-s)^r]}{t}{x,y}\right)\wv{p_{-r,s}}.
\end{align}
Here, we have used the finite-variable version of \eqref{eq:Cauchy} to rewrite the sum over the symmetric monomials $\monsym{\nu}$ and their duals $\dmonsym{\nu}{t}$ as a product of exponentials of power sums. Then, we note that $\fpowsum{m}{x}+\fpowsum{m}{y}$ is the power sum $\fpowsum{m}{x,y}$ in both \(x_i\) and \(y_j\) variables and use \eqref{eq:Cauchy} again to expand the product of exponentials in terms of Jack polynomials and their duals. Finally, we use \propref{prop:jackprops} to identify $\prod_i z_i^{-1-s}$ as a Jack polynomial in the $z_i^{-1}$, then apply the orthogonality of Jack polynomials, and lastly note that the Jack polynomial norm follows from the normalisation of \([\Delta_r]\). For $s\le -1$, the result is non-vanishing since $\rho_{-\alpha}$ and $\sigma_r$ are injective. This proves the non-triviality of $\scrs{r}$ on $\Wak{p_{r,s}}$ for all $s \in \mathbb{Z}$, as well as \eqref{eq:SV}.
A similar computation in the relaxed sector results in
\begin{align}
\scrs{r}\wv{p_{r,s};q}
&=\int_{[\Delta_r]}G_r(z;t)\prod_{i=1}^r z_i^s\cdot
\prod_{m\ge1}\exp\left(\frac{-2}{\alpha}\frac{\fpowsum{m}{z}a_{-m}}{m}\right)\cdot
\prod_{i=1}^r\beta(z_i)\wv{p_{-r,s};q}\:\dd z_1\cdots \dd z_r\notag\\
&=\int_{[\Delta_r]}G_r(z;t)\prod_{i=1}^r z_i^s\cdot
\prod_{m\ge1}\exp\left(\frac{-2}{\alpha}\frac{\fpowsum{m}{z}a_{-m}}{m}\right)\cdot
\sum_{\nu\st\ell(\nu)\le r}\beta_{-\nu_1}\cdots\beta_{-\nu_r}\fmonsym{\nu}{z}\wv{p_{-r,s};q}\: \frac{\dd z_1\cdots \dd z_r}{z_1\cdots z_r}\notag\\
&=(\rho_{-\alpha}\circ\widetilde{\sigma_r})\left(\fdjack{[-s^r]}{t}{x,y}\right)
\wv{p_{-r,s};q},
\end{align}
\end{subequations}
which is likewise non-vanishing for $s\le 0$ as $\rho_{-\alpha}$ and $\widetilde{\sigma_r}$ are injective. This proves \eqref{eq:RSV} and the non-triviality of $\scrs{r}$ on $\WakRel{p_{r,s};q}$, for all $s \in \mathbb{Z}$.
\end{proof}
Suppose now that the $\AKMA{sl}{2}$ level $k$ is admissible: $t=k+2=\frac{u}{v}$, where $u \in \mathbb{Z}_{\ge 2}$ and $v \in \mathbb{Z}_{\ge 1}$ are coprime. The $\AKMA{sl}{2}$ vacuum may be identified with the Wakimoto \hw{} vector{} $\wv{0} = \wv{p_{1,0}} = \wv{p_{-u+1,-v}} \in \Wak{p_{-u+1,-v}}$ (we note the symmetry $p_{r,s} = p_{r+u,s+v}$). \thmref{thm:ConstructSVs} then guarantees that the $\AKMA{sl}{2}$-module homomorphism $\scrs{u-1}$ acts non-trivially on $\wv{p_{u-1,-v}}$ to give a non-trivial singular vector in the vacuum module. The corresponding field then generates the non-trivial proper ideal of the universal vertex operator algebra{}, proving the following result:
\begin{cor} \label{cor:WakVacIsUniversal}
The universal vertex operator algebra{} \(\UVOA{k}\) of $\AKMA{sl}{2}$ at non-critical level $k\neq-2$ may be realised as a subalgebra of the Wakimoto vertex operator algebra{}.
\end{cor}
\noindent This is, of course, obvious if $k$ is not admissible. What it means
in the admissible level case is that calculations requiring the singular
vector of the vacuum module of $\AKMA{sl}{2}$ may be equivalently carried out
in the free field realisation using Jack symmetric polynomials.
\section{Classifying admissible modules} \label{sec:AdmMod}
In this section, we specialise to admissible levels $k=t-2$, where $t=\frac{u}{v}$ and $u\in\mathbb{Z}_{\ge2}$ and $v\in\mathbb{Z}_{\ge1}$ are coprime. Then, the universal vertex operator algebra{} $\UVOA{k}$ of $\AKMA{sl}{2}$ is not simple and the unique maximal proper ideal is generated by the field that corresponds to the singular vector{} of the vacuum module (see \secref{sec:RelEx}). As we saw in \thmref{thm:ConstructSVs} and \corref{cor:WakVacIsUniversal}, the vacuum module may be constructed as a submodule of the Wakimoto vacuum module $\Wak{0}$ and its singular vector is then explicitly given by $\scrs{u-1} \wv{p_{u-1,-v}}$.
Quotienting by this maximal proper ideal, that is, setting the singular vector{} $\scrs{u-1} \wv{p_{u-1,-v}}$ to zero, amounts to replacing the universal vertex operator algebra{} $\UVOA{k}$ by its simple counterpart \(\AdmMod{u}{v}\). Our aim in this section is to use the explicit expression for the vacuum singular vector{} to classify the possible (relaxed) highest weights and thereby determine the spectrum of $\AdmMod{u}{v}$-modules. By \corref{cor:WakVacIsUniversal}, these calculations may be performed in Wakimoto's free field realisation. More specifically, we will use symmetric polynomial technology to compute a generator of the annihilating ideal in Zhu's algebra. For readers unfamiliar with Zhu's algebra, we refer to Appendix \ref{sec:zhu} for motivation, basic definitions and a very short primer.
An old result of Frenkel and Zhu \cite{FreVer92} states that Zhu's algebra \(\zhu{\UVOA{k}}\) for \(\UVOA{k}\) is nothing but the universal enveloping algebra{} $\UEA\:\SLA{sl}{2}$ of (non-affine) \(\SLA{sl}{2}\) (see \propref{prop:zhusl2}). By \propref{prop:zhuideal}, Zhu's algebra \(\zhu{\AdmMod{u}{v}}\) for the quotient \(\AdmMod{u}{v}\) is then the quotient of \(\UEA\:\SLA{sl}{2}\) by the annihilating ideal generated by the representative of the singular vector \(\scrs{u-1}\wv{p_{u-1,-v}}\). Since Zhu's algebra is filtered by conformal weight, whereas the conformal weight of \(e(z)\), \(h(z)\) and \(f(z)\) is 1 and the conformal weight of the singular vector is \((u-1)v\), it follows that the image of the singular vector in \(\zhu{\UVOA{k}}\) is a polynomial in the \(\SLA{sl}{2}\) generators of total degree at most \((u-1)v\). Furthermore, as Zhu's algebra is just the algebra of zero modes acting on relaxed \hwvs{}{}, the polynomial corresponding to the singular vector can be determined by evaluating the zero mode of the singular vector on general relaxed \hwvs{}{}, as in \cite{FeiAnn92}, since this is equivalent to evaluating the polynomial at infinitely many points.
However, the $\SLA{sl}{2}$-weight of the vacuum singular vector{} $\scrs{u-1}\wv{p_{u-1,-v}}$ is $\lambda_{u-1,-v} = \lambda_{p_{u-1,-v}} = 2(u-1)$, which means that the corresponding field and its zero mode shifts the \(\SLA{sl}{2}\)-weight of any relaxed \hwv{}{} upon which it acts by this amount. It is far more convenient to work with a field that does not shift \(\SLA{sl}{2}\)-weights and so we instead consider the field corresponding to the $\SLA{sl}{2}$-weight $0$ vector
\begin{align}\label{eq:idealgenerator}
f_0^{u-1}\scrs{u-1}\wv{p_{u-1,-v}} &= \scrs{u-1}f_0^{u-1}\wv{p_{u-1,-v}} \notag \\
&= \lambda_{u-1,-v} \brac{\lambda_{u-1,-v} - 1} \cdots \brac{\lambda_{u-1,-v} - u+2} \scrs{u-1}\gamma_0^{u-1}\wv{p_{u-1,-v}}.
\end{align}
Since $u \ge 2$, we may renormalise this vector by dividing by the non-zero $\lambda_{u-1,-v}$-dependent factors on the right-hand side{}. The field corresponding to this renormalised vector is then
\begin{align} \label{eq:DefChi}
\chi(w)&=\int_{[\Delta_r]}\scrs{u-1}(z_1+w,\dots,z_{u-1}+w)\vop{p_{u-1,-v}}(w)\gamma(w)^{u-1} \:\dd z_1\cdots\dd z_{u-1} \notag \\
&= \int_{[\Delta_r]} \vop{-2/\alpha}(z_1+w)\cdots\vop{-2/\alpha}(z_{u-1}+w)\vop{p_{u-1,-v}}(w) \beta(z_1+w)\cdots\beta(z_{u-1}+w)\gamma(w)^{u-1} \:\dd z_1\cdots\dd z_{u-1}.
\end{align}
We note that $e_0^{u-1}$ acting on the vector $f_0^{u-1}\scrs{u-1}\wv{p_{u-1,-v}}$ gives a non-zero multiple of the singular vector{} $\scrs{u-1}\wv{p_{u-1,-v}}$. It follows that the annihilating ideal that we obtain from $\chi(w)$ will be the same as that which we would have obtained if we had instead worked with the singular vector{} directly.
\subsection{Admissible \hw{} modules{}} \label{sec:HWAdmMod}
In this section, we will evaluate the action of the zero mode of \(\chi(w)\) on a general \hw{} vector{} before moving on to the evaluation of \(\chi(w)\) on relaxed \hwvs{}{}. This highest weight{} computation will not yet yield sufficient information to determine the image of \(\chi(w)\) in Zhu's algebra \(\zhu{\UVOA{k}}\) as a polynomial in the \(\SLA{sl}{2}\) generators \(e,h\) and \(f\), but it will constrain the weights of admissible \hw{} vectors{} to a finite set.
Every \hw{} vector{} of $\UVOA{k}$ may be realised as a \hw{} vector{} of some Wakimoto module $\Wak{p}$. It follows that the Heisenberg weights $p$ of the admissible \hw{} vectors{} of $\AKMA{sl}{2}$, that is, the \hw{} vectors{} of the $\AdmMod{u}{v}$-modules, are zeroes of
\begin{multline}
\dwv{p}\chi(w)\wv{p} = \int_{[\Delta_r]} \inner{\hv{p}}{\vop{-2/\alpha}(z_1+w) \cdots \vop{-2/\alpha}(z_{u-1}+w) \vop{p_{u-1,-v}}(w) \: \hv{p}} \\
\cdot \inner{\gvac}{\beta(z_1+w) \cdots \beta(z_{u-1}+w) \gamma(w)^{u-1} \: \gvac} \: \dd z_1\cdots\dd z_{u-1}.
\end{multline}
The ghost contribution is easily evaluated by computing the operator product expansion{} of the ghost fields. Since $\gvac$ is the ghost vacuum, only the fully contracted part of the operator product expansion{} contributes:
\begin{align}
\inner{\gvac}{\beta(z_1+w) \cdots \beta(z_{u-1}+w) \gamma(w)^{u-1} \: \gvac} &= \inner{\gvac}{\normord{\beta(z_1+w) \cdots \beta(z_{u-1}+w)} \normord{\gamma(w) \cdots \gamma(w)} \: \gvac} \notag \\
&= \frac{(-1)^{u-1} (u-1)!}{z_1 \cdots z_{u-1}} \inner{\gvac}{\gvac}.
\end{align}
The contribution from the free boson part of the Wakimoto realisation is likewise easily determined. Up to non-zero constant factors, which obviously do not affect the zeroes of $\dwv{p}\chi(w)\wv{p}$, this contribution is
\begin{align} \label{eq:FBContribution}
&\inner{\hv{p}}{\vop{-2/\alpha}(z_1+w) \cdots \vop{-2/\alpha}(z_{u-1}+w) \vop{2(u-1)/\alpha}(w) \: \hv{p}} \notag \\&\mspace{200mu} = \prod_{1 \le i \neq j \le u-1} (z_i-z_j)^{1/t} \cdot \prod_{i=1}^{u-1} (z_i+w)^{-2p/\alpha} z_i^{-2(u-1)/t} \cdot w^{2(u-1)p/\alpha} \inner{\hv{p}}{\hv{p}} \notag \\
&\mspace{200mu} = G_{u-1}(z;t) \prod_{i=1}^{u-1} z_i^{-v} \cdot \prod_{i=1}^{u-1} \brac{1 + \frac{z_i}{w}}^{-2p/\alpha} \inner{\hv{p}}{\hv{p}},
\end{align}
where we have used \eqref{eq:CompVerOps}, noting that all the terms involving the $a_n$ with $n \neq 0$ either annihilate the Heisenberg vacuum $\hv{p}$ or its dual, and recognised the kernel of the symmetric polynomial inner product from \eqref{eq:DefG}.
Putting these contributions together, we find that the admissible Heisenberg weights of an $\AKMA{sl}{2}$ \hw{} vector{} (in the Wakimoto free field realisation) must satisfy
\begin{equation} \label{eq:HWConstraint}
0 = \dwv{p}\chi(w)\wv{p} = \int_{[\Delta_r]} G_{u-1}(z;t) \: \overline{\fjack{[v^{u-1}]}{t}{z}} \: \prod_{i=1}^{u-1} \brac{1 + \frac{z_i}{w}}^{-2p/\alpha} \: \frac{\dd z_1\cdots\dd z_{u-1}}{z_1 \cdots z_{u-1}} \braket{p}{p},
\end{equation}
where we have recognised the product of the $z_i$ as a Jack polynomial using \propref{prop:jackprops} (recall that the overline indicates a symmetric polynomial in the inverse variables $z_i^{-1}$). We have also, again, neglected an overall non-zero constant factor. This expression has the form of an inner product for symmetric polynomials; to evaluate it, we only need to decompose the product over $i$ into Jack polynomials. For this, we use specialisation in the form given in \eqref{eq:SpecialisingToJacks} with $x_i = -z_i/w$ for $i \le u-1$, $x_i=0$ for $i>u-1$, and $X=\alpha p$:
\begin{equation} \label{eq:FirstSpecialisation}
\prod_{i=1}^{u-1} \brac{1 + \frac{z_i}{w}}^{-2p/\alpha} = \sum_{\tau} \fjack{\tau}{t}{-z_1/w, \ldots, -z_{u-1}/w, 0, 0, \ldots} \: \func{\Xi_{\alpha p}}{\fdjack{\tau}{t}{y}} = \sum_{\tau} \frac{(-1)^{\abs{\tau}}}{w^{\abs{\tau}}} \fjack{\tau}{t}{z} \: \func{\Xi_{\alpha p}}{\fdjack{\tau}{t}{y}}.
\end{equation}
Here, we have also used the homogeneity of the Jack polynomials ($\abs{\tau}$ is the sum of the parts of the partition $\tau$).
Using the orthogonality of Jack polynomials, \eqref{eq:HWConstraint} now becomes
\begin{align}
0 &= \sum_{\tau} w^{-\abs{\tau}} \intprod{\jack{[v^{u-1}]}{t}}{\jack{\tau}{t}}_t^{u-1} \func{\Xi_{\alpha p}}{\fdjack{\tau}{t}{y}} = w^{-\abs{[v^{u-1}]}} \intprod{\jack{[v^{u-1}]}{t}}{\jack{[v^{u-1}]}{t}}_t^{u-1} \func{\Xi_{\alpha p}}{\fdjack{[v^{u-1}]}{t}{y}} \notag \\
&= w^{-(u-1)v} \: \func{\Xi_{\alpha p}}{\fdjack{[v^{u-1}]}{t}{y}} = w^{-(u-1)v} \prod_{b \in [v^{u-1}]} \frac{\alpha p + t a'(b) - l'(b)}{t \tbrac{a(b)+1} + l(b)} \notag \\
&= w^{-(u-1)v} \prod_{i=1}^{u-1} \prod_{j=1}^{v} \frac{\alpha p + t (j-1) - (i-1)}{t \tbrac{v-j+1} + u-1-i},
\end{align}
where the result of specialising the dual Jack polynomials $\djack{\tau}{t}$ is given in \propref{prop:jackprops} (or \eqref{eq:SpecialisingToJacks}) and the arm and leg (co)lengths of \eqref{eq:ArmLeg} are easy to determine for the rectangular partition $[v^{u-1}]$. As the denominators of the factors appearing in this expression are all strictly positive, we arrive at the constraint
\begin{equation}
\prod_{i=1}^{u-1} \prod_{j=0}^{v-1} \tbrac{\alpha p + tj - (i-1)} =
\prod_{i=1}^{u-1} \prod_{j=0}^{v-1} \tbrac{\alpha p - \alpha p_{i,j}} = 0.
\end{equation}
This proves the following result:
\begin{prop} \label{prop:ClassHWMods}
Let $\lambda_{r,s} = \lambda_{p_{r,s}} = \alpha p_{r,s} = r-1-ts$. Then, every \hw{} vector{} of an $\AdmMod{u}{v}$-module has $\SLA{sl}{2}$-weight of the form $\lambda_{r,s}$, where $r=1, 2, \ldots, u-1$ and $s=0, 1, \ldots, v-1$.
\end{prop}
\noindent Note that when $v=1$, so that $k=t-2=u-2 \in \mathbb{Z}_{\ge 0}$, the allowed $\SLA{sl}{2}$-weights $\lambda_{r,s} = \lambda_{r,0} = r-1$ belong to the set $\set{0, 1, \ldots, k}$. These are, of course, the highest weight{}s of the integrable $\AKMA{sl}{2}$-modules and are well known to be the $\AdmMod{k+2}{1}$-modules that arise in the Wess-Zumino-Witten{} models defined on the Lie group $\SLG{SU}{2}$.
At this point, we cannot say whether all these \hw{} vectors{} do actually appear in an $\AdmMod{u}{v}$-module. For this, we need to work out the constraints on an arbitrary relaxed \hwv{}{} because it is these constraints which allow us to write down the generator of the annihilating ideal of Zhu's algebra. It is, however, well known that the Verma module $\Ver{\lambda_{r,s}}$, with $r=1, 2, \ldots, u-1$ and $s=0, 1, \ldots, v-1$, has infinitely many linearly independent singular vectors{}; however, none have $\SLA{sl}{2}$-weights belonging to the allowed set except the generator of $\SLA{sl}{2}$-weight $\lambda_{r,s}$. It follows that there are only finitely many highest weight{} $\AdmMod{u}{v}$-modules:
\begin{cor} \label{cor:ClassHWMods}
Every highest weight{} $\AdmMod{u}{v}$-module is isomorphic to one of the simple $\AKMA{sl}{2}$-modules $\Irr{\lambda_{r,0}}$ or $\Disc{\lambda_{r,s}}$, where $r=1, 2, \ldots, u-1$ and $s=1, 2, \ldots, v-1$.
\end{cor}
\noindent Again, we have not yet proven that all these $\UVOA{k}$-modules are actually $\AdmMod{u}{v}$-modules. However, it is germane to point out, at this point, that these \hw{} modules{} are precisely the admissible modules first discussed by Kac and Wakimoto \cite{KacMod88}.
\subsection{Admissible \rhw{} modules{}} \label{sec:HWRelAdmMod}
We now turn to the more intricate, but ultimately more rewarding, analysis of the \rhw{} modules{} of $\AdmMod{u}{v}$. The goal is to determine the image of the field \(\chi(w)\) in Zhu's algebra \(\zhu{\UVOA{k}} \cong \UEA\:\SLA{sl}{2}\) as a polynomial \(I_{u,v}(e,h,f)\) in the \(\SLA{sl}{2}\) generators. Here, we identify the $\SLA{sl}{2}$ generators $e$, $h$ and $f$ with the images of $e(z)$, $h(z)$ and $f(z)$, respectively, in $\zhu{\UVOA{k}}$. Since Zhu's algebra is nothing but the algebra of zero modes acting on relaxed \hwvs{}{} (\appref{sec:zhu}), $I_{u,v}(e,h,f)$ is identical to the polynomial \(I_{u,v}(e_0,h_0,f_0)\) that describes the action of the zero mode \(\chi_0\) on relaxed \hwvs{}{}. We remark that as the $\SLA{sl}{2}$-weight of $\chi(w)$ is $0$, the polynomial $I_{u,v}$ may be expressed as a polynomial in $h_0$ and the Virasoro zero mode $L_0$.
So as in the previous section, we evaluate matrix elements containing the field \(\chi(w)\), but this time the ``bra'' and the ``ket'' will be relaxed \hwvs{}{} from a general relaxed Wakimoto module:
\begin{multline} \label{eq:RelaxedConstraint}
\dwv{p;q}\chi(w)\wv{p;q} = \int_{[\Delta_r]} \inner{\hv{p}}{\vop{-2/\alpha}(z_1+w) \cdots \vop{-2/\alpha}(z_{u-1}+w) \vop{p_{u-1,-v}}(w) \: \hv{p}} \\
\cdot \inner{\gv{q}}{\normord{\beta(z_1+w) \cdots \beta(z_{u-1}+w)} \normord{\gamma(w)^{u-1}} \: \gv{q}} \: \dd z_1\cdots\dd z_{u-1}.
\end{multline}
We recall that $\gv{q}$ satisfies $J_0 \gv{q} = \gamma_0 \beta_0 \gv{q} = q \gv{q}$.
The contribution from the free boson is exactly the same as in the non-relaxed case and was given in \eqref{eq:FBContribution}. The ghost contribution, however, requires more work. Wick's theorem lets us write this contribution in terms of contractions and normally-ordered products:
\begin{multline}
\inner{\gv{q}}{\normord{\beta(z_1+w) \cdots \beta(z_{u-1}+w)} \normord{\gamma(w)^{u-1}} \: \gv{q}} \\
= \sum_{I\subseteq\{1,\dots,u-1\}} \frac{(-1)^{\abs{I}} (u-1)!}{(u-1-\abs{I})!} \prod_{i \in I} z_i^{-1} \cdot \inner{\gv{q}}{\normord{\displaystyle \prod_{i \notin I} \beta(z_i+w) \cdot \gamma(w)^{u-1-\abs{I}}} \: \gv{q}}
\end{multline}
Here, each factor of $-z_i^{-1}$ is the contraction of $\beta(z_i+w)$ and $\gamma(w)$ and the factorials count how many such contractions are needed. As relaxed \hwvs{}{} are not necessarily annihilated by $\beta_0$, the normally-ordered factor is quite non-trivial:
\begin{align}
\inner{\gv{q}}{\normord{\displaystyle \prod_{i \notin I} \beta(z_i+w) \cdot \gamma(w)^{u-1-\abs{I}}} \: \gv{q}} &= \inner{\gv{q}}{\gamma_0^{u-1-\abs{I}} \beta_0^{u-1-\abs{I}} \: \gv{q}} \prod_{i \notin I} (z_i+w)^{-1} \notag \\
&= (u-1-\abs{I})! \binom{u-2-\abs{I}+q}{u-1-\abs{I}} \prod_{i \notin I} (z_i+w)^{-1} \inner{\gv{q}}{\gv{q}}.
\end{align}
Up to an overall non-zero constant factor, the total ghost contribution is therefore
\begin{align}
&\sum_{I\subseteq\{1,\dots,u-1\}} (-1)^{\abs{I}} \binom{u-2-\abs{I}+q}{u-1-\abs{I}} \prod_{i \in I} z_i^{-1} \cdot \prod_{i \notin I} (z_i+w)^{-1} \notag \\
&\mspace{100mu} = \prod_{i=1}^{u-1} (z_i+w)^{-1} \cdot \sum_{n=0}^{u-1} (-1)^n \binom{u-2-n+q}{u-1-n}
\felsym{n}{1 + \frac{w}{z_1},\dots,1 + \frac{w}{z_{u-1}}} \notag \\
&\mspace{100mu} = \prod_{i=1}^{u-1} \Bigl( 1 + \frac{z_i}{w} \Bigr)^{-1} \cdot \sum_{n=0}^{u-1} (-1)^n \binom{u-2-n+q}{u-1-n} \sum_{m=0}^n \binom{u-1-m}{n-m} \: \overline{\felsym{m}{z}} \: w^{m-(u-1)},
\end{align}
where $\elsym{m}$ denotes the $m$-th elementary symmetric polynomial and we have used the identity
\begin{equation}
\felsym{n}{1+x_1,\dots,1+x_{u-1}}=\sum_{m=0}^n\binom{u-1-m}{n-m}\felsym{m}{x}
\end{equation}
to get from the second to the third line.
Combining this with the free boson contribution \eqref{eq:FBContribution}, the matrix element \eqref{eq:RelaxedConstraint} is thus proportional to
\begin{align}
&\int_{[\Delta_r]} G_{u-1}(z;t) \prod_{i=1}^{u-1} z_i^{-(v-1)} \cdot \prod_{i=1}^{u-1} \brac{1 + \frac{z_i}{w}}^{-2p/\alpha - 1} \notag \\
&\mspace{100mu} \cdot \sum_{n=0}^{u-1} (-1)^n \binom{u-2-n+q}{u-1-n} \sum_{m=0}^n \binom{u-1-m}{n-m} \: \overline{\felsym{m}{z}} \: w^m \: \frac{\dd z_1\cdots\dd z_{u-1}}{z_1 \cdots z_{u-1}} \notag \\
&= \sum_{n=0}^{u-1} (-1)^n \binom{u-2-n+q}{u-1-n} \sum_{m=0}^n \binom{u-1-m}{n-m} w^m \intprod{\fjack{[(v-1)^{u-1}]}{t}{z} \fjack{[1^m]}{t}{z}}{\prod_{i=1}^{u-1} \brac{1 + \frac{z_i}{w}}^{-2p/\alpha - 1}}_t^{u-1},
\end{align}
where we recall that elementary symmetric polynomials are examples of Jack polynomials (\propref{prop:jackprops}). Using \eqref{eq:JackByRectJack} and specialisation as in \eqref{eq:FirstSpecialisation}, but with $X = \alpha p + t$, this reduces to
\begin{align} \label{eq:RelaxedConstraint'}
&\sum_{n=0}^{u-1} (-1)^n \binom{u-2-n+q}{u-1-n} \sum_{m=0}^n \binom{u-1-m}{n-m} w^m \sum_{\tau} \frac{(-1)^{\abs{\tau}}}{w^{\abs{\tau}}} \intprod{\fjack{[(v-1)^{u-1}] + [1^m]}{t}{z}}{\fjack{\tau}{t}{z}}_t^{u-1} \func{\Xi_{\alpha p + t}}{\fdjack{\tau}{t}{y}} \notag \\
&= \sum_{n=0}^{u-1} (-1)^n \binom{u-2-n+q}{u-1-n} \sum_{m=0}^n \binom{u-1-m}{n-m} \frac{(-1)^{\abs{\mu}}}{w^{\abs{\mu} - m}} \intprod{\fjack{\mu}{t}{z}}{\fjack{\mu}{t}{z}}_t^{u-1} \func{\Xi_{\alpha p + t}}{\fdjack{\mu}{t}{y}} \notag \\
&= w^{-(u-1)(v-1)} \sum_{n=0}^{u-1} \binom{u-2-n+q}{u-1-n} \sum_{m=0}^n (-1)^{m+n} \binom{u-1-m}{n-m} \intprod{\fjack{[1^m]}{t}{z}}{\fjack{[1^m]}{t}{z}}_t^{u-1} \func{\Xi_{\alpha p + t}}{\fdjack{\mu}{t}{y}},
\end{align}
where $\mu = [(v-1)^{u-1}] + [1^m] = [v^m, (v-1)^{u-1-m}]$ and $\abs{\mu} = (u-1)(v-1)+m$. In the last step, we have used \eqref{eq:JackByRectJack} again and the definition \eqref{eq:innerprod} of the symmetric polynomial inner product.
\propref{prop:jackprops} gives the norm squared of $\jack{[1^m]}{t} = \elsym{m}$ and the specialisation of $\djack{\mu}{t}$ as
\begin{subequations}
\begin{align}
\intprod{\fjack{[1^m]}{t}{z}}{\fjack{[1^m]}{t}{z}}_t^{u-1} &= \prod_{i=1}^m \frac{(u-i)(t+m-i)}{(m-i+1)(t+u-1-i)} = \binom{u-1}{m} \prod_{i=1}^m \frac{t+m-i}{t+u-1-i}, \\
\func{\Xi_{\alpha p + t}}{\fdjack{\mu}{t}{y}} &= \prod_{j=1}^{v-1} \sqbrac{\prod_{i=1}^m \frac{\alpha p + tj-i+1}{2u-1-i-t(j-1)} \cdot \prod_{i=m+1}^{u-1} \frac{\alpha p + tj-i+1}{2u-1-i-tj}} \cdot \prod_{i=1}^m \frac{\alpha p + u-i+1}{t+m-i} \notag \\
&= \frac{\displaystyle \prod_{r=1}^{u-1} \prod_{s=1}^{v-1} (\lambda_p - \lambda_{r,s}) \cdot \prod_{i=0}^{m-1} (\alpha p +u-i)}{\displaystyle \prod_{i=1}^{u-1} \prod_{j=2}^{v-1} \tbrac{2u-1-i-t(j-1)} \cdot \prod_{i=0}^{m-1} \tbrac{2(u-1)-i} \cdot \prod_{i=m+1}^{u-1} (t+u-1-i) \cdot \prod_{i=1}^m (t+m-i)}. \label{eq:RelaxedSpecialisation}
\end{align}
\end{subequations}
Here, we have assumed that $v>1$; if $v=1$, then the denominator of \eqref{eq:RelaxedSpecialisation} is just $\prod_{i=1}^m (t+m-i)$. Noting that the double product in the denominator of \eqref{eq:RelaxedSpecialisation} is a constant, independent of $m$, $n$, $p$ and $q$, the matrix element \eqref{eq:RelaxedConstraint} is thus proportional to
\begin{multline}
\prod_{r=1}^{u-1} \prod_{s=1}^{v-1} (\lambda_p - \lambda_{r,s}) \cdot \sum_{n=0}^{u-1} \binom{u-2-n+q}{u-1-n} \sum_{m=0}^n (-1)^{m+n} \binom{u-1-m}{n-m} \binom{u-1}{m} \prod_{i=0}^{m-1} \frac{\alpha p + u-i}{2(u-1)-i} \\
= \prod_{r=1}^{u-1} \prod_{s=1}^{v-1} (\lambda_p - \lambda_{r,s}) \cdot \sum_{\ell = 0}^{u-1} \binom{q-1}{\ell} \binom{u-1+\ell}{u-1} \binom{\alpha p + u}{u-1-\ell},
\end{multline}
where in addition to suppressing the denominator we have also suppressed an overall non-zero constant factor that arises when simplifying the binomial expressions.
The double product in this expression can be interpreted by noting that
\begin{equation}
(\lambda_p - \lambda_{r,s}) (\lambda_p - \lambda_{u-r,v-s}) = 4t(\Delta_p - \Delta_{r,s}), \qquad \Delta_{r,s} = \Delta_{p_{r,s}} = \frac{(r-ts)^2-1}{4t} = \frac{(vr-us)^2-v^2}{4uv},
\end{equation}
by \eqref{eq:RelaxedWts}. If we define $K(u,v)$ to be the set of pairs $(r,s) \in \set{1, \ldots, u-1} \times \set{1, \ldots, v-1}$ with $(r,s)$ and $(u-r,v-s)$ identified,
then the matrix element \eqref{eq:RelaxedConstraint} is given by
\begin{equation} \label{eq:RelaxedConstraint''}
\dwv{p;q}\chi(w)\wv{p;q}=\text{const}\cdot
\prod_{(r,s) \in K(u,v)} (\Delta_p - \Delta_{r,s}) \cdot \sum_{\ell = 0}^{u-1} \binom{q-1}{\ell} \binom{u-1+\ell}{u-1} \binom{\alpha p + u}{u-1-\ell}.
\end{equation}
In order to use this to find a generator of the annihilating ideal in Zhu's algebra, the sum factor in \eqref{eq:RelaxedConstraint''} needs to be expressible in terms of \(\SLA{sl}{2}\) data. To demonstrate this, we define a function $f$ of $u$, $p$ and $q$ by
\begin{equation} \label{eq:DefF}
f_{p;q}(u) = \sum_{\ell = 0}^{u-1} \binom{q-1}{\ell} \binom{u-1+\ell}{u-1} \binom{\alpha p + u}{u-1-\ell}.
\end{equation}
For small values of $u$, it gives polynomials in $\lambda_{p;q}$ and $\Delta_p$:
\begin{equation} \label{eq:SumFactorExs}
f_{p;q}(2) = \lambda_{p;q}, \qquad
f_{p;q}(3) = \frac{3}{4} \lambda_{p;q}^2 - t \Delta_p, \qquad
f_{p;q}(4) = \frac{5}{12} \lambda_{p;q}^3 + \brac{\frac{1}{3} - t \Delta_p} \lambda_{p;q}.
\end{equation}
Of course, $\alpha = \sqrt{2t} = \sqrt{2u/v}$ also depends upon $u$, but may be regarded as an independent variable for the following analysis because of its $v$-dependence.
\begin{prop}
For each $u \in \mathbb{Z}_{\ge 0}$, $f_{p;q}(u)$ is a polynomial in $\lambda_{p;q}$ and $\Delta_p$ that satisfies the recursion relation
\begin{equation} \label{eq:Recurse}
f_{p;q}(u+2) = \frac{(2u+1) \lambda_{p;q}}{(u+1)^2} f_{p;q}(u+1) - \frac{4t \Delta_p - (u-1)(u+1)}{(u+1)^2} f_{p;q}(u).
\end{equation}
\end{prop}
\begin{proof}
We apply Zeilberger's creative telescoping algorithm, see \cite{ZeiA=B96} for background. If we set
\begin{equation}
F_{p;q}(u,\ell) = \binom{q-1}{\ell} \binom{u-1+\ell}{u-1} \binom{\alpha p + u}{u-1-\ell},
\end{equation}
then the algorithm constructs the recursion relation
\begin{equation}
\tbrac{(\alpha p + 1)^2 - u^2} F_{p;q}(u,\ell) - (2u+1) \lambda_{p;q} F_{p;q}(u+1,\ell) + (u+1)^2 F_{p;q}(u+2,\ell) = G(u,\ell+1) - G(u,\ell),
\end{equation}
where $G(u,\ell) = R(u,\ell) F_{p;q}(u,\ell)$ and
\begin{equation}
R(u,\ell) = -\frac{2 \ell^2 (\alpha p + 1 + \ell) \tbrac{2u^2 + (2 \alpha p + 3) u + \alpha p + 1}}{u (u-\ell) (u-\ell+1)}.
\end{equation}
Summing this recursion relation over $\ell$ then yields \eqref{eq:Recurse}, upon noting that $(\alpha p + 1)^2 = 4t \Delta_p + 1$. Since $f_{p;q}(0) = 0$ and $f_{p;q}(1) = 1$, it follows from \eqref{eq:Recurse} that $f_{p;q}(u)$ is a polynomial in $\lambda_{p;q}$ and $\Delta_p$, as claimed.
\end{proof}
Let $g_{u,v}(\lambda,\Delta)$ denote the polynomial for which $f_{p;q}(u) = g_{u,v}(\lambda_{p;q}, \Delta_p)$ and let
\begin{equation}
I_{u,v}(\lambda,\Delta)=\prod_{(r,s)\in K(u,v)}\left(\Delta-\Delta_{r,s}\right)\cdot g_{u,v}(\lambda,\Delta).
\end{equation}
It is a simple corollary of \eqref{eq:SumFactorExs} and \eqref{eq:Recurse} that $g_{u,v}$ has degree $u-1$ as a polynomial in $\lambda$. If we regard $\Delta$ as having degree $2$, then the total degree of $g_{u,v}$ is also $u-1$ and that of $I_{u,v}$ is therefore $(u-1)v$.
\begin{thm} \label{thm:ComputeZhu}
Zhu's algebra of \(\AdmMod{u}{v}\) is given by the quotient
\begin{align}
\zhu{\AdmMod{u}{v}}=\frac{\UEA\:\SLA{sl}{2}}{\corrfn{I_{u,v}(h,T)}},
\end{align}
where $T = \frac{1}{2t} (\frac{1}{2} h^2 - ef - fe)$ denotes the image of $T(z)$ in $\zhu{\UVOA{k}} \cong \UEA\:\SLA{sl}{2}$.
\end{thm}
\begin{proof}
As noted above, the ideal of $\UEA\:\SLA{sl}{2}$ by which one quotients to get the Zhu algebra $\zhu{\AdmMod{u}{v}}$ is generated by the image of the null field $\chi(z)$. This image is a polynomial in \(e\), \(f\) and \(h\) of total degree at most \((u-1)v\). We have evaluated the action of the zero mode of \(\chi(z)\) on a continuum of relaxed \hwvs{}{} and thus the image of \(\chi(z)\) in Zhu's algebra is, up to non-zero constant factors, equal to the polynomial \(I_{u,v}(h,T)\) of total degree $(u-1)v$.
\end{proof}
Before we can use this presentation of Zhu's algebra of \(\AdmMod{u}{v}\), we need to know a little more about the zeroes of the polynomial \(g_{u,v}\).
\begin{prop}\label{prop:zerosofg}
For each $u \in \mathbb{Z}_{\ge 1}$, the polynomials $g_{u,v}(\lambda, \Delta_{r,0})$ evaluate to zero when $r = 1, 2, \ldots, u-1$ and $\lambda = r-1, r-3, \ldots, -r+3, -r+1$.
\end{prop}
\begin{proof}
This is trivial for $u=1$ as there are then no $r$ or $\lambda$ to check. For $u=2$, \eqref{eq:SumFactorExs} gives $g_{2,v}(\lambda, \Delta) = \lambda$ and we need only check $r=1$ and $\lambda = 0$. We may therefore assume, inductively, that the statement of the proposition is true for $g_{1,v}, g_{2,v}, \ldots, g_{u+1,v}$. Then, the recursion relation \eqref{eq:Recurse} shows that
\begin{equation}
g_{u+2,v}(\lambda, \Delta_{r,0}) = \frac{(2u+1) \lambda}{(u+1)^2} g_{u+1,v}(\lambda, \Delta_{r,0}) - \frac{4t \Delta_{r,0} - (u-1)(u+1)}{(u+1)^2} g_{u,v}(\lambda, \Delta_{r,0})
\end{equation}
will vanish for all $r = 1, 2, \ldots, u-1$ and $\lambda = r-1, r-3, \ldots, -r+3, -r+1$, because $g_{u+1,v}$ and $g_{u,v}$ do. Moreover, because $4t \Delta_{u,0} = (u-1)(u+1)$, $g_{u+2,v}$ also vanishes for $r=u$ and $\lambda = r-1, r-3, \ldots, -r+3, -r+1$.
The only remaining case is $r=u+1$. Then, $4t \Delta_{u+1,0} = u(u+2)$, hence we may identify $\alpha p$ with $u$ or $-u-2$, hence $q = \frac{1}{2} (\lambda - \alpha p)$ with $\frac{1}{2} (\lambda - u)$ or $\frac{1}{2} (\lambda + u+2)$, respectively. From \eqref{eq:DefF} and $\alpha p = u$, we now obtain
\begin{align}
g_{u+2,}(\lambda, \Delta_{u+1,0}) &= f_{u/\alpha; (\lambda - u)/2}(u+2) = \sum_{\ell=0}^{u+1} \binom{\frac{1}{2} (\lambda - u) - 1}{\ell} \binom{u+1+\ell}{\ell} \binom{2(u+1)}{u+1-\ell} \notag \\
&= \binom{2(u+1)}{u+1} \sum_{\ell=0}^{u+1} \binom{\frac{1}{2} (\lambda - u) - 1}{\ell} \binom{u+1}{u+1-\ell} = \binom{2(u+1)}{u+1} \binom{\frac{1}{2} (\lambda + u)}{u+1} \notag \\
&= \binom{2(u+1)}{u+1} \frac{(\lambda + u) (\lambda + u-2) \cdots (\lambda - u+2) (\lambda - u)}{2^{u+1} (u+1)!},
\end{align}
which clearly vanishes for $\lambda = u, u-2, \ldots, -u+2, -u$. The result is the same for $\alpha p = -u-2$.
\end{proof}
We are now in a position to classify the simple weight modules over \(\zhu{\AdmMod{u}{v}}\). Recall that \(\zhu{\AdmMod{u}{v}}\)-modules are automatically $\SLA{sl}{2}$-modules; by a weight module over \(\zhu{\AdmMod{u}{v}}\), we mean that it is a weight module over $\SLA{sl}{2}$. The classification of simple \(\SLA{sl}{2}\) weight modules was summarised in \propref{prop:FinSL2WtMods}. We also recall that the quadratic Casimir operator
\begin{equation}
Q=\frac{1}{2}h^2-ef-fe=2tT
\end{equation}
acts as a scalar multiple of the identity on any simple \(\SLA{sl}{2}\) weight module.
\begin{thm}\label{thm:ZhuClassification}
The following $\SLA{sl}{2}$-modules provide a complete list of the inequivalent isomorphism classes of simple weight modules of \(\zhu{\AdmMod{u}{v}}\):
\begin{itemize}
\item The finite-dimensional highest and \lw{} modules{} \(\FinIrr{\lambda_{r,0}}\), where \(1\le r\le u-1\).
\item The infinite-dimensional \hw{} modules{}
\(\FinDisc{\lambda_{r,s}}\), where \(1\le r\le u-1\) and \(1\le s\le v-1\).
\item The infinite-dimensional \lw{} modules{} \(\finconjmod{\FinDisc{\lambda_{r,s}}}\),
where \(1\le r\le u-1\) and \(1\le s\le v-1\).
\item The infinite-dimensional \rhw{} modules{} \(\FinRel{\lambda,\Delta_{r,s}}\), where
\((r,s)\in K(u,v)\) and \(4t\Delta_{r,s}\neq \mu(\mu+2)\) for all \(\mu\in \lambda+2\mathbb{Z}\).
\end{itemize}
\end{thm}
\begin{proof}
We first consider a simple finite-dimensional \(\zhu{\AdmMod{u}{v}}\)-module \(M\), which must therefore also be a finite-dimensional \(\SLA{sl}{2}\)-module. As the quadratic Casimir takes the value \(\frac{1}{2}(r^2-1)=2t \Delta_{r,0}\) on the \(r\)-dimensional simple \(\SLA{sl}{2}\) module, it follows that \(g_{u,v}(h,T)\) must act trivially on \(M\) because the remaining \(\lambda\)-independent factors of \(I_{u,v}(\lambda,\Delta)\) do not have \(\Delta_{r,0}\) as a root for any positive integer \(r\). By \propref{prop:zerosofg}, \(g_{u,v}(m,\Delta_{r,0})=0\) if \(1\le r\le u-1\) and \(m=r-1,r-3,\dots, -r+1\). Conversely if \(r\ge u\), then \(g_{u,v}(m,\Delta_{r,0})\neq0\) for some \(m=r-1,r-3,\dots, -1+r\), because the \(\lambda\)-degree of \(g_{u,v}(\lambda,\Delta_{r,0})\) is \(u-1\), so there cannot be more than \(u-1\) zeroes. Thus, \(M\) must be isomorphic to one of the \(\FinIrr{\lambda_{r,0}}\) for some \(1\le r\le u-1\).
Next, we consider a simple infinite-dimensional \(\zhu{\AdmMod{u}{v}}\)-module \(M\), which must therefore also be an infinite-dimensional \(\SLA{sl}{2}\) weight module. Because there must be an infinite number of weight vectors in \(M\) with distinct \(\SLA{sl}{2}\)-weights, \(g_{u,v}(h,T)\) cannot vanish identically on \(M\). In order for \(I_{u,v}(h,T)\) to then vanish, \(T\) must act as multiplication by \(\Delta_{r,s}\) for some \((r,s)\in K(u,v)\). Referring to \propref{prop:FinSL2WtMods}, it follows that the last three cases of \thmref{thm:ZhuClassification} exhaust all the possible isomorphism classes for \(M\).
\end{proof}
\noindent The correspondence (\thmref{thm:ZhuSimples}) between simple modules of a vertex operator algebra{} and its Zhu algebra then proves the following classification result:
\begin{thm} \label{thm:IrrRelMods}
The following $\AKMA{sl}{2}$-modules provide a complete list of the inequivalent isomorphism classes of simple \rhw{} modules{} of \(\AdmMod{u}{v}\):
\begin{itemize}
\item The \hw{} modules{} \(\Irr{\lambda_{r,0}}\), where \(1\le r\le u-1\).
\item The \hw{} modules{} \(\Disc{\lambda_{r,s}}\), where \(1\le r\le u-1\) and \(1\le s\le v-1\).
\item The conjugates \(\conjmod{\Disc{\lambda_{r,s}}}\), where \(1\le r\le u-1\) and \(1\le s\le v-1\).
\item The \rhw{} modules{} \(\Rel{\lambda,\Delta_{r,s}}\), where \((r,s)\in K(u,v)\) and \(4t\Delta_{r,s}\neq \mu(\mu+2)\) for all \(\mu\in \lambda+2\mathbb{Z}\).
\end{itemize}
\end{thm}
\noindent These are therefore the simple modules of the vertex operator algebra{} $\AdmMod{u}{v}$ which belong to the category $\categ{R}$ that was introduced in \secref{sec:Relaxed}.
Zhu's correspondence also extends to non-simple $\zhu{\AdmMod{u}{v}}$-modules. In particular, $\SLA{sl}{2}$ admits reducible, but indecomposable, modules similar to the $\FinRel{\lambda; \Delta}$ of \propref{prop:FinSL2WtMods} whenever
\begin{equation} \label{eq:IndecCond}
4t \Delta = \mu(\mu+2) \quad \text{for some} \quad \mu \in \lambda+2\mathbb{Z}.
\end{equation}
For $\Delta = \Delta_{r,s}$, where $(r,s) \in K(u,v)$, the only solutions are $\mu = r-1-ts = \lambda_{r,s}$ and $\mu = -r-1+ts = \lambda_{u-r,v-s}$. As $0<s<v$, we find that $\lambda_{r,s} - \lambda_{u-r,v-s} \notin 2 \mathbb{Z}$, concluding that an indecomposable with $\Delta = \Delta_{r,s}$ may have at most one weight $\mu$ satisfying \eqref{eq:IndecCond}. Thus, there are \cite{MazLec10} precisely two reducible, but indecomposable, $\SLA{sl}{2}$-modules $\FinRel{\lambda_{r,s}; \Delta_{r,s}}^+$ and $\FinRel{\lambda_{r,s}; \Delta_{r,s}}^-$ for each $1 \le r \le u-1$ and $1 \le s \le v-1$. They are determined (up to isomorphism) by the following non-split short exact sequences:
\begin{equation}
\dses{\FinDisc{\lambda_{r,s}}}{}{\FinRel{\lambda_{r,s},\Delta_{r,s}}^+}{}{\finconjmod{\FinDisc{\lambda_{u-r,v-s}}}}, \qquad
\dses{\finconjmod{\FinDisc{\lambda_{u-r,v-s}}}}{}{\FinRel{\lambda_{r,s},\Delta_{r,s}}^-}{}{\FinDisc{\lambda_{r,s}}}.
\end{equation}
Applying Zhu's construction now leads to the following result:
\begin{thm} \label{thm:IndecRelMods}
For each $1 \le r \le u-1$ and $1 \le s \le v-1$, there exist two reducible, but indecomposable, $\AdmMod{u}{v}$-modules $\Rel{\lambda_{r,s}; \Delta_{r,s}}^+$ and $\Rel{\lambda_{r,s}; \Delta_{r,s}}^-$, obtained by inducing $\FinRel{\lambda_{r,s}; \Delta_{r,s}}^+$ and $\FinRel{\lambda_{r,s}; \Delta_{r,s}}^-$ to level $k$ $\AKMA{sl}{2}$-modules and quotienting by the sum of all the submodules that trivially intersect the space of conformal weight $\Delta_{r,s}$. They are determined (up to isomorphism) by the following non-split short exact sequences:
\begin{equation}
\dses{\Disc{\lambda_{r,s}}}{}{\Rel{\lambda_{r,s},\Delta_{r,s}}^+}{}{\conjmod{\Disc{\lambda_{u-r,v-s}}}}, \qquad
\dses{\conjmod{\Disc{\lambda_{u-r,v-s}}}}{}{\Rel{\lambda_{r,s},\Delta_{r,s}}^-}{}{\Disc{\lambda_{r,s}}}.
\end{equation}
\end{thm}
A theorem of Kac and Wakimoto \cite[Prop.~1]{KacMod88} asserts that the $\AdmMod{u}{v}$-modules in category $\categ{O}$ are all semisimple. The category $\categ{O}$ $\AdmMod{u}{v}$-modules therefore consist of finite direct sums of the \hw{} modules{} of \thmref{thm:IrrRelMods}. By way of contrast, a corollary of \thmref{thm:IndecRelMods} is that the $\AdmMod{u}{v}$-modules of category $\categ{R}$ need not be semisimple (when $v \neq 1$). We remark that we have not excluded the possibility that there exist \rhw{} modules{} over $\AdmMod{u}{v}$ that extend the $\Rel{\lambda, \Delta_{r,s}}$, or the $\Rel{\lambda_{r,s}, \Delta_{r,s}}^{\pm}$, non-trivially; this would seem to require more information about the submodule structure of the relaxed Verma modules than is currently available, see \cite{FeiEqu98,SemEmb97}. However, the highest weight{} result given in \corref{cor:ClassHWMods} and the analogous results for the Virasoro minimal models suggest the following conjecture:
\begin{conj}
The $\AdmMod{u}{v}$-modules of \thmref{thm:IrrRelMods} and \thmref{thm:IndecRelMods} exhaust the indecomposable $\AdmMod{u}{v}$-modules of category $\categ{R}$.
\end{conj}
We close by demonstrating that the non-semisimplicity of \rhw{} modules{} over $\AdmMod{u}{v}$ does not imply that the Virasoro mode $L_0$ acts non-semisimply, a fact that is of interest to logarithmic conformal field theory{} studies. This result requires the finite-dimensionality of the weight spaces of the category $\categ{R}$ modules, discussed in \secref{sec:Relaxed}.
\begin{thm} \label{thm:L0ActsSemisimply}
\leavevmode
\begin{enumerate}
\item The image of the energy momentum tensor $T$ in \(\zhu{\AdmMod{u}{v}} \cong \UEA\:\SLA{sl}{2} / \corrfn{I_{u,v}(h,T)}\) acts semisimply on every weight module of \(\zhu{\AdmMod{u}{v}}\) with finite-dimensional weight spaces.
\item The Virasoro zero mode \(L_0\) acts semisimply on every \rhw{} module{} of \(\AdmMod{u}{v}\).
\end{enumerate}
\end{thm}
\begin{proof}
\leavevmode
\begin{enumerate}
\item Let \(\Mod{M}\) be an indecomposable weight module of \(\zhu{\AdmMod{u}{v}}\) on which (the image of) \(T\) acts non-semisimply. As $T$ is proportional to the image of the quadratic Casimir $Q$, it follows that $T$ has a single (generalised) eigenvalue on $\Mod{M}$. Then, by Weyl's theorem for $\SLA{sl}{2}$, \(\Mod{M}\) must be infinite-dimensional with an infinite number of distinct \(\SLA{sl}{2}\) weights. Let \(\Mod{W}\) be the submodule of \(\Mod{M}\) spanned by the eigenvectors of \(T\). As $\Mod{W}$ is non-zero, $\Mod{W}$ must also be infinite-dimensional as it possesses an infinite number of distinct \(\SLA{sl}{2}\)-weights. Thus, \(\Mod{W}\) is an eigenspace of \(T\) with eigenvalue \(\Delta_{r,s}\), for some \((r,s)\in K(u,v)\).
Next, assume that there exists a generalised eigenvector \(v\) of \(T\), so that $(T-\Delta_{r,s})v\neq0$, with \(\SLA{sl}{2}\)-weight \(\lambda_v\). The existence of \(v\) would imply that \(I_{u,v}(\lambda_v, \Delta)\) has a zero of order at least $2$ at \(\Delta=\Delta_{r,s}\). Since there are only finitely many \(\SLA{sl}{2}\)-weights \(\lambda\) for which \(\Delta=\Delta_{r,s}\) is a zero of \(I_{u,v}(\lambda, \Delta)\) of order at least $2$, the quotient \(\Mod{M}/\Mod{W}\) must be finite-dimensional. But, the eigenvalue of $T$ on $v+\Mod{W}$ is then \(\Delta_{r,0}\), for some \(1\le r\le u-1\), which is a contradiction. It follows that no such generalised eigenvectors exist, hence that $T$ acts semisimply on $\Mod{M}$.
\item On any \rhw{} module{} over $\AdmMod{u}{v}$, the action of $L_0$ on the relaxed \hwvs{}{} coincides with that of $T$ on the corresponding $\zhu{\AdmMod{u}{v}}$-module. As the latter action is semisimple, so is that of $L_0$ on the relaxed \hwvs{}{}. As these generate the whole module, $L_0$ acts semisimply. \qedhere
\end{enumerate}
\end{proof}
We stress that this result does not imply that the conformal field theories{} corresponding to the vertex operator algebras{} $\AdmMod{u}{v}$ are non-logarithmic. Indeed, it has been known for some time that there are models with $k=-\tfrac{4}{3}$ \cite{GabFus01} and $k=-\tfrac{1}{2}$ \cite{LesLog04,RidFus10} that are logarithmic. The loophole is that we may also twist by the so-called spectral flow automorphisms which, when $v \neq 1$, lead to infinitely many new simple (and indecomposable) $\AdmMod{u}{v}$-modules that do not belong to category $\categ{R}$. For $k=-\tfrac{4}{3}$ and $k=-\tfrac{1}{2}$, there exist indecomposable $\AdmMod{u}{v}$-modules that are formed from \rhw{} modules{} from different spectral flow sectors and the action of $L_0$ on these is non-semisimple. In fact, these modules are \emph{staggered} in the sense of \cite{RidSta09,CreLog13}; a detailed discussion may be found in \cite{CreMod12}. We expect that there exist staggered $\AdmMod{u}{v}$-modules whenever $v \neq 1$, hence that the associated conformal field theories{} are logarithmic, and hope to report on this in the future.
\section*{Acknowledgements}
We thank Jim Borger for help with a question of commutative algebra, J\"{u}rgen Fuchs, Masoud Kamgarpour and Christoph Schweigert for illuminating discussions regarding parabolic Verma modules, Antun Milas for correspondence concerning the current status of higher rank generalisations, Ole Warnaar for advice on symmetric function theory, and the organisers of the Erwin Schr\"{o}dinger Institute programme ``Modern trends in topological quantum field theory'' for their hospitality.
DR's research is supported by the Australian Research Council Discovery Project DP1093910.
SW's work is supported by the Australian Research Council Discovery Early Career Researcher Award DE140101825.
|
\section{Introduction}
Wigner-Dyson universality of the local correlation of energy levels
among various stochastic \cite{Meh04} and quantum-chaotic systems \cite{BGS84}
under well-defined conditions was established through the ten-fold classification of
symmetric spaces of spectral $\sigma$ models \cite{Zir96},
to which the Gutzwiller trace formula also reduces \cite{MHABH09}.
This universality has in turn provided a solid and secure ground
on which system-specific information can be decoded
by measuring deviations of spectral correlation functions from their universal forms,
or transition between two universality classes.
Prime examples of the former are the weak localization correction in
Anderson Hamiltonians \cite{Efe97} and
the nonuniversal effect of short periodic orbits (small primes) in chaotic systems
(in the Riemann $\zeta$ zeroes \cite{BK99}).
Study on the latter ``universality crossover", initiated by Dyson \cite{Dys62},
has also come to encompass a variety of settings,
an example being the GUE-GOE transition that appears in
a disordered ring \cite{DM91} and chaotic systems \cite{SNMB09} both under magnetic fields.
Recent years saw applications of the universality crossover in lattice QCD,
in an effort to explore the effects of the isospin chemical potential \cite{DHSS05}
and of the finite lattice spacing in the Wilson Dirac operator \cite{DSV10}.
These studies have revealed the power of the spectral approach
in determining the pion decay constant and the Wilsonian chPT constants
from relatively small lattices.
The aim of this work is to apply this approach to the determination of low-energy constants
in another setting, namely the two-color QCD subjected under the imaginary chemical potential
\cite{Nis12} or coupled to QED.
Our novelty is to employ the individual distributions of small Dirac eigenvalues \cite{DN01}
instead of $n$-level correlation functions, in fitting the lattice data.
Practical advantages of our method will be manifested subsequently.
\section{chGSE-chGUE crossover}
Let $\bs{A}$ and $\bs{B}$ be $N/2\times N'/2$ quaternion matrices,
represented by complex $N\times N'$ matrices as
\begin{equation}
A=\sum_{\mu=0}^3 \(A^{(\mu)}_{jk}\) \otimes {\bf e}_\mu,\ \
B=\sum_{\mu=0}^3 \(B^{(\mu)}_{jk}\) \otimes {\bf e}_\mu\ \
(j=1,\ldots,N/2,\ k=1,\ldots,N'/2).
\end{equation}
Here a set of four $2\times2$ matrices ${\bf e}_\mu=(\mathbf{1}_2, -i\vec{\sigma})$
spans the basis of the quaternion field $\mathbb{H}$.
Let the matrix elements belong to
$
A^{(\mu)}_{jk}\in \mathbb{R}
$
and
$
B^{(\mu)}_{jk}\in \mathbb{C},
$
so that the matrix $\bs{A}$ is quaternion-real and $\bs{B}$ is not (i.e. a generic $N\times N'$ complex matrix).
We consider
$A^{(\mu)}_{jk}$, ${\rm Re}\, B^{(\mu)}_{jk}$, and ${\rm Im}\, B^{(\mu)}_{jk}$
to be independent random variables distributed according to
the Gaussian distributions
${\rm e}^{-\frac12{\rm tr}\, A A^\dagger}$ and ${\rm e}^{-{\rm tr}\, B B^\dagger}$,
respectively, and introduce
an ensemble of $(N+N') \times (N+N') $ Hermitian matrices $H$ of the form
\begin{equation}
H=
\left(
\begin{array}{cc}
\mathbf{0}_{N\times N} &C\\
C^{\dagger} &\mathbf{0}_{N'\times N'}
\end{array}
\right),
\ \ C={\rm e}^{-\tau}A+\sqrt{1-{\rm e}^{-2\tau}} B.
\label{H}
\end{equation}
Here a real parameter $\tau$ plays the role of
fictitious time for the Brownian motion
of the eigenvalues \cite{Dys62}.
This ensemble enjoys
the chiral symmetry $\{H,\gamma_5\}=0$ with $\gamma_5={\rm diag}(\mathbf{1}_{N},-\mathbf{1}_{N'})$,
implying that the spectrum of $H$ consists of $N$
$\pm$ pairs of nonzero eigenvalues and $\nu=|N'-N|$ zero eigenvalues.
The presence of $B$ violates the quaternion-reality of $C$ and the selfduality of $H$,
lifting the Kramers degeneracy of nonzero eigenvalues of $H$.
Accordingly this ensemble interpolates the two limiting cases,
chiral GSE at $\tau=0$ and chiral GUE at $\tau\to\infty$,
depending on a single parameter $\tau$.
We consider the case in which the Kramers degeneracy is weakly broken by $\tau\ll 1$.
Then the spectral density of $H$ in the large-$N$ limit
is identical to that of the chGSE ($\tau=0$), i.e. Wigner's semi-circle
$\bar{\rho}(\lambda)=\sqrt{4N-\lambda^2}/\pi$.
We magnify the vicinity of the origin of the $\lambda$ axis by introducing
unfolded variables $x_i=\lambda_i/\varDelta$
with $\varDelta=1/\bar{\rho}(0)=\pi/\sqrt{4N}$.
In order to realize a nontrivial crossover behavior,
we take the triple-scaling limit $N, N'\to\infty, \lambda_i\to 0, \tau\to 0$
while keeping the combinations $\rho=\sqrt{\tau}/\varDelta$, $\nu=N'-N(\geq 0)$, and $x_i$ fixed finite.
Then the j.p.d.~of $N$ positive unfolded eigenvalues $P_N(x_1,\ldots,x_N)$ is expressed
as a Pfaffian of the dynamical Bessel kernel $K(x,y)$ \cite{FNH99},
\begin{eqnarray}
&&
P_N(x_1,\ldots,x_N)={\rm Pf}\(Z \left[K(x_i, x_j)\right]_{i,j=1}^N\),
\
K(x, y)=\left[
\begin{array}{cc}
S(x,y) & I(x, y)\\
D(x,y) & S(y, x)
\end{array}
\right],\
Z=\left[
\begin{array}{cc}
0 & 1 \\
-1 & 0
\end{array}
\right]\otimes \mathbf{1},
\label{Pfaffian}
\\
&&S(x,y)=\pi \sqrt{x y} \left\{
\frac{J_{\nu}(\pi x) y J_{\nu-1}(\pi y)-x J_{\nu-1}(\pi x) J_\nu(\pi y)}{x^2-y^2}
-\frac{J_\nu(\pi x)}{2}\!\!\int_0^\pi d\upsilon\,{\rm e}^{\rho^2 (\upsilon^2-\pi^2)}J_\nu(\upsilon y)\right\}
,\nonumber\\
&&D(x,y)=
\frac{\sqrt{xy} }{2} \int_0^\pi d\upsilon\,\upsilon \int_0^1 du\,{\rm e}^{\rho^2 \upsilon^2(1+u^2)}
\left\{J_\nu(\upsilon u x) J_\nu(\upsilon y)- J_\nu(\upsilon x) J_\nu(\upsilon u y)\right\}
,\nonumber\\
&&I(x,y)=
\frac{ \sqrt{x y}}{2} \int_\pi^\infty d\upsilon\, \upsilon^2 \,{\rm e}^{-2 \rho^2 \upsilon^2}
\left\{J_{\nu}(\upsilon x) y J_{\nu-1}(\upsilon y)- x J_{\nu-1}(\upsilon x) J_{\nu}(\upsilon y)\right\}.
\nonumber
\end{eqnarray}
Due to the recursion relation
$\int_0^\infty dx_k {\rm Pf}\(Z\left[K(x_i, x_j)\right]_{i,j=1}^k \)=(N-k+1){\rm Pf}\(Z\left[K(x_i, x_j)\right]_{i,j=1}^{k-1}\),$
correlation functions of $n$ eigenvalues are given by
\begin{equation}
R_n(x_1,\ldots,x_n)=
\frac{N!}{(N-n)!}
\int_0^\infty dx_{n+1}\ldots dx_N\,P_N(x_1,\ldots,x_N)
={\rm Pf}\(Z \left[K(x_i, x_j)\right]_{i,j=1}^n\).
\label{Pfaffian2}
\end{equation}
\section{Individual eigenvalue distributions}
The Pfaffian forms in (\ref{Pfaffian})$\sim$(\ref{Pfaffian2})
originate from quaternion determinants (Tdet)
composed of a quaternionic kernel, $\left[\mathcal{K}(x_i, x_j)\right]_{i,j}$,
whose $\mathbb{C}$-number representative is
the antisymmetric matrix $Z \left[{K}(x_i, x_j)\right]_{i,j}$.
Accordingly,
the probability $E_k(s)$ for an interval $[0, s]$ to contain exactly $k$ eigenvalues is also given
in terms of the Fredholm Tdet of a quaternionic integral operator $\hat{{\cal K}}_s$,
i.e.\ the square root of the corresponding Fredholm determinant of $\hat{K}_s$
(i.e.~Fredholm Pfaffian of $Z\hat{K}_s$),
\begin{equation}
E_k(s)=
\frac{1}{k!} (-\partial_\xi)^k \left. {\rm Det} (1-\xi \hat{K}_s)^{1/2}\right|_{\xi=1}.
\label{FredholmPfaffian}
\end{equation}
Here $\hat{K}_s$ denotes an integral operator with the dynamical Bessel kernel $K(x,y)$ (\ref{Pfaffian})
acting on the space of two-component $L^2$-functions over the interval $[0,s]$.
First few $E_k(s)$'s are expressed as
\begin{eqnarray}
&&E_0(s)={{\rm Det} (1-\hat{K}_s)^{1/2}},\quad
E_1(s)=E_0(s)\frac{T_1}{2},\quad
E_2(s)=E_0(s)\frac{1}{2!} \left(\frac{T_1^2}{4}-\frac{T_2}{2}\right),
\label{Ek}\\
&&E_3(s)=E_0(s)\frac{1}{3!} \left(\frac{T_1^3}{8}-\frac34 T_1 T_2+T_3\right),\quad
E_4(s)=E_0(s)\frac{1}{4!} \left(\frac{T_1^4}{16}- \frac34 T_1^2 T_2+ \frac34 T_2^2+2T_1 T_3-3T_4\right),
\nonumber
\end{eqnarray}
where
$T_n(s)={\rm Tr} \bigl(\hat{K}_s(I-\hat{K}_s)^{-1}\bigr)^n$
denote functional traces of the resolvents of $\hat{K}_s$.
Probability distribution $p_k(s)$ of the $k^{\rm th}$ smallest positive eigenvalue is then given as
$p_k(s)=-\partial_s \sum_{\ell=0}^{k-1} E_{\ell}(s).$
An efficient way of numerically evaluating the Fredholm determinant of
a trace-class operator $\hat{K}_s$ acting on $L^2$-functions over an interval $[0,s]$
is the Nystr\"{o}m-type discretization \cite{Bor10}
\begin{equation}
{\rm Det}(1-\hat{K}_s)\simeq \det(I-\mathbf{K}_s),\ \
\mathbf{K}_s= \left[K(x_i,x_j) \sqrt{w_i\,w_j}\right]_{i,j=1}^m ~.
\label{Nystrom}
\end{equation}
\begin{figure}[b]
\begin{center}
\includegraphics[bb=0 0 259 165,width=74mm]{fig1a.pdf}
\includegraphics[bb=0 0 258 165,width=73.3mm]{fig1b.pdf}
\caption{
First four eigenvalue distributions $p_{1\sim 4}(s)$ (left) for $0.04\leq \rho \leq 0.70$ (step 0.01, purple to red)
and the spectral density $R_1(x)$ for $0.01\leq \rho \leq 1.00$ (step 0.01)
for the chGSE (black) to chGUE (grey) crossover.}
\end{center}
\end{figure}
Here we employ a quadrature rule consisting of a set of points $\{x_i\}$
taken from the interval $[0,s]$ and associated weights $\{w_i\}$ such that
${\int_0^s f(x)dx \simeq \sum_{i=1}^m f(x_i) w_i}$.
Similarly, the resolvents in (\ref{Ek}) are approximated as
$T_n(s)\simeq {\rm tr} \left(\mathbf{K}_s(I-\mathbf{K}_s)^{-1}\right)^n.$
For a practical purpose we choose the Gauss quadrature rule, i.e.\ sampling $\{x_i\}$ from
the nodes of Legendre polynomials normalized to $[0,s]$.
Previously we applied the Nystr\"{o}m-type method to the dynamical Bessel kernels
interpolating chGSE-chGUE (\ref{Pfaffian}) and chGOE-chGUE
and evaluated the smallest eigenvalue distributions $p_1(s)$ \cite{Nis12}.
In this work we extend our computation to the first four eigenvalues, aiming to reduce the fitting errors
in determining the low-energy constants.
We set the approximation order $m$ to be
at least 100, and confirmed the stability of the results for increasing $m$ (up to $200\sim 400$).
The distributions
$p_1(s), \cdots, p_4(s)$ for the $\nu=0$ case,
computed from (\ref{Pfaffian})$\sim$(\ref{Nystrom})
for $\rho \leq 0.70$ are exhibited in Fig.~1L.
A practical advantage of using individual eigenvalue distributions
over the spectral density $R_1(x)=\sum_{k=1}^\infty p_k(x)=S(x,x)$ (Fig.~1R)
for fitting the lattice data is clear from the figure:
the oscillation of the latter immediately becomes structureless and insensitive to
the interpolation parameter $\rho$ due to the overlapping of multiple peaks of the former, whereas
the quasi-Gaussian shape of each peak is clearly distinguishable and is extremely sensitive to $\rho$.
Another advantage specific to the current case originates from the fact that
$p_{2k-1}(s)$ and $p_{2k}(s)$ move in opposite directions as $\rho$ is increased to break the
Kramers degeneracy. By combining the two best-fitting values of $\rho$ for these two distributions,
any error present in the mean level spacing $\varDelta$ of the Dirac spectrum,
which would result in shifting the unfolded data of $2k-1^{\rm th}$ and $2k^{\rm th}$ eigenvalues
to the same direction, is expected to be cancelled.
We have confirmed this by generating $10^5$ samples of
crossover random matrix ensembles with $N=N'=6^4$ and various $\rho\leq 0.50$ and
by fitting histograms of first four eigenvalues to the analytic results.
Combined values of $\rho$ from these four fittings have reproduced the true input values
within a few per mil of systematic error
(max.~0.5\%), an order of magnitude closer to the input values than using any single individual distribution.
Such an accuracy could neither be hoped for had we used the spectral density $R_1(x)$ for fitting.
\section{Effective theory and low-energy constants}
The Dirac operator $D\!\!\!\!\!/$ of a QCD-like theory with
quarks in a pseudoreal (real) representation,
such as the fundamental of Sp$(2N)$ (SO$(N)$),
possesses an antiunitary symmetry unlike
QCD with quarks in a complex representation \cite{Ver94}:
$D\!\!\!\!\!/$ commutes with ${\cal C}Z \ast$ (${\cal C}\ast$),
with ${\cal C}$ being the charge conjugation and $\ast$ the complex conjugation.
As $({\cal C}Z\ast)^2=+1$ ($({\cal C}\ast)^2=-1$), $D$ can be brought to a real symmetric
(quaternion selfdual) matrix by a similarity transformation.
Due to this property, the distinction between
left-handed quarks and conjugated right-handed quarks is lost,
leading to the Pauli-G\"{u}rsey extension of the flavor symmetry from ${\rm SU}(N_F)_L\times {\rm SU}(N_F)_R$
to ${\rm SU}(2N_F)=:G$ and its vector subgroup
from ${\rm SU}(N_F)_V$ to ${\rm Sp}(2N_F)$ or ${\rm SO}(2N_F)=:H$.
Accordingly its low-energy effective theory becomes a nonlinear
$\sigma$ model on an exotic Nambu-Goldstone manifold $G/H$.
Since the Dirac operator charged under the ${\rm U(1)}$ gauge field is complex,
coupling QCD-like theories with electromagnetism or even
subjecting them to the constant ${\rm U(1)}$ background
breaks the antiunitary symmetry of $D\!\!\!\!\!/$ and the Pauli-G\"{u}rsey extended flavor symmetry.
In the latter case that is equivalent to putting on a weak imaginary chemical potential $\mu=i \mu_I$,
its effect on the low-energy Lagrangian is systematically incorporated by the
flavor covariantization of the derivatives \cite{KSTVZ00}.
Furthermore, if the theory is in a finite volume $V=L^4$ and the
Thouless energy $E_c\simeq {{F^2}/{\Sigma L^2}}$ is much larger than $m$,
the path integral is dominated by the zero-mode integration (the $\varepsilon$ regime),
\begin{equation}
Z=\int_{{\rm SU}(2N_F)} \!\!\!\!\!\!\!\!\!\!\!\!\! dU\,\,\,\exp\(
\frac12 V\Sigma m \,{\rm Re}\,{\rm tr}\, \hat{M} U
-V\mu_I^2 F^2\,{\rm tr}\, (\hat{B} U^\dagger\hat{B} U+\hat{B}\hat{B})
\).
\label{Zchiral}
\end{equation}
Here $U$ is an SU($2N_F$) matrix-valued Nambu-Goldstone field,
$\hat{B}=\sigma_3 \otimes \mathbf{1}_{N_F}$,
$\hat{M}=i\sigma_2\otimes \mathbf{1}_{N_F}$\ $\(\sigma_1 \otimes \mathbf{1}_{N_F}\)$
for quarks in a pseudoreal (real) representation.
$\Sigma=\<\bar{\psi}\psi\>/N_F$ denotes the chiral condensate
and $F$ the pseudo-scalar decay constant,
both measured in the chiral and zero-chemical potential limit.
Note that the above 0D $\sigma$ model for the case of fermions in a real representation
can as well be derived from the random matrix ensemble (\ref{H}) through the standard procedure:
(i) introduce $N_F$ species of complex Grassmannian $(N+N')$-vectors $\psi_f, \bar{\psi}_f$
and consider a replicated spectral determinant
$\left<\det (\lambda-H)^{N_F}\right>=
\left<\int d\psi d\bar{\psi}\,{\rm e}^{\sum_f \bar{\psi}_f (\lambda-H) \psi_f}\right>$,
where $\left<\cdots\right>$ denotes averaging over $A$ and $B$,
(ii) perform Gaussian integrations over $A$ and $B$,
(iii) introduce a $2N_F\times 2N_F$-matrix valued Hubbard-Stratonovich variable $Q$ and
open up the 4-fermi term,
(iv) perform Gaussian integrations over $\psi$ and $\bar{\psi}$,
(v) take the aforementioned triple-scaling limit and denote the angular part of $Q$
(not fixed by the large-$N$ saddle point equation) as $U$.
Then the coefficients of the mass and chemical-potential terms are identified as
$V\Sigma m=i\pi x$ and $2VF^2\mu_I^2=\pi^2\rho^2$.
By substituting $m\to i \lambda_{\rm Dirac}$ which turns the QCD partition function
into the Dirac spectral determinant, the former equality provides
the definition of unfolded Dirac eigenvalues $x=\lambda_{\rm Dirac}/\varDelta$
due to the Banks-Casher relation $\Sigma=\pi/\varDelta V$.
The latter equality is used to determine $F^2$ from the slope of the $\mu_I$-$\rho$ plot.
\section{Fitting Dirac spectra of SU(2)$\times$U(1) gauge theory}
As the aim of this work is
to demonstrate the validity and advantage of the method and not to approach the continuum, chiral, or thermodynamic limit,
we chose the simplest possible setting on the lattice side:
(i) generate $10^4$ samples of quenched SU(2)=Sp(2) gauge fields $U_\mu(x)$ on an (intentionally)
small lattice $V=6^4$,
with a plaquette action at $\beta_{{\rm SU(2)}}=6/g_{\rm SU(2)}^2=0\sim 1.75$ (step .25), using
the standard heat-bath/overrelaxation algorithm.
(ii-a) multiply the ${\rm SU(2)}$ fields on temporal links $U_0(x)$ by a constant phase ${\rm e}^{i\mu_I}$
with $\mu_I=0.00524\sim.05240$ (step .00524), or
(ii-b) generate quenched noncompact ${\rm U(1)}$ gauge fields $A_\mu(x)$ under the Coulomb gauge-fixing condition
\cite{BDHIY07} and multiply the ${\rm SU(2)}$ fields $U_\mu(x)$
by $\exp({ie_{{\rm U(1)}} A_\mu(x)})$, with
$e_{{\rm U(1)}}=0.0004\sim.0024$ (step .0004),
(iii) substitute the gauge fields into an unimproved staggered Dirac operator and diagonalize.
Due to the absence of the ${\cal C}$ matrix, the antiunitary symmetries of staggered Dirac operators
are swapped between real and pseudoreal representations \cite{HV95}.
Accordingly, our case with SU(2)$\times$U(1) fundamental fermions indeed
corresponds to the chGSE-chGUE crossover (\ref{H}).
The low-energy constants are determined by the following steps:
(I) fit the histogram of each of the two smallest Dirac eigenvalues (i.e.~four counting the Kramers degeneracy) of
the pure SU(2) case to the rescaled chGSE ($\rho=0$) prediction $p_k(\lambda_k/\varDelta)/\varDelta$
by varying $\varDelta$,
(II) combine two optimal values of $\varDelta$ and their variances to determine $\bar{\varDelta}$
and thus $\Sigma=\pi/\bar{\varDelta} V$,
(III) fit the histogram of each of the four smallest {\em unfolded} Dirac eigenvalues $x_k=\lambda_k/\bar{\varDelta}$
of (a) SU(2)$+\mu_I$ or (b) SU(2)$\times$U(1) case
to the chGSE-chGUE prediction $p_k(x_k)$ by varying $\rho$,
(IV) combine four optimal values of $\rho$ and their variances to determine $\bar{\rho}$ and thus
$F^2\mu_I^2=(\pi^2/2)\bar{\rho}^2/V$.
We first observe that the four values of $\varDelta$ obtained in the step (I) are mutually consistent,
giving rise to combined relative errors in $\Sigma$ that are
extremely small, $\sim 0.1\%$ (Table 1, top).
One-parameter fittings in the steps (I), (III-a), or (III-b) are quite satisfactory,
with $\chi^2/{\rm dof}=0.5\sim 1.5$ for all range of parameters in concern (exemplified in Fig.~2, above).
We also confirmed our expectation that the best-fitting values of $\rho$ for $k=1, 3$ and
those for $k=2, 4$ have a tendency to counter-move,
in favor of cancelling the unfolding ambiguity due to a tiny error within $\varDelta$.
Relative errors in $\bar{\rho}$ are considerably reduced by the combined use of
four individual eigenvalue distributions (Fig.~2 below), and are no larger than $\pm .018$(stat)$\pm .005$(sys).
Linear response of $\bar{\rho}$ on $\mu_I$ or $e_{\rm U(1)}$ is confirmed
for the SU(2)$+\mu_I$ case (Fig.~3, left), and
the pseudo-scalar decay constant $F^2$ at various values of $\beta_{\rm SU(2)}$ is obtained from the slopes
(Table 1, middle).
For the SU(2)$\times$U(1) case,
the coefficients (equivalent of $F^2\mu_I^2$) of the ${\rm tr}\,\hat{B} U^\dagger\hat{B} U$ term in
(\ref{Zchiral}) divided by $e_{\rm U(1)}^2$,
extrapolated to $e_{\rm U(1)}\to 0$
are summarized in Table 1, bottom.
Complete lattice results, and details of analytic and numerical computations presented
in \S {\bf 2} and \S {\bf 3} will be reported in a subsequent publication.
\begin{table}[h]
$\!\!\!\!\!\!\!\!$
\begin{tabular}{l|llllllll}
$\beta_{\rm SU(2)}$ & 0 & 0.25 & 0.50 & 0.75 & 1.00 & 1.25 & 1.50 & 1.75 \\
\hline
$\Sigma\,a^3$
& 1.310(2) & 1.255(2) & 1.199(1) & 1.139(1) & 1.070(1) & .987(1) & .883(1) & .743(1)\\
$F^2 a^2$
& .284(2) & .268(2) & .247(2) & .226(2) & .205(1) & .178(1) & .153(1) & .115(1)\\
$F^2\mu_I^2 a^4/e_{\rm U(1)}^2$
& 220(2) & 198(2) & 186(2) & 163(1) & 145(1) & 123(1) & 99.5(8) & 68.0(6)\\
\end{tabular}
\caption{Chiral condensate $\Sigma$ from quenched SU(2) [top],
pseudo-scalar decay constant $F^2$ from SU(2)+$\mu_I$ [middle],
and an equivalent of $F^2\mu_I^2$ (divided by $e_{\rm U(1)}^2$) from
SU(2)$\times$U(1) [bottom], all in the lattice unit.}
\end{table}
\begin{figure}[ht]
$\!\!\!\!\!\!\!\!\!\!$
\includegraphics[bb=0 0 360 235,width=5.32cm]{fig2a.pdf}
\includegraphics[bb=0 0 360 258,width=4.85cm]{fig2b.pdf}
\includegraphics[bb=0 0 360 258,width=4.85cm]{fig2c.pdf}
$\!\!\!\!\!\!\!\!\!\!\!\!$
\includegraphics[bb=0 0 188 110,width=5.1cm]{fig2d.pdf}~~
\includegraphics[bb=0 0 175 110,width=4.8cm]{fig2e.pdf}~~
\includegraphics[bb=0 0 174 109,width=4.8cm]{fig2f.pdf}
\caption{Histograms of $k^{\rm th}$ unfolded Dirac eigenvalues for $\beta_{\rm SU(2)}=0.25$ and
$e_{\rm U(1)}=0.0008, .0012, .0016$
[above, left to right] and best-fitting $p_k(s)\ (k=1\sim4)$ from the chGSE-chGUE crossover.
The $\rho$ parameter determined for each $k$,
their combined values [dots] and statistical errors [band] are shown below each graph.}
\vspace{5mm}
\centering
\includegraphics[bb=0 0 178 106,width=6cm,clip]{fig3a.pdf}~~~~
\includegraphics[bb=0 0 176 106,width=6cm,clip]{fig3b.pdf}
\caption{Ratios $\bar{\rho}/\mu_I$ for SU(2)+$\mu_I$ at $\beta_{\rm SU(2)}=0.5$ [left] and
$\bar{\rho}/e_{\rm U(1)}$ for SU(2)$\times$U(1) at $\beta_{\rm SU(2)}=0.25$ [right].
}
\end{figure}
|
\section{Introduction}
The Schr\"odinger-Newton (SN) system of equations has been studied in various contexts since its introduction in~\cite{Bon} as a model for self-gravitating (quantum) particles. The single-particle formulation of the problem treats the central body's quantum mechanical distribution as a mass density sourcing the gravitational field (whose influence affects the central body):
\begin{equation}\label{NSet}
\begin{aligned}
i \, \hbar \, \frac{\partial \Psi(\vecbf r,t)}{\partial t} &= -\frac{\hbar^2}{2 \, m} \nabla^2 \Psi(\vecbf r,t) + m\, \Phi(\vecbf r, t) \, \Psi(\vecbf r,t) \\
\nabla^2 \Phi(\vecbf r, t) &= 4 \, \pi \, G \, m \, \Psi^*(\vecbf r, t) \, \Psi(\vecbf r, t)
\end{aligned}
\end{equation}
where $\Psi(\vecbf r, t)$ is the wave function associated with the particle of mass $m$ moving under the influence of the potential energy $\Phi(\vecbf r, t)$.
Regardless of its motivation (and there are a variety of motivations for studying this set -- some of these are described in~\cite{Carlip,Giulini2} and references therein -- we side-step that discussion in the current work, focusing instead on a comparative study of the solutions to~\refeq{NSet} and our proposed modification), much is known about the solutions in both the time-independent~\cite{Moroz, Tod} and time-dependent form~\cite{HarrisonN,Meter, Manfredi} (see also the dissertations~\cite{Harrison,Salzman} for additional background and review).
We propose to augment the gravitational sourcing in~\refeq{NSet} to include the energy density present in the gravitational field itself. The motivation for doing this comes originally from special relativity and the universal coupling of gravity -- if mass density acts as a source, then so can energy density, and the energy density of the gravitational field is available as a source prior to any additional external sources (like the energy density associated with electromagnetic fields, for example). The governing equation for $\Phi$, the gravitational potential energy, is then~\cite{Einstein}:
\begin{equation}\label{EG}
\nabla^2 \Phi(\vecbf r) = \frac{4 \, \pi \, G }{c^2} \, \rho(\vecbf r)\, \Phi(\vecbf r) + \frac{1}{2 \, \Phi(\vecbf r)} \, \nabla \Phi(\vecbf r) \cdot \nabla \Phi(\vecbf r).
\end{equation}
This is the field equation for a scalar $\Phi(\vecbf r)$ that is sourced by a distribution of mass given by the mass density $\rho(\vecbf r)$ and its own energy density (represented by the second term). This field equation was developed originally by Einstein~\cite{Einstein}, and a (special) relativistic version appeared in~\cite{FandN} -- more recently, the self-coupled case was re-derived in~\cite{GiuliniSC}.
If one were to similarly couple a second-rank field theory to itself by making the field's stress tensor a source -- motivated by the universal nature of gravity -- one ends up with general relativity (GR)~\cite{Deser} (and for a review of that process~\cite{Franklin}) together with its geometric interpretation. The sourcing in~\refeq{EG} is a scalar version of that self-coupling (and was worked out in that context in~\cite{DandH}). Our fundamental question will be how this relatively simple nonlinearity, showing up in the gravitational field equation \footnote{Any theory of gravity should include some sort of nonlinearity due to self-sourcing, so it is important to introduce such terms, even at the toy scalar level.} changes the spectrum of the resulting Schr\"odinger-coupled system.
We'll start with a review of the self-coupling that leads to the (static) field theory~\refeq{EG}, then develop the spherically symmetric vacuum solution of~\refeq{EG}, and use that to estimate the ground state energy as a function of particle mass. Then we will introduce a numerical method to calculate the ground state energy of both SN (to test correct numerical behavior) and the new self-coupled gravitational field, and compare the ground state energy to the predicted form. For both the ground state and the first excited state, the Bohr model estimates provide qualitatively correct predictions. By calculating the energy difference between the first excited state and ground state of the self-coupled gravity, we can compare the transition energy emission of this theory with the one predicted by SN. Finally, we calculate the total energy of $N$ self-gravitating bosons using the self-coupled form, and compare that with the total energy as calculated using SN in~\cite{Bon}.
\section{Motivation}
The static gravitational field theory given by~\refeq{EG} was first proposed by Einstein {\it en route} to general relativity. We will generate the field equation here (a recent discussion of this derivation is in~\cite{FranklinAJP}) -- the issue that motivates its development is the lack of self-coupling of the original field equation of Newtonian gravity: $\nabla^2 \Phi = 4 \, \pi \, G \, \rho$ describes a field $\Phi$ sourced by a mass density $\rho$, but this Poisson form lacks the energy density source that comes from $\Phi$ itself.
Let's start by calculating the energy density associated with a field $\Phi$ that comes from Newtonian gravity (so that it satisfies the Poisson equation: $\nabla^2 \Phi = 4 \, \pi \, G \, \rho$). Proceeding as usual, the work done in building a distribution of mass, $\rho$, is given by:
\begin{equation}
W = \frac{1}{2} \, \int_{\hbox{\tiny{all space}}} \rho \, \Phi \, d\tau,
\end{equation}
then using the Poisson equation to eliminate $\rho$ and integrating by parts gives
\begin{equation}
W = \int \of{ -\frac{1}{8 \, \pi \, G} \, \nabla\Phi \cdot \nabla \Phi } \, d\tau,
\end{equation}
and we would call the integrand the energy density of the field:
\begin{equation}\label{ufirstpass}
u_\Phi = -\frac{1}{8 \, \pi \, G} \, \nabla \Phi \cdot \nabla \Phi.
\end{equation}
According to special relativity, we should use an associated effective mass density $\rho_\Phi = u_\Phi/c^2$ as a source in the field equation for $\Phi$. That is the nonlinearity promised by a theory of gravity -- the field's self-energy must act as a source. So we are tempted to start with:
\begin{equation}\label{PetersP}
\nabla^2 \Phi = 4 \, \pi \, G \, \of{\rho + \rho_\Phi} = 4 \, \pi \, G \, \rho - \frac{1}{2 \, c^2} \nabla \Phi \cdot \nabla\Phi,
\end{equation}
and this form was considered in~\cite{Peters}. The problem is that the new field equation~\refeq{PetersP} leads to an energy density (obtained as above, but with the new field equation instead of the Poisson equation) that is {\it not}~\refeq{ufirstpass}, and so the self-coupling is not consistent.
We'll now generate the correct self-coupled theory using the machinery of field theory, where the Hamiltonian density is the energy density (for a static, free field).
Suppose we start with a free field theory that has Hamiltonian $\mathcal{H} = \frac{f(\Phi)}{8 \, \pi \, G} \, \nabla \Phi \cdot \nabla \Phi$, i.e.\ we augment the usual scalar Hamiltonian with a function of $\Phi$, $f(\Phi)$, that we will fix by demanding that it lead to a source term that looks like $\rho_{\mathcal{H}}= \mathcal{H}/c^2$. The field equation for this $\mathcal{H}$ is
\begin{equation}\label{protP}
\frac{f(\Phi)}{4 \, \pi \, G} \, \nabla^2 \Phi + \frac{f'(\Phi)}{8 \, \pi \, G} \, \nabla \Phi \cdot \nabla \Phi = 0,
\end{equation}
or
\begin{equation}
\nabla^2 \Phi = -\frac{f'(\Phi)}{2 \, f(\Phi)} \, \nabla\Phi \cdot \nabla \Phi.
\end{equation}
It is the right-hand side of this equation that we would like to set equal to $4 \, \pi \, G\, \of{\mathcal{H}/c^2} = \frac{f(\Phi)}{2 \, c^2} \, \nabla \Phi \cdot \nabla \Phi$, giving us the self-consistent energy density source, and when we do that, we get an ODE for $f(\Phi)$:
\begin{equation}
\frac{f(\Phi)}{2\, c^2} = -\frac{f'(\Phi)}{2 \, f(\Phi)} \longrightarrow f(\Phi) = \frac{c^2}{\Phi}.
\end{equation}
Our final free Hamiltonian is: $\mathcal H = \frac{c^2}{8 \, \pi \, G \, \Phi} \, \nabla \Phi \cdot \nabla \Phi$, and this is the energy density of the field in the theory, given also in~\cite{Einstein}. Using this solution for $f(\Phi)$ gives the field equation:
\begin{equation}
\nabla^2 \Phi = \frac{1}{2 \, \Phi} \, \nabla\Phi \cdot \nabla \Phi,
\end{equation}
in vacuum, and we recover~\refeq{EG} when we introduce massive sources (that would put a factor of $\rho$ on the right-hand side of~\refeq{protP}). When computing solutions to this self-consistent, self-coupled form of gravity, it is worth noting that the field equation~\refeq{EG} can be written linearly in $\sqrt{\Phi}$, where it reads (again, from~\cite{Einstein}):
\begin{equation}\label{Einsimple}
\nabla^2 \of{ \sqrt{\Phi}} = \frac{2 \, \pi \, G}{c^2} \, \rho \, \of{\sqrt{\Phi}}.
\end{equation}
In order to recover Newtonian gravity, we must take $\Phi = c^2 + \Phi_N$, then inserting this into~\refeq{Einsimple} (or~\refeq{EG}) and collecting in powers of $c$ gives back $\nabla^2 \Phi_N = 4 \, \pi \, G \, \rho$ to zeroth order, a requirement of the weak-field limit.
\section{Point Source Solution}
We'll start by looking at solutions to the field equation:
\begin{equation}\label{Pfield}
\nabla^2 \Phi = \frac{4 \, \pi \, G \, \rho}{c^2} \, \Phi + \frac{1}{2 \, \Phi} \, \nabla \Phi \cdot \nabla \Phi,
\end{equation}
in regions where $\rho = 0$ (so we are in vacuum) and $\Phi(\vecbf r) = \Phi(r)$. Such a solution would be appropriate for a spherically symmetric source of mass $m$ localized near the origin.
Under our assumptions, the field equation reduces to
\begin{equation}
\left( r \, \Phi(r) \right)'' = \frac{r}{2 \, \Phi(r)} \, \left( \Phi'(r) \right)^2
\end{equation}
with primes denoting $r$-derivatives. The solution to this equation comes with two integration constants, $\alpha$ and $\beta$:
\begin{equation}
\Phi(r) = \beta \, \left[ 4 \, \alpha^2 - \frac{4 \, \alpha}{r} + \frac{1}{r^2} \right].
\end{equation}
If we ask that $\Phi(r)$ look like a Newtonian point source (of mass $m$) as $r$ approaches spatial infinity, then we can fix $\alpha$
\begin{equation}\label{prebchoice}
\Phi(r) = -\frac{G \, m}{r} + \frac{\beta}{r^2} + \frac{G^2 \, m^2}{4 \, \beta}.
\end{equation}
To recover the Poisson form of Newtonian gravity for $\Phi$ small (compared to $c^2$), we must have $\Phi(\infty) = c^2$ as a boundary condition, and that sets the constant $\beta$. Our final spherically symmetric vacuum solution looks like
\begin{equation}\label{ppSC}
\Phi(r) = -\frac{G \, m}{r} + \frac{G^2 \, m^2}{4 \, c^2 \, r^2} + c^2.
\end{equation}
The energy density of this field is, from $\mathcal H = \frac{c^2}{8 \, \pi \, G \, \Phi} \, \nabla \Phi \cdot \nabla \Phi$,
\begin{equation}
\mathcal{H} = \frac{G \, m^2}{8 \, \pi \, r^4},
\end{equation}
which is everywhere positive.
We are interested in the vacuum solution because it is the one that is relevant to the Bohr approach to estimating quantum mechanical energies. But it is easy to compute (especially from~\refeq{Einsimple}) the solution in cases other than vacuum -- for example, if we had a sphere of radius $R$ with constant density $\rho_0$ inside it, then the interior solution to~\refeq{EG} is just:
\begin{equation}
\Phi_i(r) = \of{\frac{A \, \sinh\of{r/r_0}}{r}}^2 \, \, \, \, \, \, \, \, \, \, r_0 \equiv \frac{c}{\sqrt{2 \, \pi \, G \, \rho_0}},
\end{equation}
where $A$ is a constant that we would use to match up to an exterior solution (at $r > R$, for example) and we have chosen the solution that is finite at the origin.
\section{Scaling of the Ground State}
For the SN system, with constants $G$, $\hbar$, and $m$, there is only one way to make an energy
\begin{equation}
E \sim \frac{G^2}{\hbar^2} \, m^5,
\end{equation}
so we expect, up to constants out front, that the energy spectrum of~\refeq{NSet} scales like $m^5$ (as indeed it does~\cite{Harrison, Bernstein}).
With the introduction of $c$ appearing in~\refeq{EG}, we can form the Planck mass: $M_p = \sqrt{\frac{\hbar \, c}{G}}$, and this means that any power of $m$ could appear in the energy spectrum of the self-coupled system (where we expect $E \sim \frac{G^2}{\hbar^2} \, m^q \, M_p^s$ with $q + s = 5$, but otherwise unconstrained). We'd like a way to estimate relevant combinations of $M_p$ and $m$ that might appear in our new spectrum. To that end, we will use the Bohr model (originally for hydrogen, of course, but applied in this gravitational setting) with potential given by~\refeq{ppSC} -- the idea is that if the ground state is localized close to the origin, then far away, the potential associated with that ground state should go roughly like~\refeq{ppSC}, and so the spectrum of the spherically symmetric vacuum solution could provide some relevant approximate information.
To start, we'll apply the Bohr method to $\Phi(r) = -\frac{G \, m}{r} + c^2$, just the Newtonian point particle potential (with an offset at spatial infinity so as to match~\refeq{ppSC}). According to the rules of old quantization (for circular orbits, updated to elliptical orbits in the Wilson-Sommerfeld formulation), we start with
the total energy for a particle moving in a circle of radius $r$ (so that $v = \sqrt{G \, m/r}$):
\begin{equation}
E= m\, c^2 -\frac{1}{2} \, \frac{G \, m^2}{r}.
\end{equation}
Next, we assume angular momentum is quantized: $L = m \, v \, r = n \, \hbar$ for integer $n$, giving us a value for the radius: $r = \frac{n^2\, \hbar^2}{G \, m^3}$ -- then using this radius in $E$ we get a discrete set of energies:
\begin{equation}
E_n = m \, c^2 - \frac{1}{2} \, \frac{G^2 \, m^5}{n^2\, \hbar^2}.
\end{equation}
The ground state corresponds to $n=1$:
\begin{equation}\label{BFitSN}
E_1 = m \, c^2 -\frac{1}{2} \, \frac{G^2}{\hbar^2} \, m^5 = m\, c^2 -\frac{1}{2} \, \frac{c^2}{M_p^4} m^5,
\end{equation}
which is what we expect, namely a linear (in $m$) offset (associated with the shift at spatial infinity) and $m^5$ scaling.
Performing the same procedure for the ``point potential" in~\refeq{ppSC} gives
\begin{equation}\label{BohrSC}
E_n = \frac{2 \, m \, c^2 \, M_p^4 \, n^2}{m^4 + 2 \, M_p^4 \, n^2},
\end{equation}
with ground state energy:
\begin{equation}\label{BFitSCSC}
E_1 = \frac{2 \, m \, c^2 \, M_p^4}{m^4 + 2 \, M_p^4},
\end{equation}
where again, we only care about the mass scaling here -- our estimate cannot predict constant offsets and/or overall constants out front. We'll come back to those later on.
Note that the expression~\refeq{BFitSCSC} reduces to~\refeq{BFitSN} in the $m \ll M_P$ limit, as it should:
\begin{equation}
\frac{2 \, m \, c^2 \, M_p^4}{m^4 + 2 \, M_p^4} = m\, c^2 -\frac{1}{2} \, \frac{c^2}{M_p^4} m^5 + O\of{\of{\frac{m}{M_p}}^8} \, m \, c^2
\end{equation}
for $m$ small.
\section{Numerical Approach}
We'll start by specializing to spherically symmetric solutions, then we'll render the equations of interest dimensionless, and put them in time-independent form to define the eigenvalue problem of interest. From there, we'll introduce the iterative finite difference approach that can be applied to either SN or self-coupled gravity to find the ground state energies.
\subsection{Dimensionless Form}
Since we are interested in the ground state energies, we will focus on spherically symmetric solutions to both SN and the modified system. For SN, we have
\begin{equation}
\begin{aligned}
-\frac{\hbar^2}{2 \, m} \, \frac{1}{r} \, \left(r\, \Psi(r,t) \right)'' + m \, \Phi(r,t) \, \Psi(r,t) &= i \, \hbar \, \frac{\partial \Psi(r,t)}{\partial t} \\
\frac{1}{r} \, \left( r \, \Phi(r,t) \right)'' &= 4 \, \pi \, G \, m \, \Psi(r,t)^* \, \Psi(r,t),
\end{aligned}
\end{equation}
and we can simplify further by setting $P(r,t) \equiv r \, \Psi(r,t)$, then the above becomes
\begin{equation}\label{SNII}
\begin{aligned}
-\frac{\hbar^2}{2 \, m} \, P(r,t)'' + m \, \Phi(r,t) \, P(r,t) &= i \, \hbar \, \frac{\partial P(r,t)}{\partial t} \\
\left( r \, \Phi(r,t) \right)'' &= \frac{4 \, \pi \, G \, m}{r} \, P(r,t)^* \, P(r,t).
\end{aligned}
\end{equation}
The wave function is normalized to $1$, so that our $P(r,t)$ has:
\begin{equation}\label{normit}
4 \, \pi \, \int_0^\infty \| P(r,t) \|^2 \, dr = 1.
\end{equation}
The modified theory becomes, under the same assumptions and substitutions,
\begin{equation}\label{SCSCII}
\begin{aligned}
-\frac{\hbar^2}{2 \, m} \, P(r,t)'' + m \, \Phi(r,t) \, P(r,t) &= i \, \hbar \, \frac{\partial P(r,t)}{\partial t} \\
\left( r \, \Phi(r,t) \right)'' &= \frac{4 \, \pi \, G \, m}{c^2 \, r} \, P(r,t)^* \, P(r,t) \, \Phi(r,t) + \frac{r}{2 \, \Phi(r,t)} \, \left(\Phi(r,t)'\right)^2.
\end{aligned}
\end{equation}
We can render the equations dimensionless by introducing $r = r_0 \, R$, $t = t_0 \, T$ for dimensionless $R$ and $T$, and taking $\Phi = \Phi_0 \, \bar\Phi$, $P = P_0 \, \bar P$, where $\Phi_0$ is a speed$^2$ and $P_0$ has dimension $1/\sqrt{\hbox{length}}$. Finally, let $m = m_0 \, \bar m$ for Planck mass $m_0 = \sqrt{\frac{\hbar \, c}{G}}$ and set $\bar P(r,t) = e^{-i \, \bar E \, T} \, \bar P(r)$ for dimensionless energy $\bar E = E/E_0$. Then our pairs take the form of an eigenvalue problem
\begin{equation}\label{SNIII}
\begin{aligned}
-\frac{1}{\bar m} \, \frac{\partial^2 \bar P}{\partial R^2} + \bar m \, \bar \Phi \, \bar P &= \bar E \, \bar P\\
\frac{\partial^2}{\partial R^2} \left( R \, \bar \Phi \right) &= \frac{\bar m}{R} \, \bar P^* \, \bar P,
\end{aligned}
\end{equation}
and
\begin{equation}\label{SCSCIII}
\begin{aligned}
-\frac{1}{\bar m}\, \frac{\partial^2 \bar P}{\partial R^2} + \bar m\, \bar \Phi \, \bar P &= \bar E \, \bar P \\
\frac{\partial^2}{\partial R^2} \left( R \, \bar \Phi \right) &= \frac{\bar m}{R} \, \bar P^* \, \bar P \, \bar \Phi + \frac{R}{2 \, \bar\Phi} \left( \frac{\partial \bar \Phi}{\partial R}\right)^2
\end{aligned}
\end{equation}
with
\begin{equation}
r_0 = \frac{\hbar}{\sqrt{2} \, m_0 \, c} \, \, \, \, \, \, \, \, \, \, \, \, t_0 =\frac{\hbar}{m_0 \, c^2} \, \, \, \, \, \, \, \, \, \, \, \, P_0 =\frac{c}{\sqrt{4 \, \pi \, m_0 \, G}} \, \, \, \, \, \, \, \, \, \, \, \, \Phi_0 = c^2 \, \, \, \, \, \, \, \, \, \, \, \, E_0 = m_0 \, c^2.
\end{equation}
The normalization of the wave function~\refeq{normit} now reads:
\begin{equation}\label{finalnorm}
\int_0^\infty \bar P^* \, \bar P \, dR = \frac{1}{4 \, \pi \, P_0^2 \, r_0} = \frac{\sqrt{2} \, G \, m_0^2}{\hbar \, c} = \sqrt{2}.
\end{equation}
\subsection{Method}
The numerical method is the same in both cases -- we discretize in $R$ by taking $R_j \equiv j \, \Delta R$, where $j = 0$ is a boundary point (at the origin -- for $\Psi$ finite at the origin, we must have $\bar P = 0$ there), and we take $R_{J+1} = R_\infty$ to be a numerical approximation to infinity, where we again require $\bar P = 0$. Let $\bar P_j \equiv \bar P(R_j)$ (and $\bar\Phi_j \equiv \bar \Phi(R_j)$), then we can discretize Schr\"odinger's equation using finite differences:
\begin{equation}\label{Schdiff}
-\frac{1}{\bar m} \, \left[ \frac{\bar P_{j+1} - 2 \, \bar P_j + \bar P_{j-1}}{\Delta R^2} \right] + \bar m \, \bar \Phi_j \, \bar P_j = \bar E \, \bar P_j
\end{equation}
for $j = 1 \ldots J$. We can define the vector $\bar{\vecbf P} \in \mat R^J$ to have entries that are precisely the unknown $\bar P_j$ values (and similarly for the vector $\bar{\bm \Phi}$), and then~\refeq{Schdiff} can be written as a matrix eigenvalue problem in the usual way:
\begin{equation}
\mat D(\bar{\bm \Phi}) \, \bar{\vecbf P} = \bar E \, \bar{\vecbf P}
\end{equation}
with
\begin{equation}\label{DDef}
\mat D(\bar{\bm \Phi}) \dot =\left(\begin{array}{cccc} \frac{2}{\bar m\, \Delta R^2} + \bar m \, \bar \phi_1 & -\frac{1}{\bar m \, \Delta R^2} & 0 & \ldots \\
-\frac{1}{\bar m \, \Delta R^2} & \frac{2}{\bar m \, \Delta R^2} + \bar m \, \bar \phi^n_2 & -\frac{1}{\bar m\, \Delta R^2} & 0 \\
0 & \ddots & \ddots & \ddots \end{array} \right).
\end{equation}
As a matrix eigenvalue problem, it is relatively easy to construct $\mat D$ and then find the eigenvector associated with the smallest eigenvalue -- that eigenvector is an approximation to the ground state. Once we have $\bar{\vecbf P}$ in hand, we can construct the entries of $\bar{\bm \Phi}$ in either the SN or augmented case using Verlet. We can find the values for $\bar\Phi_j$ by working backwards from $j = J$ to $1$ using the recursion
\begin{equation}\label{Verlet}
\bar\Phi_{j-1} = \frac{1}{R_{j-1}} \, \left[ 2 \, R_j \, \bar \Phi_j - R_{j+1} \, \bar{\Phi}_{j+1} + \frac{\bar m}{R_j} \, \| \bar P_j \|^2 \, \Delta R^2\right],
\end{equation}
where we take $\bar\Phi_{j} = 1 - \frac{\sqrt{2} \, \bar m}{R_{j}}$ for $j = J$ and $J+1$ -- since $R_{J+1} = R_\infty$, we want the potential to approximate its value out at spatial infinity, and this is the dimensionless form of $\Phi_j = c^2 - \frac{G \, m}{r_0\, R_j}$. For the gravitational field equation in~\refeq{SCSCIII}, the analogous Verlet recursion looks like
\begin{equation}\label{VerletSC}
\bar\Phi_{j-1} = \frac{1}{R_{j-1}} \, \biggl[ 2 \, R_j \, \bar \Phi_j - R_{j+1} \, \bar{\Phi}_{j+1} + \frac{\bar m}{R_j} \, \| \bar P_j \|^2 \, \bar{\Phi}_j \, \Delta R^2 + \frac{R_j}{2 \, \bar \Phi_j} \, \left( \bar \Phi_{j+1} - \bar \Phi_j \right)^2
\biggr]
\end{equation}
with the same boundary conditions as above.
So, if we had the entires of $\bar{\vecbf P}$, we could construct the entries of $\bar{\bm \Phi}$ in either case, but in order to get $\bar{\vecbf P}$, we need $\bar{\bm \Phi}$ -- the matrix $\mat D$ in~\refeq{DDef} depends on the values of the potential $\bar{\bm \Phi}$. We can use an iterative approach (similar to the one in~\cite{Bernstein}) to get around the problem. Let $\bar P^k_j$ be the $k^{\hbox{\tiny{th}}}$ iteration for $\bar P_j$, then we can construct $\bar \Phi^k_j$ using~\refeq{Verlet}, and update by finding the smallest eigenvalue/vector of $\mat D(\Phi^k)$, so
\begin{equation}\label{smalleval}
\bar{\vecbf P}^{k+1} = \hbox{smallest eigenvector of $\mat D(\bar{\bm \Phi}^k)$ normalized so that $\sum_{j=1}^n \| \bar P^{k+1}_j\|^2 \, \Delta R = \sqrt{2}$}.
\end{equation}
From this, we can construct $\bar{\bm \Phi}^{k+1}$ and iterate until:
\begin{equation}\label{stopper}
\| \bar{\vecbf P}^{k+1} - \bar{\vecbf P}^k \| \le \epsilon,
\end{equation}
for $\epsilon$, some user-specified tolerance.
\section{Energies}
\subsection{Schr\"odinger Newton}
The ground state energies for the SN system were calculated for $\bar m = .7$ to $\bar m = 2$ in steps of $.02$. In~\reffig{fig:SNPPhi}, we show $\bar P$ and $\bar \Phi$ for the masses $\bar m =.7$ (top) and $\bar m = 2$ (bottom) -- we chose these mass limits because at our value of $R_\infty = 200$ and $\Delta R = \frac{R_\infty}{5001}$, masses less than $\bar m = .7$ begin to violate our assumption that $\bar P(\infty) = 0$, and masses larger than $\bar m = 2$ get localized to only a few grid points near the origin. We can expand the mass range by changing $R_\infty$, allowing us to probe smaller masses, and by decreasing $\Delta R$, allowing us to move up in mass.
\begin{figure}[htbp]
\centering
\includegraphics[width=4in]{PPhiSN}
\caption{The numerical solutions for $\bar P$ and $\bar\Phi$ for $\bar m = .7$ (top) and $\bar m = 2$ (bottom) for the SN system. Here, $R_\infty = 200$ with $\Delta R =200/5001$, and we use $\epsilon = 10^{-8}$ as the tolerance for the iteration (see~\refeq{stopper}).}
\label{fig:SNPPhi}
\end{figure}
The energies themselves are shown in~\reffig{fig:SNenergy}. There, the dots are the numerically-determined values -- we fit the dots to a curve of the form $A \, \bar m + B \, \bar m^5$, motivated by~\refeq{BFitSN}, using a nonlinear Levenberg-Marquardt fit of the data. The curve is:
\begin{equation}\label{BfitSNC}
C(\bar m) = 1.00123\, \bar m - 0.163181\, \bar m^5.
\end{equation}
\begin{figure}[htbp]
\centering
\includegraphics[width=4in]{EbarSN}
\caption{Ground state energy as a function of mass for the SN pair -- the points are numerically determined using the numerical method described in the previous section, and the curve is the best fit curve, $C(\bar m)$, from~\refeq{BfitSNC}.}
\label{fig:SNenergy}
\end{figure}
The best fit curve has two interesting features -- first, it correctly identifies the linear $m \, c^2$ offset from~\refeq{BFitSN}, and this provides an estimate of the error in the method -- we should have a value of $1.0$ in front of the $\bar m$ term in~\refeq{BfitSNC}, but instead we get $1.00123$, an error of $\approx .1\%$. As for the coefficient of the $\bar m^5$ term, it agrees well with the accepted value of $-.163$~\cite{ Harrison, Bernstein}.
\subsection{Self-consistent, Self-Coupled}
We use the same setup and parameters to find the ground state energies for the modified, self-sourcing gravity system. In~\reffig{fig:SCSCPPhi}, we have the plots corresponding to~\reffig{fig:SNPPhi}, showing the functions $\bar P$ and $\bar\Phi$ for the smallest and largest mass values.
\begin{figure}[htbp]
\centering
\includegraphics[width=4in]{PPhiSCSC}
\caption{The numerical solutions for $\bar P$ and $\bar\Phi$ for $\bar m = .7$ (top) and $\bar m = 2$ (bottom) for the self-coupled system. Here, $R_\infty = 200$ with $\Delta R =200/5001$, and we use $\epsilon = 10^{-8}$ as the tolerance for the iteration (see~\refeq{stopper}).}
\label{fig:SCSCPPhi}
\end{figure}
This time, we fit to the function: $A \, \frac{2 \, \bar m}{2 + B \, \bar m^4}$, guided by the form of~\refeq{BFitSCSC}, and obtain
\begin{equation}\label{BfitSCSCC}
C(\bar m) = \frac{2.0205\, \bar m}{2 + 0.45448\, \bar m^4}.
\end{equation}
The numerically-determined ground state values and best fit curve are shown in~\reffig{fig:E1SCSC}.
\begin{figure}[htbp]
\centering
\includegraphics[width=4in]{EbarSCSC}
\caption{Ground state energy as a function of mass for self-coupled gravity -- the points are numerically determined using the numerical method described in the previous section, and the curve is the best fit curve, $C(\bar m$), from~\refeq{BfitSCSCC}.}
\label{fig:E1SCSC}
\end{figure}
While the fit curve captures the basic behavior of the ground state energies in this case, it does not fit as well as in the SN case. The freedom in generating energies using $M_p$ in addition to $m$ allows for a more complicated spectrum (for SN, $m^5$ is the only possibility), and we do not expect the simple estimate from the Bohr method to work as well. Still, the structure of the spectrum is described by that estimate.
\subsection{Comparison}
We have the numerical spectrum from the self-coupled case, and can compare that with the $\bar m -.163 \, \bar m^5$ scaling from the SN system -- that is shown in~\reffig{fig:SNSCE} -- for small masses, the two ground state energies agree well, but they begin to diverge near $\bar m = 1.2$. Note that both energies are offset by the same amount (these are bound state energies, so we expect them to be negative, they have been shifted upwards by the constant factor $m \, c^2$, the value of the potential energy at infinity). The energies are off by around $8 \%$ by $\bar m = 2$.
\begin{figure}[htbp]
\centering
\includegraphics[width=4in]{EbarSCSN}
\caption{Points show the ground state energy for the self-coupled system, while the solid line is the SN curve for the ground state energy.}
\label{fig:SNSCE}
\end{figure}
While the SN energies become arbitrarily negative as $\bar m$ increases, the self-coupled spectrum looks like it asymptotically approaches zero (or $-m \, c^2$, relative to the value of the potential at spatial infinity) -- indeed, this is predicted by the Bohr estimate~\refeq{BohrSC}, where the energy scales like $1/m^3$ for $m$ large.
Another way to compare the spectrum of the self-coupled case with the SN energies is to look at the energy emitted during a transition from the first excited state to the ground state. In the SN case, the first excited state has energy~\cite{Bernstein,Harrison}: $E_2 = m \, c^2 - .0308 \, G^2 \, m^5/\hbar^2$, and so the difference between the first excited state and the ground state is:
\begin{equation}
\Delta_{\hbox{\tiny{SN}}} = .1322 \, \frac{G^2 \, m^5}{\hbar^2},
\end{equation}
or $\Delta_{\hbox{\tiny{SN}}} = .1322 \, \bar m^5$ in our dimensionless variables.
To compute the first excited state using our approach, we simply perform our iteration using the eigenvector associated with the second smallest eigenvalue (a simple modification of~\refeq{smalleval})-- the energy, as a function of $\bar m$ together with the best fit from~\refeq{BFitSCSC} with $n = 2$ (so that the fit function is $A \, \bar m/(8 + B \, \bar m^4)$) is shown in~\reffig{fig:SCSCFES}.
\begin{figure}[htbp]
\centering
\includegraphics[width=4in]{EbarSCSCFES}
\caption{The energy of the first excited state for the self-coupled case, together with its best-fit curve from the Bohr estimate.}
\label{fig:SCSCFES}
\end{figure}
Subtracting the ground state for each mass, we get the $\Delta_{\hbox{\tiny SC}}$ to compare with the SN case -- that difference is shown in~\reffig{fig:SCSCDELTA}, together with the $\Delta_{\hbox{\tiny{SN}}}$ from above.
\begin{figure}[htbp]
\centering
\includegraphics[width=4in]{RAD}
\caption{The energy difference between the first excited state and ground state for SN (solid curve) and the self-coupled case (points).}
\label{fig:SCSCDELTA}
\end{figure}
Notice there that by $\bar m = 2$, the energy difference for SN is roughly ten times that of the self-coupled case -- the frequency of emission (if radiation were emitted) for SN at this mass value would be ten times the frequency in the self-coupled case, providing a clear signature for one over the other. If~\refeq{BohrSC} continues to hold qualitatively for larger $n$, we see that the self-coupled case has a maximum energy of $m\, c^2$ (the limit as $n \rightarrow \infty$), so any transition energy is bounded in this case, while the transition energies of SN will always be of the form: $\alpha \, \bar m^5$ for some constant $\alpha$.
\subsection{Total Energy in a Boson Collapse Model}
In the original application of the SN equations, appearing in~\cite{Bon}, the total energy of $N$ bosons in their ground state was calculated -- the Schr\"odinger-Newton pair appears as
\begin{equation}
\begin{aligned}
i \, \hbar \, \frac{\partial \Psi(\vecbf r,t)}{\partial t} &= -\frac{\hbar^2}{2 \, m} \nabla^2 \Psi(\vecbf r,t) + m\, \Phi(\vecbf r, t) \, \Psi(\vecbf r,t) \\
\nabla^2 \Phi(\vecbf r, t) &= 4 \, \pi \, G \, N\, m \, \Psi^*(\vecbf r, t) \, \Psi(\vecbf r, t)
\end{aligned}
\end{equation}
where $N$ particles of mass $m$ are interacting gravitationally in the same state. This is a very different application of the SN system (as compared with single-particle collapse), and yet we can use our self-consistent scalar gravity in place of the gravitational field equation here, just as for the single particle case. We'll replace the Poisson equation with
\begin{equation}\label{Ninplace}
\nabla^2 \Phi(\vecbf r) = \frac{4 \, \pi \, G }{c^2} \, N \, m\, \Psi^*(\vecbf r, t) \, \Psi(\vecbf r, t) \, \Phi(\vecbf r) + \frac{1}{2 \, \Phi(\vecbf r)} \, \nabla \Phi(\vecbf r) \cdot \nabla \Phi(\vecbf r).
\end{equation}
as usual.
The total energy, from the SN approach, is $E_{\hbox{\tiny{tot}}} = N \, m \, c^2 - .163 \, N^3 \, G^2 \, m^5/\hbar^2$, a dimensionless energy of $\bar E_{\hbox{\tiny{tot}}} = \bar m\, N - .163 \, \bar m^5 \, N^3$ -- that is plotted in~\reffig{fig:bonEtot} (the choice of $m$ just changes the scale in $N$, so we have left an unscaled axis there). There are three distinct regions in the total energy curve -- in the first, where the derivative of the total energy is positive, adding particles adds energy. In the second, where the derivative is negative, but $\bar E_{\hbox{\tiny{tot}}}$ is still positive, adding particles decreases the total energy. Finally, the total energy becomes negative when using the SN equations.
\begin{figure}[htbp]
\centering
\includegraphics[width=4in]{BonEtot}
\caption{The total energy, from~\cite{Bon}, for SN applied to a system of $N$ self-gravitating bosons in the ground state.}
\label{fig:bonEtot}
\end{figure}
We can calculate the corresponding total energy using the self-consistent scalar gravity in the form~\refeq{Ninplace} -- taking $\bar m = .3$ and $\bar m = .5$, we find the total energy and plot them with the SN result at those masses in~\reffig{fig:BOSTOTS}.
\begin{figure}[htbp]
\centering
\includegraphics[width=4in]{Etotbos}
\caption{Total energy for $N$ bosons, each with mass $\bar m = .3$ (top plot) or $\bar m = .5$ (bottom plot). The points represent the total energy as calculated using self-consistent scalar gravity, while the solid curve is the SN prediction from~\cite{Bon}.}
\label{fig:BOSTOTS}
\end{figure}
The total energies are different between the two cases -- the peak that separates the region in which the total energy grows with additional particles from the region in which the total energy decreases with additional particles has shifted (upward in $N$ for the self-coupled case), and the self-coupled energy does not become negative. There is additional physics which must be introduced (and is considered in detail in~\cite{Bon}) as the regimes change, and our goal here is only to once again draw a distinction (in principle detectable) between using the Poisson form of Newtonian gravity and using the nonlinear, self-coupled scalar gravity.
\section{Conclusion}
We have introduced an iterative numerical method to find the ground state energies for coupled quantum mechanical/gravitational problems. We tested the method with the familiar Schr\"odinger-Newton pair of equations, and found that the method worked well, agreeing with known results for that system. In addition, we used the Bohr method to predict the form of the ground-state energy for SN as a function of mass. Then, we used the same approach for the self-consistent, self-sourced scalar gravity from~\cite{GiuliniSC}. The spectrum has a very different mass dependence there, and this is to be expected given the additional (and necessary, for a theory of gravity) nonlinearity appearing in the gravitational field equation. Once again, the Bohr method provided a qualitatively relevant estimate for the energy dependence on mass for both the ground state and first excited state. The difference in energies provides a distinct signature for the self-coupled gravity -- the transition energy from the first excited state to the ground state is much less in the self-coupled case than in SN.
It would be interesting to explore the dynamics of the new, self-coupled case, and compare with the dynamics of SN, which are known. Aside from specific computational targets, we use special relativity to motivate the self-sourcing in the gravitational field equation, but we use the non-relativistic Schr\"odinger equation to capture the quantum mechanical behavior -- we could probe the high energy solutions by using the Klein-Gordon or Dirac equations. In these cases, it would be interesting to see how the self-coupled gravitational field arises in the context of~\cite{Giulini2} in which the authors show that Schr\"odinger-Newton is a natural limit of fields coupled to gravity, and the Poisson equation for the gravitational field appears as the first term in their expansion.
\subsection*{Acknowledgements}
The authors thank David Griffiths for useful commentary and feedback.
|
\section{Introduction and Problem Formulation}
Let $\mathcal{K}\subset \mathbb{R}^n$ be a convex set, and let $F:\mathbb{R}^n\to\mathbb{R}$ be an approximately convex function over $\mathcal{K}$ in the sense that
\begin{align}
\label{eq:linfty_approx}
\sup_{x\in\mathcal{K}}|F(x)-f(x)|\leq \epsilon/n
\end{align}
for some convex function $f:\mathbb{R}^n\to\mathbb{R}$ and $\epsilon>0$. In particular, $F$ may be discontinuous. We seek to find $x\in\mathcal{K}$ such that
\begin{align}
\label{eq:opt_objective}
F(x) - \min_{y \in \mathcal{K}} F(y) \leq \epsilon
\end{align}
using only function evaluations of $F$. This paper presents a randomized method based on simulated annealing that satisfies \eqref{eq:opt_objective} in expectation (or with high probability). Moreover, the number of required function evaluations of $F$ is at most $\mathcal{O}^*(n^{4.5})$ (see Corollary~\ref{cor:4.5}), where $\mathcal{O}^*$ hides polylogarithmic factors in $n$ and $\epsilon^{-1}$. Our method requires only a membership oracle for the set $\mathcal{K}$. In Section \ref{sec:de-fluc}, we consider the case when the amount of non-convexity in \eqref{eq:linfty_approx} can be much larger than $\epsilon/n$ for points away from the optimum.
In the oracle model of computation, access to function values at queried points is referred to as the zeroth-order information.
Exact function evaluation of $F$ may be equivalently viewed as approximate function evaluation of the convex function $f$, with the oracle returning a value
\begin{align}
\label{eq:approx_oracle}
F(x) \in [f(x)-\epsilon/n, f(x)+\epsilon/n].
\end{align}
A closely related problem is that of convex optimization with a \emph{stochastic} zeroth order oracle. Here, the oracle returns a noisy function value $f(x)+\eta$. If $\eta$ is zero-mean and subgaussian, the function values can be averaged to emulate, with high probability, the approximate oracle \eqref{eq:approx_oracle}. The randomized method we propose has an $\mathcal{O}^*(n^{7.5}\epsilon^{-2})$ oracle complexity for convex stochastic zeroth order optimization, which, to the best of our knowledge, is the best that is known for this problem. We refer to Section~\ref{sec:stoch_zeroth} for more details.
The motivation for studying zeroth-order optimization is plentiful, and we refer the reader to \cite{conn2009introduction} for a discussion of problems where derivative-free methods are essential. In Section~\ref{sec:apps} we sketch three areas where the algorithm of this paper can be readily applied: private computation with distributed data, two-stage stochastic programming, and online learning algorithms.
\section{Prior Work}
The present paper rests firmly on the long string of work by Kannan, Lov\'asz, Vempala, and others \citep{lovasz1993random,kannan1997random,kalai2006simulated,lovasz2006fast,lovasz2006hit,lovasz2007geometry}. In particular, we invoke the key lower bound on conductance of Hit-and-Run from \cite{lovasz2006fast} and use the simulated annealing technique of \cite{kalai2006simulated}. Our analysis extends Hit-and-Run to approximately log-concave distributions which required new theoretical results and implementation adjustments. In particular, we propose a unidimensional sampling scheme that mixes fast to a truncated approximately log-concave distribution on the line.
Sampling from $\beta$-log-concave distributions was already studied in the early work of \cite{applegate1991sampling} with a discrete random walk based on a discretization of the space. In the case of non-smooth densities and unrestricted support, sampling from approximate log-concave distributions has also been studied in \cite{belloni2009computational} where the hidden convex function $f$ is quadratic. This additional structure was motivated by the central limit theorem in statistical applications and leads to faster mixing rates. Both works used ball walk-like strategies. Neither work considered random walks that allow for long steps like Hit-and-Run.
The present work was motivated by the question of information-based complexity of zeroth-order stochastic optimization. The paper of \cite{AgaFosHsuKakRak13siam} studies a somewhat harder problem of regret minimization with zeroth-order feedback. Their method is based on the pyramid construction of \cite{NemYud83} and requires $\mathcal{O}(n^{33}\epsilon^{-2})$ noisy function evaluations to achieve a regret (and, hence, an optimization guarantee) of $\epsilon$. The method of \cite{liang2014zeroth} improved the dependence on the dimension to $\mathcal{O}^*(n^{14})$ using a Ball Walk on the epigraph of the function in the spirit of \cite{bertsimas2004solving}. The present paper further reduces this dependence to $\mathcal{O}^*(n^{7.5})$ and still achieves the optimal $\epsilon^{-2}$ dependence on the accuracy. The best known lower bound for the problem is $\Omega(n^2\epsilon^{-2})$ (see \cite{Shamir12}).
Other relevant work includes the recent paper of \cite{dyer2013simple} where the authors proposed a simple random walk method that requires only approximate function evaluations. As the authors mention, their algorithm only works for smooth functions and sets $\mathcal{K}$ with smooth boundaries --- assumptions that we would like to avoid. Furthermore, the effective dependence of \cite{dyer2013simple} on accuracy is worse than $\epsilon^{-2}$.
\section{Preliminaries}
Throughout the paper, the functions $F$ and $f$ satisfy \eqref{eq:linfty_approx} and $f$ is convex. The Lipschitz constant of $f$ with respect to $\ell_\infty$ norm will be denoted by $L$, defined as the smallest number such that $|f(x) - f(y)| \leq L \|x - y \|_{\infty}$ for $x, y \in \mathcal{K}$. Assume the convex body $\mathcal{K}\subseteq \mathbb{R}^n$ to be well-rounded in the sense that there exist $r,R>0$ such that $\mathcal{B}_2^n(r) \subseteq \mathcal{K} \subseteq \mathcal{B}_2^n(R)$ and $R/r \leq \mathcal{O}(\sqrt{n})$.\footnote{This condition can be relaxed by applying a pencil construction as in \cite{lovasz2007geometry}.} For a non-negative function $g$, denote by $\pi_g$ the normalized probability measure induced by $g$ and supported on $\mathcal{K}$.
\begin{definition}
A function $h:\mathcal{K}\to\mathbb{R}_+$ is log-concave if
$$h(\alpha x + (1-\alpha)y) \geq h(x)^\alpha h(y)^{1-\alpha}$$
for all $x,y\in \mathcal{K}$ and $\alpha\in [0,1]$. A function is called $\beta$-log-concave for some $\beta\geq 0$ if
$$h(\alpha x + (1-\alpha)y) \geq e^{-\beta} h(x)^\alpha h(y)^{1-\alpha}$$
for all $x,y\in \mathcal{K}$ and $\alpha\in [0,1]$.
\end{definition}
\begin{definition}
A function $g:\mathcal{K}\to\mathbb{R}_+$ is $\xi$-approximately log-concave if there is a log-concave function $h:\mathcal{K}\to\mathbb{R}_+$ such that
$$\sup_{x\in \mathcal{K}} |\log h(x) - \log g(x)| \leq \xi.$$
\end{definition}
\begin{lemma}
\label{lem:linfty_beta_concave}
If the function $g$ is $\beta/2$-approximately log-concave, then $g$ is $\beta$-log-concave.
\end{lemma}
For one-dimensional functions, the above lemma can be reversed:
\begin{lemma}[\cite{belloni2009computational}, Lemma 9]
\label{lem:1dsandwich}
If $g$ is a unidimensional $\beta$-log-concave function, then there exists a log-concave function $h$ such that
\begin{align*}
e^{-\beta} h(x) \leq g(x) \leq h(x) \ \ \mbox{for all} \ x \in \mathbb{R}.
\end{align*}
\end{lemma}
\begin{remark}[Gap Between $\beta$-Log-Concave Functions and $\xi$-Approximate Log-Concave Functions]
A consequence of Lemma \ref{lem:1dsandwich} is that $\beta$-log-concave functions are equivalent to $\beta$-approximately log-concave functions when the domain is unidimensional. However, such equivalence no longer holds in higher dimensions.
In the case the domain is $\mathbb{R}^n$, \cite{green1952approximately,cholewa1984remarks} established that $\beta$-log-concave functions are $\frac{\beta}{2}\log_2 (2n)$-approximately log-concave. \cite{laczkovich1999local} showed that there are functions such that the factor that relates these approximations cannot be less than $\frac{1}{4}\log_2 (n/2)$. \qed
\end{remark}
We end this section with two useful lemmas that can be found in \cite{lovasz2007geometry}.
\begin{lemma}[\cite{lovasz2007geometry}, Lemma 5.19]
\label{lem:volume}
Let $h: \mathbb{R}^n\rightarrow \mathbb{R}$ be a log-concave function. Define $M_h := \max h$ and $L_h(t) = \{ x\in \mathbb{R}^n: h(x)\geq t \}$. Then for $0<s<t<M_h$
\begin{align*}
\frac{{\sf vol}(L_h(s))}{{\sf vol}(L_h(t))} \leq \left( \frac{\log(M_h/s)}{\log(M_h/t)} \right)^n .
\end{align*}
\end{lemma}
\begin{lemma}[\cite{lovasz2007geometry}, Lemma 5.6(a)]
\label{lem:tails}
Let $X$ be a random point drawn from a log-concave distribution $h: \mathbb{R} \rightarrow \mathbb{R}_+$ and let $M_h := \max_{x\in \mathbb{R}} h(x)$. Then for every $t \geq 0$
\begin{align*}
\mathbb{P}(h(X) \leq t) < \frac{t}{M_h}.
\end{align*}
\end{lemma}
\section{Sampling from Approximate Log-Concave Distributions via Hit-and-Run}
In this section we analyze the Hit-and-Run procedure to simulate random variables from a distribution induced by an approximate log-concave function. The Hit-and-Run algorithm is as follows.
\vspace{1cm}
\begin{algorithm}[H]
\SetAlgoLined
\label{alg:hit_run}
\KwIn{a target distribution $\pi_g$ on $\mathcal{K}$ induced by a nonnegative function $g$; $x\in{\sf dom}(g)$; linear transformation $\Sigma$; number of steps $m$}
\KwOut{a point $x' \in {\sf dom}(g)$ generated by one-step Hit-and-Run walk}
initialization: a starting point $x \in {\sf dom}(g)$ \;
\For{$i=1,\ldots,m$}{
1. Choose a random line $\ell$ that passes through $x$. The direction is uniform from the surface of ellipse given by $\Sigma$ acting on sphere \;
2. On the line $\ell$ run the unidimensional rejection sampler with $\pi_g$ restricted to the line (and supported on $\mathcal{K}$) to propose a successful next step $x'$ \;
}
\caption{Hit-and-Run}
\end{algorithm}
\vspace{0.5cm}
In order to handle approximate log-concave functions we need to address implementation issues and address the theoretical difficulties caused by deviations from log-concavity which can include discontinuities. The main implementation difference lies is the unidimensional sampler. No longer a binary search yields the maximum over the line and its end points since $\beta$-log-concave functions can be discontinuous and multimodal.
We now turn to these questions.
\subsection{Unidimensional sampling scheme}
As a building block of the randomized method for solving the optimization problem \eqref{eq:opt_objective}, we introduce a one-dimensional sampling procedure. Let $g$ be a unidimensional $\beta$-log-concave function on a bounded line segment $\ell$, and let $\pi_g$ be the induced normalized measure. The following guarantee will be proved in this section.
\begin{lemma}
\label{lem:1dsampler}
Let $g$ be a $\beta$-log-concave function and let $\ell$ be a bounded line segment $\ell$ on $\mathcal{K}$. Given a target accuracy $\tilde{\epsilon}\in(0, e^{-2\beta}/2)$, Algorithm~\ref{alg:rej_samp} produces a point $X\in \ell$ with a distribution $\hat \pi_{g,\ell}$ such that
$$d_{\sf tv}(\pi_{g,\ell},\hat\pi_{g,\ell})\leq 3e^{2\beta}\tilde{\epsilon}.$$
Moreover, the method requires $\mathcal{O}^*(1)$ evaluations of the unidimensional $\beta$-log-concave function $g$ if $\beta$ is $\mathcal{O}(1)$.
\end{lemma}
\vspace{1cm}
\begin{algorithm}[H]
\SetAlgoLined
\label{alg:rej_samp}
\KwIn{unidimensional $\beta$-log-concave function $g$ defined on a bounded segment $\ell=[\uu{x},\bb{x}]$; accuracy $\tilde{\epsilon}>0$}
\KwOut{A sample $x$ with distribution $\hat\pi_{g,\ell}$ close to $\pi_{g,\ell}$}
Initialization: \textbf{(a)} compute a point $p \in \ell$ s.t. $g(p) \geq e^{-3\beta} \max_{z \in \ell} g(z)$ \;
\textbf{(b)} given target accuracy $\tilde{\epsilon}$, find two points $e_{-1},e_1$ on two sides of $p$ s.t.
\begin{equation}\label{eq:init_ei}
\begin{array}{c}
e_{-1}=\uu{x} \ \ \mbox{if }\ \ g(\uu{x})\geq \frac{1}{2}e^{-\beta} \tilde{\epsilon} g(p), \ \ \ \frac{1}{2}e^{-\beta} \tilde{\epsilon} g(p) \leq g(e_{-1}) \leq \tilde{\epsilon} g(p) \ \ \mbox{otherwise}\\
\\
e_{1}=\bb{x} \ \ \mbox{if }\ \ g(\bb{x})\geq \frac{1}{2}e^{-\beta} \tilde{\epsilon} g(p), \ \ \ \frac{1}{2}e^{-\beta} \tilde{\epsilon} g(p) \leq g(e_1) \leq \tilde{\epsilon} g(p) \ \ \mbox{otherwise};
\end{array} \end{equation}
\While{sample rejected}{
pick $x \sim {\sf unif}([e_{-1},e_1])$ and pick $r \sim {\sf unif}([0,1])$ independently\;
\eIf{$r \leq g(x)/\{g(p)e^{3\beta}\}$}{
accept $x$ and stop\;
}{
reject $x$ \;
}
}
\caption{Unidimensional rejection sampler}
\end{algorithm}
\vspace{0.5cm}
The proposed method for sampling from the $\beta$-log-concave function $g$ is a \emph{rejection} sampler that requires two initialization steps. We first show how to implement step $\mathbf{(a)}$.
\begin{algorithm}[H]
\SetAlgoLined
\label{alg:initial}
\KwIn{unidimensional $\beta$-log-concave function $g$ defined on a bounded interval $\ell = [\uu{x},\bb{x}]$}
\KwOut{a point $p \in \ell$ s.t. $g(p) \geq e^{-3\beta} \max_{z \in \ell} g(z)$}
\While{did not stop}{
set $x_l = \frac{3}{4} \uu{x} + \frac{1}{4} \bb{x}$, $x_c = \frac{1}{2} \uu{x} + \frac{1}{2} \bb{x}$ and $x_r = \frac{1}{4} \uu{x} + \frac{3}{4} \bb{x}$ \;
\eIf{$|\log g(x_l) - \log g(x_r)|>\beta $}{
set $[\uu{x},\bb{x}]$ as either $[x_l, \bb{x}]$ or $[\uu{x}, x_r]$ accordingly \;
}{
\uIf{$|\log g(x_l) - \log g(x_c)|>\beta$}{
set $[\uu{x},\bb{x}]$ as either $[x_l, \bb{x}]$ or $[\uu{x}, x_c]$ accordingly \;
}
\uElseIf{$|\log g(x_r) - \log g(x_c)|>\beta$}{
set $[\uu{x},\bb{x}]$ as either $[\uu{x}, x_r]$ or $[x_c, \bb{x}]$ accordingly \;
}
\Else{ output $p = {\displaystyle \arg\max_{x\in\{x_l,x_c,x_r\}}} f(x)$ and stop \;}
}
}
\caption{Initialization Step \textbf{(a)}}
\end{algorithm}
\vspace{0.5cm}
For the $\beta$-log-concave function $g$, let $h$ be a log-concave function in Lemma~\ref{lem:1dsandwich} and let $\tilde{L}$ denote the Lipschitz constant of the convex function ~$-\log h$. In the following two results, the $\mathcal{O}^*$ notation hides a $\log(\tilde{L})$ factor.
\begin{lemma}[Initialization Step \textbf{(a)}]\label{Lemma:InitStepa}
Algorithm~\ref{alg:initial} finds a point $p \in \ell$ that satisfies $g(p) \geq e^{-3\beta} \max_{z \in \ell} g(z)$.
Moreover, this step requires $\mathcal{O}^*(1)$ function evaluations.
\end{lemma}
\begin{lemma}[Initialization Step \textbf{(b)}]\label{Lemma:InitUni}
Let $\ell=[\uu{x},\bb{x}]$ and $p\in \ell$. The binary search algorithm finds $e_{-1}\in [\uu{x},p]$ and $e_1 \in [ p, \bb{x}]$ such that \eqref{eq:init_ei} holds.
Moreover, this step requires $\mathcal{O}^*(1)$ function evaluations.
\end{lemma}
According to Lemmas~\ref{lem:1dsampler},~\ref{Lemma:InitStepa},~\ref{Lemma:InitUni}, the unidimensional sampling method produces a sample from a distribution that is close to the desired $\beta$-log-concave distribution. Furthermore, the method requires a number of queries that is logarithmic in all the parameters.
\subsection{Mixing time}
In this section, we will analyze mixing time of the Hit-and-Run algorithm with a $\beta/2$-approximate log-concave function $g$, namely
\begin{align}
\label{eq:near_log_concave}
\exists ~~\text{ log-concave }~~~ h ~~~\text{ s.t. }~~~~~ \sup_{\mathcal{K}}| \log g - \log h|\leq \beta/2.
\end{align}
In particular, this implies that $g$ is $\beta$-log-concave, according to Lemma~\ref{lem:linfty_beta_concave}. In this section, we provide the analysis of Hit-and-Run with the linear transformation $\Sigma=I$ and remark that the results extend to other linear transformations employed to round the log-concave distributions.
The mixing time of a geometric random walk can be bounded through the spectral gap of the induced Markov chain. In turn, the spectral gap relates to the so called conductance which has been a key quantity in the literature. Consider the transition probability of Hit-and-Run with a density $g$, namely
$$
P_u^{g}(A) = \frac{2}{n \pi_n} \int_A \frac{g(x) dx}{\mu_g (u,x) |x - u|^{n-1}}
$$
where
$\mu_g(u,x) = \int_{\ell(u,x) \cap \mathcal{K}} g(y) dy$. Let $\pi_g(x) = \frac{g(x)}{\int_{y \in \mathcal{K}} g(y) dy}$ be the probability measure induced by the function $g$. The conductance for a set $S\subset \mathcal{K}$ with $0<\pi_g(S)<1$ is defined as
$$
\phi^g(S) = \frac{\int_{x \in S} P_x^g(\mathcal{K} \backslash S) d \pi_g}{ \min\{ \pi_g(S), \pi_g(\mathcal{K}\backslash S) \}},
$$
and $\phi^g$ is the minimum conductance over all measurable sets. The $s$-conductance is, in turn, defined as
$$
\phi_s^g = \inf_{S\subset \mathcal{K}, s<\pi_g(S)\leq 1/2} \frac{\int_{x \in S} P_x^g(\mathcal{K} \backslash S) d \pi_g}{ \pi_g(S) -s }.
$$ By definition we have $\phi^g \leq \phi^g_s$ for all $s>0$.
The following theorem provides us an upper bound on the mixing time of the Markov chain based on conductance. Let $\sigma^{(0)}$ be the initial distribution and $\sigma^{(m)}$ the distribution of the $m$-th step of the random walk of Hit-and-Run with exact sampling from the distribution $\pi_g$ restricted to the line.
\begin{theorem}[\cite{lovasz1993random}; \cite{lovasz2007geometry}, Lemma 9.1]
\label{thm:contraction_LV}
Let $0 < s \leq 1/2$ and let $g:\mathcal{K}\to\mathbb{R}_+$ be arbitrary. Then for every $m \geq 0$,
$$
d_{\sf tv}(\pi_g,\sigma^{(m)}) \leq H_0 \left(1 - \frac{(\phi^g)^2}{2}\right)^m\ \ \ \mbox{and} \ \ \ d_{\sf tv}(\pi_g,\sigma^{(m)}) \leq H_s + \frac{H_s}{s} \left(1 - \frac{(\phi_s^g)^2}{2}\right)^m
$$
where
$$
H_0=\sup_{x\in \mathcal{K}} \pi_g(x)/\sigma^{0}(x) \ \ \ \mbox{and} \ \ \ H_s = \sup\{|\pi_g(A) - \sigma^{(0)}(A) |: \pi_g(A) \leq s \}.
$$
\end{theorem}
Building on \cite{lovasz2006fast}, we prove the following result that provides us with a lower bound on the conductance for Hit-and-Run induced by a log-concave $h$. The proof of the result below follows the proof of Theorem 3.7 in \cite{lovasz2006fast} with modifications to allow unbounded sets $\mathcal{K}$ without truncating the random walk.
\begin{theorem}[Conductance Lower Bound for Log-concave Measures with Unbounded Support]
\label{thm:cond_lower_bound}
Let $h$ be a log-concave function in $\mathbb{R}^n$ such that the level set of measure $\frac{1}{8}$ contains a ball of radius $r$. Define $R = (\mathbb{E}_h \| X - z_h \|^2 )^{1/2}$, where $z_h = \mathbb{E}_h X$ and $X$ is sampled from the log concave measure induce by $h$. Then for any subset $S$, with $\pi_h(S) = p \leq \frac{1}{2}$, the conductance of Hit-and-Run satisfies
$$
\phi^h(S) \geq \frac{1}{C_1 n \frac{R}{r} \log^2 \frac{nR}{rp}}
$$
where $C_1>0$ is a universal constants.
\end{theorem}
Although Theorem \ref{thm:cond_lower_bound} is new, very similar conductance bounds allowing for unbounded sets were establish before. Indeed in Section 3.3 of \cite{lovasz2006fast} the authors discuss the case of unbounded $\mathcal{K}$ and propose to truncate the set to its effective diameter and use the fact that this distribution would be close to the distribution of the unrestricted set. Such truncation needs to be enforces which requires to change the implementation of the algorithm and lead to another (small) layer of approximation errors. Theorem \ref{thm:cond_lower_bound} avoids this explicit truncation and truncation is done implicitly in the proof only. We note that when applying the simulated annealing technique, even if we start with a bounded set, by diminishing the temperature, we are effectively stretching the sets which would essentially require to handle unbounded sets.
We now argue that conductance of Hit-and-Run with $\beta$-approximate log-concave measures can be related to the conductance with log-concave measures.
\begin{theorem}[Conductance Lower Bound for Approximate Log-concave Measures]
\label{thm:compare_cond}
Let $g$ be a $\beta/2$-approximate log-concave measure and $h$ be any log-concave function with the property \eqref{eq:near_log_concave}. Then the conductance and $s$-conductance of the random walk induced by $g$ are lower bounded as
$$\phi^g\geq e^{-3\beta} \phi^h ~~~~~\text{and}~~~~~\phi_s^g \geq e^{-3\beta} \phi_{s/e^{\beta}}^h.
$$
\end{theorem}
We apply Theorem~\ref{thm:contraction_LV} to show contraction of $\sigma^{(m)}$ to $\pi_g$ in terms of the total variation distance.
\begin{theorem}[Mixing Time for Approximately-log-concave Measure]
\label{thm:mixing_LC}
Let $\pi_g$ is the stationary measure associated with the Hit-and-Run walk based on a $\beta/2$-approximately log-concave function $g$, and $M=\|\sigma^{(0)}/ \pi_g\| = \int (d\sigma^{(0)}/d\pi_g) d\sigma^{(0)}$.
There is a universal constant $C<\infty$ such that for any $\gamma \in (0,1/2)$, if$$
m \geq C n^2 \frac{e^{6\beta}R^2}{r^2} \log^4 \frac{e^\beta MnR}{r \gamma^2} \log \frac{M}{\gamma}
$$
then $m$ steps of the Hit-and-Run random walk based on $g$ yield
$$d_{\sf tv}(\pi_g, \sigma^{(m)}) \leq \gamma.$$
\end{theorem}
\begin{remark}
The value $M$ in Theorem~\ref{thm:mixing_LC} bounds the impact of the initial distribution $\sigma^{(0)}$ which can be potentially far from the stationary distribution. In the Simulated Annealing application of next section, we will show in Lemma \ref{lma: warm-start} that we can ``warm start'' the chain by carefully picking an initial distribution such that $M=\mathcal{O}(1)$.
\end{remark}
Theorem \ref{thm:mixing_LC} shows $\gamma$-closeness between the distribution $\sigma^{(m)}$ and the corresponding stationary distribution. However, the stationary distribution is not exactly $g$ since the unidimensional sampling procedure described earlier truncates the distribution to improve mixing time.
The following theorem shows that these concerns are overtaken by the geometric mixing of the random walk.
Let $\hat \pi_{g,\ell}$ denote the distribution of the unidimensional sampling scheme (Algorithm~\ref{alg:rej_samp}) along the line $\ell$ and $\pi_{g,\ell}$ denote the distribution of the unidimensional sampling scheme proportional to $g$ along the line $\ell$.
\begin{theorem}\label{Thm:Allerrors}
Let $\hat{\sigma}^{(m)}$ denote the distribution of the Hit-and-Run with the unidimensional sampling scheme (Algorithm~\ref{alg:rej_samp}) after $m$ steps. For any $0<s<1/2$, the algorithm maintains that
$$d_{\sf tv}( \hat{\sigma}^{(m)}, \pi_{g}) \leq 2d_{\sf tv}( \hat{\sigma}^{(0)}, \sigma^{(0)}) + m \sup_{\ell \subset \mathcal{K}} d_{tv}(\hat \pi_{g,\ell},\pi_{g,\ell} ) + \left\{H_s + \frac{H_s}{s} \left(1 - \frac{(\phi_s^g)^2}{2}\right)^m\right\} $$
where the supremum is taken over all lines $\ell$ in $\mathcal{K}$. In particular, for a target accuracy $\gamma \in (0,1/e)$, if $d_{\sf tv}( \hat{\sigma}^{(0)}, \sigma^{(0)}) \leq \gamma /8$, $s$ such that $H_s \leq \gamma/4$, $m \geq \{2/(\phi_s^g)^2\} \log(\{H_s/s\}\{4/\gamma\})$, and the precision of the unidimensional sampling scheme to be $\tilde\epsilon= \gamma e^{-2\beta}/\{12m\}$, we have
$$d_{\sf tv}( \hat{\sigma}^{(m)}, \pi_{g}) \leq \gamma.$$
\end{theorem}
\section{Optimization via Simulated Annealing}
We now turn to the main goal of the paper: to exhibit a method that produces an $\epsilon$-minimizer of the nearly convex function $F$ in expectation. Fix the pair $f,F$ with the property \eqref{eq:linfty_approx}, and define a series of functions
$$h_{i}(x) = \exp(-f/T_i), ~~~~~ g_i(x) = \exp(-F/T_i)$$
for a chain of temperatures $\{T_i, i= 1,\ldots,K\}$ to be specified later. It is immediate that $h_i$'s are log-concave. Lemma~\ref{lem:linfty_beta_concave}, in turn, implies that $g_i$'s are $\frac{2\epsilon}{nT_i}$-log-concave.
We now introduce the simulated annealing method that proceeds in epochs and employs the Hit-and-Run procedure with the unidimensional sampler introduced in the previous section. The overall simulated annealing procedure is identical to the algorithm of \cite{kalai2006simulated}, with differences in the analysis arising from $F$ being only approximately convex.
\vspace{1cm}
\begin{algorithm}[H]
\SetAlgoLined
\label{alg:annealing}
\KwIn{A series of temperatures $\{T_i, 1\leq i \leq K\}$, $K$=number of epochs, $x \in {\rm int}\mathcal{K}$}
\KwOut{a candidate point $x$ for which $F(x) \leq \min_{y \in \mathcal{K}}F(y) + \epsilon$ holds}
initialization: well-rounded convex body $\mathcal{K}$ and $\{X_0^j, 1\leq j\leq N \}$ i.i.d. samples from uniform measure on $\mathcal{K}$, $N$-number of strands, set $\mathcal{K}_0 = \mathcal{K}$, and $\Sigma_0=I$ \;
\While{$i$-th epoch, $1 \leq i \leq K$}{
1. calculate the $i$-th rounding linear transformation $\mathcal{T}_i$ based on $\{X_{i-1}^j, 1\leq j\leq N \}$ and let $\Sigma_i = \mathcal{T}_i \circ \Sigma_{i-1}$
\;
2. draw $N$ i.i.d. samples $\{X_i^j, 1\leq j\leq N \}$ from measure $\pi_{g_i}$ using Hit-and-Run algorithm with linear transformation $\Sigma_i$ and with $N$ warm-starting points $\{X_{i-1}^j, 1\leq j\leq N \}$ \;
}
output ~~$x=\argmin{1\leq j\leq N, 1\leq i\leq K} F(X_i^j) $.
\vspace{0.3cm}
\caption{Simulated annealing}
\end{algorithm}
\vspace{0.5cm}
Before stating the optimization guarantee of the above simulated annealing procedure, we prove that the warm-start property of the distributions between successive epochs and the rounding guarantee given by $N$ samples.
\subsection{Warm start and mixing}
We need to prove that the measures between successive temperatures are not too far away in the $\ell_2$ sense, so that the samples from the previous epoch can be treated as a warm start for the next epoch. The following result is an extension of Lemma 6.1 in \citep{kalai2006simulated} to $\beta$-log-concave functions.
\begin{lemma}
\label{lma: warm-start}
Let $g(x) = \exp(-F(x))$ be a $\beta$-log-concave function. Let $\mu_i$ be a distribution with density proportional to $\exp\{-F(x)/T_i\}$, supported on $\mathcal{K}$. Let $T_i = T_{i-1}\left(1-\frac{1}{\sqrt{n}}\right)$. Then
$$\|\mu_{i}/\mu_{i+1}\|\leq C_\gamma = 5 \exp(2\beta/T_i)$$
\end{lemma}
Next we account for the impact of using the final distribution from the previous epoch $\sigma^{(0)}$ as a ``warm-start."
\begin{theorem}\label{Thm:Warm}
Fix a target accuracy $\gamma \in (0,1/e)$ and let $g$ be an $\beta/2$-approximately log-concave function in $\mathbb{R}^n$. Suppose the simulated annealing algorithm (Algorithm~\ref{alg:annealing}) is run for $K=\sqrt{n} \log (1/\rho)$ epochs with temperature parameters $T_i=(1-1/\sqrt{n})^i, 0\leq i\leq K$. If the Hit-and-Run with the unidimensional sampling scheme (Algorithm~\ref{alg:rej_samp}) is run for $m=\mathcal{O}^*(n^3)$ number of steps prescribed in Theorem~\ref{thm:mixing_LC}, the algorithm maintains that
\begin{align}
\label{eq:keep_it_close2}
d_{\sf tv}( \hat{\sigma}^{(m)}_{i} , \pi_{g_{i}}) \leq e\gamma
\end{align}
at every epoch $i$, where $\hat{\sigma}^{(m)}_{i}$ is the distribution of the $m$-th step of Hit-and-Run. Here, $m$ depends polylogarithmically on $\rho^{-1}$.
\end{theorem}
\subsection{Rounding for $\beta$-log-concave functions}
\label{sec:rounding}
The simulated annealing procedure runs $N=\mathcal{O}^*(n)$ strands of random walk to round the log-concave distribution into near-isotropic position (say $1/2$-near-isotropic) at each temperature. The $N$ strands do not interact and thus the computation within each epoch can be parallelized, further reducing the time complexity of the algorithm. For $N$ i.i.d isotropic random vectors $X_i \in \mathbb{R}^n, 1\leq i\leq N$ sampled from a log-concave measure, the following concentration holds when $N$ is large enough:
\begin{align*}
\sup_{\| v \|_{\ell_2}=1 } \left| v^T \left( \frac{1}{N} \sum_{i=1}^N X_i X_i^T \right) v - 1 \right| \leq \frac{1}{2}
\end{align*}
or, equivalently,
\begin{align}
\label{eq:spectral.concentration}
\frac{1}{2} \leq \sigma_{\min} \left(\frac{1}{N} \sum_{i=1}^N X_i X_i^T \right) \leq \sigma_{\max} \left(\frac{1}{N} \sum_{i=1}^N X_i X_i^T \right) \leq \frac{3}{2}.
\end{align}
Theorems of this type have been first achieved for uniform measures on the convex body $\mathcal{K}$ (measures with bounded Orlicz $\psi_2$ norm). \cite{bourgain1996random} proved this holds as long as $N\geq C n \log^3 n$. \cite{rudelson1999random} improved this bound to $N\geq C n \log^2 n$. For log-concave measures (with bounded Orlicz $\psi_1$ norm through Borell's lemma), \cite{guedon2007lp} proved a stronger version where $N \geq C n \log n$. See also \citep{adamczak2010quantitative} for further improvements. For bounded (almost surely) vectors, we can instead appeal to the the following literature. Theorem 5.41 in \citep{vershynin2010introduction} yields a spectral concentration bound for heavy tail random matricies with isotropic independent rows. (See also \cite{tropp2012user} Theorem 4.1 for matrix Bernstein's type inequalities.)
For our problem, we need to prove \eqref{eq:spectral.concentration} for independent near isotropic rows with $\beta$-log-concave measures. There are two ways to achieve this goal. The first is to invoke the \cite{guedon2007lp}'s result. Random vectors sampled from $\beta$-log-concave still belong to the Orlicz $\psi_1$ family, thus $N \geq \mathcal{O}(n \log n)$ is enough to achieve the goal with high probability. The second way is through the following lemma:
\begin{lemma}[\cite{vershynin2010introduction}, Theorem 5.41]
Let $X$ be an $N \times n$ matrix whose rows $X_i$ are
independent isotropic random vectors in $\mathbb{R}^n$. Let R be a number such that $\| X_i \|_{\ell_2} \leq R$ almost surely for all $i$. Then for every $t \geq 0$, one has
$$\sqrt{N} - tR \leq \sigma_{\min} (X) \leq \sigma_{\max} (X) \leq \sqrt{N} + tR $$
with probability at least $1 - 2n \exp(-c t^2)$, where $c > 0$ is an absolute constant.
\end{lemma}
Clearly if we take $N \geq C R^2 \log n$, we have
$$
\frac{1}{2} \leq \sigma_{\min} \left(\frac{1}{N} \sum_{i=1}^N X_i X_i^T \right) \leq \sigma_{\max} \left(\frac{1}{N} \sum_{i=1}^N X_i X_i^T \right) \leq \frac{3}{2}
$$
with probability at least $1 - n^{-C}$, since $\mathcal{K}$ is uniformly bounded within $R = \mathcal{O}(\sqrt{n}) r$ and isotropic condition implies $r = \mathcal{O}(1)$ (which translates into $\| X_i \|_{\ell_2} \leq \mathcal{O}(\sqrt{n})$). Thus we conclude that $N = \mathcal{\Theta}(n\log n)$ is enough for bringing a $\beta$-log concave measure into isotropic position.
\subsection{Optimization guarantee}
We prove an extension of Lemma 4.1 in \citep{kalai2006simulated}:
\begin{theorem}
\label{eq:opt_guarantee}
Let $f$ be a convex function. Let $X$ be chosen according to a distribution with density proportional to $\exp\{-f(x)/T\}$. Then
$$\mathbb{E}_f f(X) - \min_{x\in\mathcal{K}} f(x) \leq (n+1)T$$
Furthermore, if $F$ is such that $|F-f|_\infty\leq \rho$, for $X$ chosen from a distribution with density proportional to $\exp\{-F(x)/T\}$, we have
$$
\mathbb{E}_F f(X) - \min_{x \in \mathcal{K}} f(x) \leq (n+1)T \cdot \exp(2\rho/T)
$$
\end{theorem}
The above Theorem implies that the final temperature $T_K$ in the simulated annealing procedure needs to be set as $T_K = \epsilon/n$. This, in turn, leads to $K=\sqrt{n}\log(n/\epsilon)$ epochs. The oracle complexity of optimizing $F$ is then, informally,
\begin{center}
$\mathcal{O}^*(n^3)$ queries per sample ~~$\red{\times}$~~ $\mathcal{O}^*(n)$ parallel strands ~~$\red{\times}$~~ $\mathcal{O}^*(\sqrt{n})$ epochs ~~$ \red{=} ~~~~\mathcal{O}^*(n^{4.5})$
\end{center}
The following corollary summarizes the computational complexity result:
\begin{corollary}
\label{cor:4.5}
Suppose $F$ is approximately convex and $|F-f|\leq \epsilon/n$ as in \eqref{eq:linfty_approx}. The simulated annealing method with $K=\sqrt{n}\log(n/\epsilon)$ epochs produces a random point such that
$$\En f(X) - \min_{x\in\mathcal{K}} f(x) \leq \epsilon,$$
and thus
$$\En F(X) - \min_{x\in\mathcal{K}} F(x) \leq 2\epsilon.$$
Furthermore the number of oracle queries required by the method is $\mathcal{O}^*(n^{4.5})$.
\end{corollary}
\section{Stochastic Convex Zeroth Order Optimization}
\label{sec:stoch_zeroth}
Let $f:\mathcal{K}\to\mathbb{R}$ be the unknown convex $L$-Lipschitz funciton we aim to minimize. Within the model of convex optimization with stochastic zeroth-order oracle $\boldsymbol{\mathrm{O}}$, the information returned upon a query $x\in\mathcal{K}$ is $f(x) + \epsilon_x$ where $\epsilon_x$ is the zero mean noise. We shall assume that the noise is sub-Gaussian with parameter $\sigma$. That is,
$$
\mathbb{E} \exp(\lambda \epsilon_x) \leq \exp(\sigma^2 \lambda^2/2).
$$
It is easy to see from Chernoff's bound that for any $t \geq 0$
$$
\mathbb{P} (|\epsilon_x| \geq \sigma t) \leq 2\exp(-t^2/2).
$$
We can decrease the noise level by repeatedly querying at $x$. Fix $\tau>0$, to be determined later. The average $\bar{\epsilon}_x$ of $\tau$ observations is concentrated as
$$
\mathbb{P} \left( |\bar{\epsilon}_{x}| \geq \sigma t/\sqrt{\tau} \right) \leq 2\exp(-t^2/2).
$$
To use the randomized optimization method developed in this paper, we view $f(x)+\bar{\epsilon}_x$ as the value of $F(x)$ returned upon a single query at $x$. Since the randomized method does not re-visit $x$ with probability $1$, the function $F$ is ``well-defined''.
Let us make the above discussion more precise by describing three oracles. Oracle $\boldsymbol{\mathrm{O}}'$ draws noise $\epsilon_x$ for each $x\in\mathcal{K}$ prior to optimization. Upon querying $x\in\mathcal{K}$, the oracle deterministically returns $f(x)+\epsilon_x$, even if the same point is queried twice. Given that the optimization method does not query the same point (with probability one), this oracle is equivalent to an \emph{oblivious} version of oracle $\boldsymbol{\mathrm{O}}$ of the original zeroth order stochastic optimization problem.
To define $\boldsymbol{\mathrm{O}}_\alpha$, let $\boldsymbol{\mathcal{N}}_\alpha$ be an $\alpha$-net in $\ell_\infty$ which can be taken as a box grid of $\mathcal{K}$. If $\mathcal{K}\subseteq R B_\infty$, the size of the net is at most $(R/\alpha)^n$. The oracle draws $\epsilon_x$ for each element $x\in\boldsymbol{\mathcal{N}}_\epsilon$, independently. Upon a query $x'\in\mathcal{K}$, the oracle deterministically returns $f(x)+\epsilon_x$ for $x\in\boldsymbol{\mathcal{N}}_\alpha$ which is closest to $x'$. Note that $\boldsymbol{\mathrm{O}}_\alpha$ is no more powerful than $\boldsymbol{\mathrm{O}}'$, since the learner only obtains the information on the $\alpha$-net.
Oracle $\boldsymbol{\mathrm{O}}_\alpha^\tau$ is a small modification of $\boldsymbol{\mathrm{O}}_\alpha$. This modification models a repeated query at the same point, as described earlier. Parametrized by $\tau$ (the number of queries at the same point), oracle $\boldsymbol{\mathrm{O}}_\alpha^\tau$ draws random variables $\epsilon_x$ for each $x\in \boldsymbol{\mathcal{N}}_\alpha$, but sub-Gaussian parameter of $\epsilon_x$ is $\sigma/\sqrt{\tau}$. The optimization algorithm pays for $\tau$ oracle calls upon a single call to $\boldsymbol{\mathrm{O}}_\alpha^\tau$.
We argued that $\boldsymbol{\mathrm{O}}_\alpha^\tau$ is no more powerful than the original zeroth order oracle given that the algorithm does not revisit the point. In the rest of the section, we will work with $\boldsymbol{\mathrm{O}}_\alpha^\tau$ as the oracle model. For any $x$, denote the projection to the $\boldsymbol{\mathcal{N}}_\alpha$ to be $\mathcal{P}_{\boldsymbol{\mathcal{N}}_\alpha} (x)$. Define $F:\mathcal{K}\mapsto \mathbb{R}$ as
$$F(x) = f(\mathcal{P}_{\boldsymbol{\mathcal{N}}_\alpha} (x))+\epsilon_{\mathcal{P}_{\boldsymbol{\mathcal{N}}_\alpha} (x)}$$
where $\mathcal{P}_{\boldsymbol{\mathcal{N}}_\alpha} (x)$ is the closest to $x$ point of $\boldsymbol{\mathcal{N}}_\alpha$ in the $\ell_\infty$ sense. Clearly,
\begin{align}
\label{eq:linfty_bound}
|F-f|_\infty \leq \max_{x\in\boldsymbol{\mathcal{N}}_\alpha} |\epsilon_x| + \alpha L
\end{align}
where $L$ is the ($\ell_\infty$) Lipschitz constant. Since $(\epsilon_x)_{x\in\boldsymbol{\mathcal{N}}_\alpha}$ define a finite collection of sub-Gaussian random variables with sub-Gaussian parameter $\sigma$, we have that with probability at least $1-\delta$
$$\max_{x\in \boldsymbol{\mathcal{N}}_\alpha} |\epsilon_x|\leq \sigma \sqrt{\frac{2n\log(R/\alpha)+2\log(1/\delta)}{\tau}}$$
From now on, we condition on this event, which we call $\mathcal{E}$. To guarantee \eqref{eq:linfty_approx}, we set
$$\frac{\epsilon}{2n} = \sigma \sqrt{\frac{2n\log(R/\alpha)+2\log(1/\delta)}{\tau}} = \alpha L$$
where $\tau$ is the parameter from oracle $\boldsymbol{\mathrm{O}}_\alpha^{\tau}$. We use the first equality to solve for $\tau$ and the second to solve for $\alpha$:
$$\tau = \frac{\sigma^2 n^2(8n\log(R/\alpha)+8\log(1/\delta))}{\epsilon^2} = \frac{\sigma^2 n^2(8n\log(2LRn/\epsilon)+8\log(1/\delta))}{\epsilon^2} = \mathcal{O}^*(n^3/\epsilon^2)$$
and $\alpha=\epsilon/(2Ln)$.
Note here $L$ affects $\tau$ only logarithmically, and, in particular, we could have defined the Lipschitz constant with respect to $\ell_2$. We also observe that the oracle model depends on $\alpha$ and, hence, on the target accuracy $\epsilon$. However, because the dependence on $\alpha$ is only logarithmic, we can take $\alpha$ to be much smaller than $\epsilon$.
Together with the $\mathcal{O}^*(n^{4.5})$ oracle complexity proved in the previous section for optimizing $F$, the choice of $\tau=\mathcal{O}^*(n^3 \epsilon^{-2})$ evaluations per time step yields a total oracle complexity of
$$\mathcal{O}^*(n^{7.5}\epsilon^{-2})$$
for the problem of stochastic convex optimization with zeroth order information. We observe that a factor of $n^2$ in oracle complexity comes from the union bound over the exponential-sized discretization of the set. This (somewhat artificial) factor can be reduced or removed under additional assumptions on the noise, such as a draw from a Gaussian process with spatial dependence over $\mathcal{K}$. Alternatively, this $n^2$ factor could be removed completely if we could take a union bound over the polynomial number of points visited by the algorithm. Such an argument, however, appears to be tricky.
\section{Optimization of Non-convex Functions with Decreasing Fluctuations}
\label{sec:de-fluc}
Assume the non-convex function $F(x)$ has the property that the ``amount of non-convexity'' is decreasing as $x$ gets close to its global minimum $x^*$. If one has some control on the rate of this decrease, it is possible to break the optimization problem into stages, where at each stage one optimizes to the current level of non-convexity, redefines the optimization region to guarantee a smaller amount of ``non-convexity,'' and proceeds to the next stage. We are not aware of optimization methods for such a problem, and it is rather surprising that one may obtain provable guarantees through simulated annealing.
As one example, consider the problem of stochastic zeroth order optimization where the noise level decreases as one approaches the optimal point. Then, one would expect to obtain a range of oracle complexities between $\log(1/\epsilon)$ and $1/\epsilon^2$ in terms of the rate of the noise decrease.
Let us formalize the above discussion. Suppose there exists a $1$-Lipschitz $\alpha$-strongly convex function $f(x)$ with minimum achieved at $x^*\in\mathcal{K}$:
\begin{align*}
f(x) - f(x^*) \geq \langle \nabla f(x^*), x - x^* \rangle + \frac{\alpha}{2} \| x - x^*\|^2 \geq \frac{\alpha}{2} \| x - x^*\|^2.
\end{align*}
Define a measure of ``non-convexity'' of $F$ with respect to $f$ in a ball of radius $r$ around $x^*$:
\begin{align*}
\Delta(r) := \sup_{x \in \mathcal{B}_2^n(x^*,r)} |F(x) - f(x)|.
\end{align*}
We have in mind the situation where $\Delta(r)$ decreases as $r$ decreases to $0$.
At the $t$th stage of the optimization problem, suppose we start with an $\ell_2$ ball $\mathcal{B}_2^n(x_{t-1}, 2r_t)$ of radius $2r_t$ with the property
$$
\mathcal{B}_2^n(x^*, 3r_{t}) \supset \mathcal{B}_2^n(x_{t-1}, 2r_{t}) \supset \mathcal{B}_2^n(x^*, r_{t}).
$$
Next, we run the simulated annealing procedure for the approximately log-concave function defined over this ball. After $\mathcal{O}^*(n^{4.5})$ queries, we are provided with a point $x_t$ such that in expectation (or high probability)
\begin{align*}
f(x_t) - f(x^*) \leq C n \cdot \Delta(3r_t)
\end{align*}
with some universal constant $C>0$. Thanks to strong convexity,
\begin{align*}
\frac{\alpha}{2} \| x_t - x^*\|^2 \leq f(x_t) - f(x^*) \leq C n \cdot \Delta(3r_t)
\end{align*}
which suggests the recursive definition of $r_{t+1}$:
\begin{align*}
\frac{\alpha}{2Cn} r_{t+1}^2 := \frac{\alpha}{2Cn} \| x_t - x^*\|_{\ell_2}^2 \leq \Delta(3r_t).
\end{align*}
At stage $t+1$ we restrict the region to be $ \mathcal{B}_2^n(x_{t}, 2r_{t+1}) \supset \mathcal{B}_2^n(x^*, r_{t+1})$ and run the optimization algorithm again with the new parameter of approximate convexity. The recursion formula for the radius from $r_t$ to $r_{t+1}$ satisfies
$$
\frac{\alpha}{2Cn} r_{t+1}^2 \leq \Delta(3r_t).
$$
The recursion formula yields a fixed point --- a ``critical radius'' $r^*$ where no further improved can be achieved, with $\frac{\alpha}{2Cn} (r^*)^2 = \Delta(3r^*)$. Let us explore two examples:
\begin{align*}
&\text{Polynomial}:~~~~ \Delta(r) = c r^{p}, ~0 < p < 2, \\
&\text{Logarithmic}:~~~~ \Delta(r) = c \log(1+d r),
\end{align*}
where $c,d>0$ are constants. For the polynomial case, the critical radius is $r^* = \left(\frac{2 \cdot 3^pcC n}{\alpha}\right)^{\frac{1}{2-p}}$, and the required number of epochs is at most $\frac{\log \log (r_0/r^*)+ \log(1/\epsilon)}{\log (2/p)}$ if we want to get $r_t = (1+\epsilon) r$.
For the logarithmic case, the critical radius is the unique non-zero solution to
$$
\frac{2cC n}{\alpha} \log(1+3d r) = r^2.
$$
We conclude that at an $\mathcal{O}^*(1)$ multiplicative overhead on the number of oracle calls, we can optimize to any level of precision above the fixed point $r^*$ of the non-convexity decay function.
\section{Further Applications}
\label{sec:apps}
In this section, we sketch several applications of the zeroth-order optimization method we introduced. Our treatment is cursory, meant only to give a sense of the range of possible domains.
\subsection{Private computation with distributed data}
Suppose $i=1,\ldots,n$ are entities---say, hospitals---that each possess private data in the form of $m$ covariate-response pairs $\left\{(x_{i,j},y_{i,j})\right\}_{j=1}^{m}$. A natural approach to analyzing the aggregate data is to compute a minimizer $w^*$ of
\begin{align}
\label{eq:reg_erm}
f(w) = \frac{1}{mn}\sum_{i, j} \ell(x_{i,j},y_{i,j}; w) + R(w)
\end{align}
for some convex regularization function $R$ and a convex (in $w$) loss function $\ell$. For instance, $\ell(x_{i,j}, y_{i,j}; w)=(y_{i,j}-x_{i,j}\cdot w)^2$ and $R(w)=0$ would correspond to the problem of linear regression.
Given that the hospitals are not willing to release the data to a central authority that would perform the computation, how can the objective \eqref{eq:reg_erm} be minimized? We propose to use the simulated annealing method of this paper. To this end, we need to specify what happens when the value $f(w_t)$ at the current point $w_t$ is requested. Consider the following idea. The current $w_t$ is passed to a randomly chosen hospital $I_t\sim \text{unif}(1,\ldots,n)$. The hospital, in turn, privately chooses an index $J_t\sim \text{unif}(1,\ldots,m)$, computes the loss $\ell(x_{I_t,J_t}, y_{I_t,J_t}; w_t)$, adds zero-mean noise $\eta_t\sim N(0,1)$, and passes the resulting value
$$v_t = \ell(x_{I_t,J_t}, y_{I_t,J_t}; w_t) + \eta_t$$
back to the central authority. Since the computation is done privately by the hospital, the only value released to the outside world is the noisy residual. It is easy to check that $v_t$ is an unbiased estimate of $f(w_t)$:
$$\mathbb{E}[v_t] = f(w_t)$$
with respect to the random variables $(I_t,J_t)$ and $\eta_t$. Moreover, the noise level with respect to each source of randomness is of constant order. By repeatedly querying for the noisy value at $w_t$, the algorithm can reduce the noise variance, as in Section~\ref{sec:stoch_zeroth}, yet---importantly---the returned value is for a potentially different random choice of the hospital and the data point. This latter fact means that repeated querying does not allow the central authority to learn a specific data point. Interestingly, the additional layer of privacy given by the zero-mean noise $\eta_t$ presents no added difficulty to the minimization procedure, except for slightly changing a constant in the number of required queries.
\subsection{Two-stage stochastic programming}
\cite{dyer2013simple} discuss the following mathematical programming formulation:
\begin{align}
\label{eq:two_stage}
\max~~~ &px + \mathbb{E}\left[\max \left\{qy | Wy\leq Tx - \xi,~ y\in \mathbb{R}^{n_1}\right\}\right] \\
\text{subject to} ~~~&Ax\leq b, \notag
\end{align}
where $q\in\mathbb{R}^{n_1}$, $W\in\mathbb{R}^{d\times n_1}$, and $T\in\mathbb{R}^{d\times n}$. The expectation is taken over the random variable $\xi$. This problem is concave in $x$, and can be solved in two stages. If, given $x$, an approximate value for the inner expected maximum can be computed, the problem falls squarely into the setting of zeroth order optimization with approximate function evaluations. While the method of \cite{dyer2013simple} is simpler, its dependence on the target accuracy $\epsilon$ is worse. Additionally, the method of this paper can deal directly with constraint sets with non-smooth boundaries; the method can also handle more general functions in \eqref{eq:two_stage} that are not smooth.
\subsection{Online learning via approximate dynamic programming}
Online learning is a generic name for a set of problems where the forecaster makes repeated predictions (or decisions). For concreteness, suppose that on each round $t=1,\ldots,T$, the forecaster observes some side information $s_t\in S$, makes a prediction $\widehat{y}_t\in\mathcal{K}$, and observes an outcome $y_t\in\Y$. The goal of the forecaster is to ensure small regret, defined as
$$\sum_{t=1}^T \ell(\widehat{y}_t,y_t) - \inf_{f\in\F}\sum_{t=1}^T \ell(f(s_t),y_t)$$
where $\F$ is a class of strategies, mapping $S$ to $\mathcal{K}$, and $\ell:\mathcal{K}\times\Y\to\mathbb{R}$ is a cost function, which we assume to be convex in the first argument. The vast majority of online learning methods can be written as solutions to the following optimization problem (see \citealt{rakhlin2012relax}):
\begin{align*}
\widehat{y}_t = \argmin{\widehat{y}\in\mathcal{K}} \max_{y_t\in\Y} \left\{ \ell(\widehat{y},y_t) + \Phi_t(s_1,y_1,\ldots,s_t,y_t)\right\}
\end{align*}
where $\Phi_t$ is a relaxation on the minimax optimal value. One of the tightest relaxations is the so-called sequential Rademacher complexity, which itself involves an expectation over a sequence of Rademacher random variables and a supremum over the class $\F$. While the gradient of $\Phi_t$ might not be available, it is often possible to approximately evaluate this function and solve the saddle point problem approximately.
\begin{appendix}
\section{Proofs of Section 3}
\begin{proof}[Proof of Lemma \ref{lem:linfty_beta_concave}]
The proof is straightforward:
\begin{align*}
g(\alpha x + (1-\alpha)y) &\geq e^{-\beta/2} h(\alpha x + (1-\alpha)y) \geq e^{-\beta/2} h(x)^\alpha h(y)^{1-\alpha} \\
&\geq e^{-\beta/2} (e^{-\beta/2}g(x))^\alpha (e^{-\beta/2}g(y))^{1-\alpha} \geq e^{-\beta} g(x)^\alpha g(y)^{1-\alpha}.
\end{align*}
\end{proof}
\section{Proofs of Section 4.1}
\begin{proof}[Proof of Lemma \ref{Lemma:InitStepa}]
Consider a unidimensional $\beta$-log-concave function $g : \mathbb{R} \rightarrow \mathbb{R}$. In view of Lemma~\ref{lem:1dsandwich}, $g$ can be ``sandwiched'' by a log-concave function $h$ such that $e^{-\beta} h(x) \leq g(x) \leq h(x)$.
Given $\ell$, we want to find $p \in \ell$ such that $g(p) \geq e^{-3\beta} \max_{z\in \ell}g(z)$. We use the following 3-point method, inspired by \cite{AgaFosHsuKakRak13siam}, to provide such a point. Let us work with the convex function
\begin{align*}
\tilde{h} &= - \log h
\end{align*}
and a nearly-convex function $$\tilde{g} = - \log g.$$
The sandwiching guarantee can be written as
\begin{align*}
\tilde{h} (x) \leq \tilde{g}(x) \leq \tilde{h}(x) + \beta.
\end{align*}
We now claim that each iteration of the ``while'' loop of Algorithm~\ref{alg:initial} maintains the following property: either the length of the interval is reduced by at least $3/4$ while still containing the optimal point, or we have the output point $p$ that satisfies
\begin{align*}
\tilde{g}(p) &\leq \min \tilde{g} + 3 \beta
\end{align*}
In the latter case, $g(p) \geq e^{-3\beta} \max_{z \in \ell} g(z)$ as desired.
There are essentially two cases. First, if $\tilde{g}(x_l) - \tilde{g}(x_r) > \beta$ (or similarly we can argue for $|\tilde{g}(x_l) - \tilde{g}(x_c)| > \beta$ and $|\tilde{g}(x_r) - \tilde{g}(x_c)| > \beta$), we have
$$
\tilde{h} (x_l) + \beta \geq \tilde{g} (x_l) > \tilde{g}(x_r) + \beta \geq \tilde{h} (x_r) +\beta
$$
and thus $\tilde{h}(x_l) > \tilde{h}(x_r)$. Because of convexity of $\tilde{h}$ we can safely remove $[\uu{x},x_l]$ with the remaining interval still containing the point we are looking for. Second case is when
\begin{align*}
|\tilde{g}(x_l) - \tilde{g}(x_r)| \leq \beta, ~~~~|\tilde{g}(x_l) - \tilde{g}(x_c)| \leq \beta, ~~~~|\tilde{g}(x_r) - \tilde{g}(x_c)| \leq \beta
\end{align*}
Here, we can show the function $g(x)$ is flat enough for $[\uu{x}, \bb{x}]$ and thus the best of $x_l, x_c, x_r$ are good enough. It is not hard to see that
\begin{align*}
|\tilde{h}(x_l) - \tilde{h}(x_r)| \leq 2\beta, ~~~~|\tilde{h}(x_l) - \tilde{h}(x_c)| \leq 2\beta, ~~~~|\tilde{h}(x_r) - \tilde{h}(x_c)| \leq 2\beta.
\end{align*}
Consider the point $x_l$. By convexity of $\tilde{h}$, there must be a supporting line $k_l(x)$ that is below the convex function $\tilde{h}$ and such that $k_l(x_l)=\tilde(x_l)$. Thus
\begin{align*}
\min_{[\uu{x},x_{l}]} \tilde{h}(x) \geq \min_{[\uu{x},x_{l}]} k_l(x) \geq k_l(x_l) - 2\beta = \tilde{h}(x_l) - 2\beta
\end{align*}
using the fact that $|\uu{x}-x_{l}|=|x_l-x_c|$. Similarly we can prove
\begin{align*}
\min_{[x_{l},x_{r}]} \tilde{h}(x) \geq \tilde{h}(x_c) - 2\beta, ~~~~\min_{[x_{r},\bb{x}]} \tilde{h}(x) \geq \tilde{h}(x_r) - 2\beta .
\end{align*}
Thus
\begin{align*}
\min_{[\uu{x},\bb{x}]} \tilde{h}(x) \geq \min(\tilde{h}(x_l),\tilde{h}(x_c),\tilde{h}(x_r)) - 2\beta.
\end{align*}
By sandwiching
\begin{align*}
\min_{[\uu{x},\bb{x}]} \tilde{g}(x) & \geq \min(\tilde{g}(x_l),\tilde{g}(x_c),\tilde{g}(x_r)) - 3\beta
\end{align*}
and, hence,
\begin{align*}
\tilde{g}(p) &\leq \min_{[\uu{x},\bb{x}]} \tilde{g}(x) + 3\beta.
\end{align*}
It remains to show that the algorithm will terminate in an $\mathcal{O}^*(1)$ number of steps. Let $\tilde{L}$ be the Lipschitz constant of $\tilde{h}$. By the time the interval is shrunk to $|\uu{x}-\bb{x}|\leq \beta/\tilde{L}$, the algorithm must have entered the second case above and terminated.
\end{proof}
\begin{proof}[Proof of Lemma \ref{Lemma:InitUni}]
Consider a unidimensional $\beta$-log-concave function $g : \mathbb{R} \rightarrow \mathbb{R}$. In view of Lemma~\ref{lem:1dsandwich}, $g$ can be ``sandwiched'' by a log-concave function $h$ such that $e^{-\beta} h(x) \leq g(x) \leq h(x)$.
We consider the interval $[x_l,x_r] = [p,\bb{x}]$ (the other case follows similarly). If $g(\bb{x}) \geq \frac{1}{2}e^{-\beta}\tilde{\epsilon} g(p)$, set $e_1 = \bb{x}$. Otherwise we have $g(\bb{x}) < \frac{1}{2}e^{-\beta} \tilde{\epsilon} g(p)$ and we proceed. The procedure always query the midpoint $x_m$ of current interval $[x_l,x_r]$. If $g(x_m) > \tilde{\epsilon} g(p)$ set $x_l=x_m$, or if $g(x_m) < \frac{1}{2}e^{-\beta} \tilde{\epsilon} g(p)$ set $x_r = x_m$, and continue the search. Either operation halves the interval. If the midpoint $x_m$ is such that $\frac{1}{2}e^{-\beta} \tilde{\epsilon} g(p) \leq g(x_m) \leq \tilde{\epsilon} g(p)$, stop the process and return $e_1 = x_m$.
At every iteration, the interval $[x_l,x_r]$ is such that $g(x_l)>\tilde \epsilon g(p)$ and $g(x_r) < \frac{1}{2}e^{-\beta} \tilde{\epsilon} g(p)$. We now claim that the algorithm must terminate in an $\mathcal{O}^*(1)$ number of steps.
Let $\tilde{h} = -\log h$, and let $\tilde{L}$ be the Lipschitz constant of $\tilde{h}$. As soon as the length of the current interval $|x_l-x_r|<1/(2\tilde{L})$, we have $|\tilde{h}(x_l)-\tilde{h}(x_r)|<1/2$. Thus $h(x_l)/h(x_r) < e^{1/2}$ and $g(x_l)/g(x_r)< e^{1/2+\beta}$, implying that both $g(x_l)>\tilde \epsilon g(p)$ and $g(x_r) < \frac{1}{2}e^{-\beta} \tilde{\epsilon} g(p)$ cannot be true at the same time as $2e^{\beta}> e^{1/2+\beta}$. Hence, the algorithm terminates in a number of steps that is logarithmic in $\tilde{L}$.
\end{proof}
\begin{proof}[Proof of Lemma~\ref{lem:1dsampler}]
Let $h$ be the log-concave function associated with the $\beta$-log-concave function $g$ in the sense of Lemma~\ref{lem:1dsandwich}, so that $e^{-\beta}h(x)\leq g(x)\leq h(x)$ for all $x\in\ell$, and $L_f(t)$ denote the (upper) level set of a function $f$ at level $t$.
We note that since $L_g(t) \subset L_h(t)$ and $e^{-\beta}h(x)\leq g(x)\leq h(x)$, \eqref{eq:init_ei} implies that
\begin{equation}\label{AuxEq}L_{h}(e^\beta\tilde{\epsilon} g(p)) \subseteq [e_{-1},e_{1}] \subseteq L_{h}(\mbox{$\frac{1}{2}$}e^{-\beta} \tilde{\epsilon} g(p)).\end{equation}
Moreover, either $e_{-1}=\uu{x}$ or $g(e_{-1})\leq \tilde\epsilon g(p)$ which implies $h(e_{-1})\leq e^\beta\tilde\epsilon g(p)<\frac{1}{2}e^{-\beta}g(p)$ if $\tilde\epsilon<\frac{1}{2}e^{-2\beta}$.
The stationary distribution for this sampling scheme is a truncated distribution according to the $\beta$-log-concave distribution $g$ restricted to $[e_{-1},e_{1}]$. (Indeed, this correspond to the classic Accept-Reject method to simulate $g$ based on the uniform distribution with constant $M:=g(p)e^{3\beta}$, see \cite{robert2004monte} page 49.)
Therefore
$$
\begin{array}{rl}
d_{\sf tv}(\pi_{g,\ell},\hat \pi_{g,\ell}) & = d_{\sf tv}(\pi_{g,\ell}1\{\ell\setminus [e_{-1},e_1]\},\hat \pi_{g,\ell}1\{\ell\setminus[e_{-1},e_1]\}) + d_{\sf tv}(\pi_{g,\ell}1\{[e_{-1},e_1]\},\hat \pi_{g,\ell}1\{[e_{-1},e_1]\}) \\
&\leq \mathbb{P}_{z\sim g} \left(z \notin [e_{-1},e_1] \right) + \frac{\mathbb{P}_{z\sim g} \left(z \notin [e_{-1},e_1] \right)}{1-\mathbb{P}_{z\sim g} \left(z \notin [e_{-1},e_1] \right)}.
\end{array}
$$
Next we verify that the truncation error (in the total variation norm) of restricting $g$ to $[e_{-1},e_{1}]$ instead of $\ell$ is of the desired order. By Lemma~\ref{lem:tails} which quantifies the tail decay of unidimensional log-concave measures, we have
\begin{align*}
\mathbb{P}_{z\sim g} \left(z \notin [e_{-1},e_1] \right) & \leq e^\beta \cdot \mathbb{P}_{z\sim h} \left(z \notin [e_{-1},e_1]\right) \leq e^\beta \cdot \mathbb{P}_{z\sim h} \left(z \notin L_{h}(e^\beta\tilde{\epsilon} g(p)) \right) \\
& \leq e^\beta \cdot \mathbb{P}_{z\sim h} \left(z \notin L_{h}(e^\beta\tilde{\epsilon} M_h) \right) \leq e^\beta \cdot \mathbb{P}_{z\sim h}\left(h(z) \leq e^\beta \tilde{\epsilon} M_h \right) \leq e^{2\beta} \cdot \tilde{\epsilon}
\end{align*}
where we used (\ref{AuxEq}).
Thus, provided that $e^{2\beta}\tilde{\epsilon} \leq 1/2$, the total variation distance between the truncated measure $\hat\pi_{g,\ell}$ supported on $[e_{-1}, e_1]$ satisfies
$$
d_{\sf tv}(\pi_{g,\ell},\hat \pi_{g,\ell}) \leq 3e^{2\beta} \tilde{\epsilon}.
$$
In order to bound the number of evaluations we first bound the probability of the event the event $\{g(x) > \frac{1}{2}e^{-2\beta} g(p)\}$. Indeed, by (\ref{AuxEq}), if $X \sim U([e_{-1},e_{1}])$ and $Z \sim U\big(L_h(\mbox{$\frac{1}{2}$}\tilde\epsilon e^{-\beta} g(p))\big)$, we have
\begin{align*}
\mathbb{P}_{X} \left(g(X) \geq \frac{1}{2}e^{-2\beta}g(p)\right) & \geq \mathbb{P}_{X} \left(h(X) \geq \frac{1}{2}e^{-\beta} g(p) \right) \\
& \geq \mathbb{P}_{Z} \left(h(Z) \geq \frac{1}{2}e^{-\beta}h(p)\right) \\
& = \frac{{\sf vol}(L_h(\frac{1}{2}e^{-\beta} h(p)))}{{\sf vol}(L_h( \mbox{$\frac{1}{2}$}\tilde{\epsilon} e^{-\beta} h(p)))} \end{align*}
By Lemma~\ref{lem:volume} with $s = \mbox{$\frac{1}{2}$}\tilde{\epsilon} e^{-\beta} h(p)$ and $t=\mbox{$\frac{1}{2}$}e^{-\beta} h(p)$, it follows that
\begin{align*}
\frac{{\sf vol}(L_h(\mbox{$\frac{1}{2}$}e^{-\beta} h(p)))}{{\sf vol}(L_h( \mbox{$\frac{1}{2}$}\tilde{\epsilon} e^{-\beta} h(p)))} & \geq \frac{\log \frac{\max h}{\mbox{$\frac{1}{2}$}e^{-\beta}h(p)}}{\log \frac{\max h}{\mbox{$\frac{1}{2}$}\tilde{\epsilon} e^{-\beta} h(p)}} = \frac{\log \frac{\max h}{h(p)}+ \log 2 + \beta}{\log \frac{\max h}{ h(p)}+ \log 2 + \log 1/\tilde{\epsilon} + \beta} \geq \frac{\log 2 + \beta}{\log 2 + \beta + \log 1/\tilde{\epsilon} }\geq \frac{\log 2}{\log (2/\tilde{\epsilon})}.\end{align*}
Then, since $r\sim U([0,1])$ we have
\begin{align*}
& \mathbb{P}\left(r \leq \frac{e^{-3\beta} g(x)}{g(p)} \mid g(x) \geq \frac{1}{2}e^{-2\beta} g(p) \right) \geq \frac{1}{2}e^{-5\beta}
\end{align*}
and thus
\begin{align*}
\mathbb{P}\left(r \leq \frac{e^{-3\beta} g(x)}{g(p)} \right) \geq \frac{e^{-5\beta}\log 2}{2\log (2/\tilde{\epsilon})}.
\end{align*}
Since we have a lower bound on the acceptance probability on each sampling step, the number of iterations we need to sample is of the order $\frac{e^{-5\beta}\log 2}{2\log (2/\tilde{\epsilon})}$. This quantity is $\mathcal{O}(\log(1/\tilde\epsilon))$ if $\beta$ is $\mathcal{O}(1)$.
\end{proof}
\section{Proofs of Section 4.2}
\begin{proof}[Proof of Theorem \ref{thm:compare_cond}]
Define the shorthand $\rho=e^{\beta/2}$. By sandwiching,
$$
\rho^{-1} h(x) \leq g (x) \leq \rho h(x)
$$
Then
\begin{align*}
\rho^{-2} \pi_h(x) \leq \pi_g(x) \leq \rho^2 \pi_h(x) \\
\rho^{-2} P_u^h(A) \leq P_u^g(A) \leq \rho^2 P_u^h(A)
\end{align*}
for any $x, u \in \mathcal{K}$ and $A \subset \mathcal{K}$.
Thus we have
$$
\phi^{g}(S) \geq \rho^{-6} \phi^{g}(S).
$$
The $s$-conductance bound can be derived as follows.
\begin{align*}
\phi_s^g & = \inf_{A \subset \mathcal{K}, s \leq \pi_g(A) \leq 1/2} \frac{\int_{x \in A} P_x^g(K \backslash A) d \pi_g}{\pi_g(A) - s} \\
&\geq \rho^{-6} \inf_{A \subset \mathcal{K}, s \leq \pi_g(A) \leq 1/2} \frac{\int_{x \in A} P_x^h(K \backslash A) d \pi_h}{\pi_h(A) - s/\rho^2} \\
&\geq \rho^{-6} \inf_{A \subset \mathcal{K}, s/\rho^2 \leq \pi_h(A) \leq 1/2} \frac{\int_{x \in A} P_x^h(K \backslash A) d \pi_h}{\pi_h(A) - s/\rho^2} = \rho^{-6} \phi_{s/\rho^2}^h
\end{align*}
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm:mixing_LC}]
The $H_s$ defined in Theorem~\ref{thm:contraction_LV} can be upper bounded by
\begin{align*}
H_s & = \sup_{A:\pi_g(A) \leq s} \int_{A} \left| \frac{d \sigma^{(0)}}{d \pi_g } - 1 \right| d \pi_g \\
& \leq \sup_{A:\pi_g(A) \leq s} \left\{ \int_{A} \left( \frac{d \sigma^{(0)}}{d \pi_g } - 1 \right)^2 d \pi_g \cdot \int_{A} d \pi_g \right\}^{1/2} \\
& \leq s^{1/2} \sup_{A:\pi_g(A) \leq s} \left\{ \int_{A} \left( \frac{d \sigma^{(0)}}{d \pi_g } - 1 \right)^2 d \pi_g \right\}^{1/2} \leq s^{1/2} \| \sigma^{(0)}/\pi_g \|^{1/2} = s^{1/2} M^{1/2}.
\end{align*}
Let us now use upper bound of Theorem~\ref{thm:contraction_LV} with $s = \left(\frac{\gamma}{2M^{1/2}}\right)^2$ and $D = \frac{2R}{r}$, as well as Theorem~\ref{thm:compare_cond}. We obtain
\begin{align*}
d_{tv}(\pi_g,\sigma^{(m)}) & \leq \frac{\gamma}{2} + \frac{2M}{\gamma} \left(1 - \frac{(\phi_{s}^g)^2}{2} \right)^m \\
& \leq \frac{\gamma}{2} + \frac{2M}{\gamma} \left(1 - \frac{(\rho^{-6}\phi_{s/\rho^2}^h)^2}{2} \right)^m \\
& \leq \frac{\gamma}{2} + \frac{2M}{\gamma} \exp\left( - \frac{m (\rho^{-6}\phi_{s/\rho^2}^h)^2 }{2} \right)
\end{align*}
where $\rho=e^{\beta/2}$. In view of Theorem~\ref{thm:cond_lower_bound},
$$\phi_{s/\rho^2}^h \geq \frac{cr}{nR\log^2\left(\frac{\rho^2 nR}{r\left(\gamma/2M^{1/2}\right)^2}\right)}$$
we arrive at
\begin{align*}
d_{tv}(\pi_g,\sigma^{(m)}) & \leq \frac{\gamma}{2} + \frac{2M}{\gamma} \exp\left( - \frac{m}{2} \left(\frac{\rho^{-6}c r}{n R \log^2 \frac{\rho^2 n R M}{r\gamma^2}} \right)^2 \right)
\end{align*}
Hence, if
$$
m \geq C n^2 \frac{\rho^{12}R^2}{r^2} \log^4 \frac{\rho^2 M nR}{r \gamma^2} \log \frac{M}{\gamma}
$$
then
$$d_{\sf tv}(\pi_g, \sigma^{(m)}) \leq \gamma.$$
\end{proof}
\begin{proof}[Proof of Theorem \ref{Thm:Allerrors}]
Step 1. (Main Step) By the triangle inequality we have
\begin{equation}\label{TriIneq}
d_{\sf tv}(\hat \sigma^{(m)}, \pi_g) \leq d_{\sf tv}(\hat\sigma^{(m)}, \sigma^{(m)}) + d_{\sf tv}(\sigma^{(m)}, \pi_g)
\end{equation}
The last term in (\ref{TriIneq}) converges to zero at a geometric rate in $m$ by Theorem~\ref{thm:mixing_LC}. Specifically we have
$$d_{\sf tv}(\sigma^{(m)}, \pi_g) \leq H_s + \frac{H_s}{s} \left(1 - \frac{(\phi_s^g)^2}{2}\right)^m.$$
To bound $d_{\sf tv}(\hat\sigma^{(m)}, \sigma^{(m)})$, the total variation distance after $m$ steps between the two random walks from their corresponding starting distributions $\hat{\sigma}^{(0)}$ and $\sigma^{(0)}$, write for any measurable set $A$
$$
\hat{\sigma}^{(m)}(A) = \int (P_x^{\hat{g}})^{(m)}(A) \hat{\sigma}^{(0)} d x \ \ \mbox{and} \ \ \sigma^{(m)}(A) = \int (P_x^{g})^{(m)}(A) \sigma^{(0)} d x
$$
so that
\begin{align*}
d_{\sf tv}\left( \sigma^{(m)}, \hat{\sigma}^{(m)}\right) \leq \sup_{u\in \mathcal{K}} ~ d_{\sf tv}\left((P_u^{\hat{g}})^{(m)}, (P_u^{g})^{(m)} \right) + 2d_{\sf tv}\left(\hat{\sigma}^{(0)},\sigma^{(0)} \right).
\end{align*}
The result follows from Step 2 that shows $\sup_{u\in \mathcal{K}} ~ d_{\sf tv}\left((P_u^{\hat{g}})^{(m)}, (P_u^{g})^{(m)} \right) \leq m \sup_{\ell \subset \mathcal{K}} d_{tv}(\hat \pi_{g,\ell},\pi_{g,\ell} ).$
Step 2. (Error Propagation Bound in $m$ Steps) The unidimensional sampling scheme produces a sample from a truncated distribution (see Lemma~\ref{lem:1dsampler}). That is, at each step of the Hit-and-Run algorithm, we are sampling from a truncated measure according to a truncated function $\hat{g}$ along each line $\ell$ of $g$ (approximately-log-concave function in $\mathbb{R}^n$). Let us denote the transition probability kernel starting from $u$ for this truncated function to be $P_u^{\hat{g}}$
and the kernel for the original function is $P_u^{g}$. Let us bound the total variation distance between these two kernels through the spherical (elliptical) coordinate system, and with $p(\cdot)$ being the density corresponding to the measure $P(\cdot)$).
Suppose that $\sup_{\ell \subset \mathcal{K}} d_{tv}(\hat \pi_{g,\ell},\pi_{g,\ell} ) \leq \tilde\epsilon$.
Since $p_u^{\hat{g}}(\theta) = p_u^{g}(\theta)=p(\theta)$, it holds that
\begin{align*}
2 d_{\sf tv}(P_u^{\hat{g}}, P_u^{g}) & = \int |p_u^{\hat{g}}(r|\theta)p_u^{\hat{g}}(\theta) - p_u^{g}(r|\theta) p_u^{g}(\theta)| dr d\theta \\
& \leq \int \left\{ \int |p_u^{\hat{g}}(r|\theta) - p_u^{g}(r|\theta)| dr \right\} p(\theta) d\theta \\
& \leq 2\tilde{\epsilon} \int p(\theta) d\theta = 2\tilde{\epsilon}
\end{align*}
where on each line (over all $\theta$ according to the measure given by the linear transformation composed with uniform direction) the truncated distribution is an $\tilde{\epsilon}$ approximation to $\pi_g$. We now claim that the $m$-fold iterate of the Hit-and-Run kernel satisfies $d_{\sf tv}\left((P_u^{\hat{g}})^{(m)}, (P_u^{g})^{(m)}\right) \leq m \tilde{\epsilon}$. Let us prove this by induction. Suppose it holds for $m-1$ steps. Then
\begin{align*}
2 d_{\sf tv}\left((P_u^{\hat{g}})^{(m)}, (P_u^{g})^{(m)} \right) &= \int | \int (p_u^{\hat{g}})^{(m-1)}(y) p_y^{\hat{g}}(x) - (p_u^g)^{(m-1)}(y) p_y^g(x) dy | dx \\
&\leq \int | \int (p_u^{\hat{g}})^{(m-1)}(y) p_y^{\hat{g}}(x) - (p_u^g)^{(m-1)}(y) p_y^{\hat{g}}(x) dy | dx \\
& \quad + \int | \int (p_u^g)^{(m-1)}(y) p_y^{\hat{g}}(x) - (p_u^g)^{(m-1)}(y) p_y^g(x) dy | dx \\
& \leq \int |(p_u^{\hat{g}})^{(m-1)}(y) - (p_u^g)^{(m-1)}(y) | \left( \int p_y^{\hat{g}}(x) dx \right) dy \\
& \quad + \int (p_u^g)^{(m-1)}(y) \left( \int| p_y^{\hat{g}}(x) - p_y^g(x)| dx\right) dy \\
& \leq 2 d_{\sf tv}\left((P_u^{\hat{g}})^{(m-1)}, (P_u^{g})^{(m-1)} \right) + 2 \max_y ~ d_{\sf tv}(P_y^{\hat{g}}, P_y^{g}) \\
& \leq 2(m-1)\tilde{\epsilon} + 2\tilde{\epsilon} = 2m\tilde{\epsilon}
\end{align*}
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm:cond_lower_bound}]
The proof follows closely the arguments in the proof of Theorem 3.7 in \cite{lovasz2006fast} for bounded sets with modifications to avoid the truncation device discussed in Section 3.3 of \cite{lovasz2006fast}. Define the step-size $F(x)$ by $P(\|x-y\|\leq F(x))=1/8$ where $y$ is a random step from $x$. Next define $\lambda(x,t)= {\rm vol}((x+tB)\cap L(\frac{3}{4}f(x)))/{\rm vol}(tB)$ and $s(x)=\sup\{t>0:\lambda(x,t)\geq 63/64\}$. Finally, $\alpha(x) = \inf\{ t \geq 3 :P(f(y)\geq t f(x)) \leq 1/16\}$ where $y$ is a hit-and-run step from $x$.
Let $\mathcal{K} = S_1 \cup S_2$ be a partition into measurable sets, where $S_1 = S$ and $p = \pi_h(S_1) \leq \pi_h(S_2)$. For for $D = R \log (C' n^2/p )$ we will prove that
\begin{align}
\label{cond.eqtn}
\int_{S_1} P_x(S_2) dx \geq \frac{1}{C nD \log \frac{nD}{p}} \pi_h(S_1).
\end{align}
Consider the points that are deep inside these sets with respect to 1-step distribution
$$
S_1' = \{ x \in S_1 : P_x(S_2) < 1/1000 \}
\ \
\mbox{and}
\ \
S_2' = \{ x \in S_2: P_x(S_1) < 1/1000 \},
$$
and the complement $S_3' = \mathcal{K} \backslash S_1' \cup S_2'$.
Suppose $\pi_h(S_1') < \pi_h(S_1)/2$. Then
$$
\int_{S_1} P_x(S_2) dx \geq \frac{1}{1000} \pi_h(S_1 \backslash S_1') \geq \frac{1}{2000} \pi_h(S_1)
$$
which proves \eqref{cond.eqtn}. Thus we can assume
\begin{align}
\pi_h(S_1') \geq \pi_h(S_1)/2 \ \ \mbox{and} \ \
\pi_h(S_2') \geq \pi_h(S_2)/2.
\end{align}
Define the exceptional subset $W$ as set of points $u$ for which $\alpha(u)$ is very large
$$
W = W_1 \cup W_2, \ \ \mbox{where} \ W_1 := \{ u \in S:\alpha(u) \geq 2^{27} nD/p\} \ \ \mbox{and} \ \ W_2 := \{u \in \mathcal{K} : \|x-z_h\| \geq D \}.
$$
By Lemma 6.10 in \cite{lovasz2007geometry}, $\pi_h(W_1) \leq p/\{2^{23} nD\}$ and, by Lemma 5.17 in \cite{lovasz2007geometry}, $\pi_h(W_2) \leq \frac{p}{Cn^2}$. Now for any $u \in S_1' \backslash W$ and $v \in S_2' \backslash W$
$$
d_{tv}(P_u,P_v) \geq P_u(S_1) - P_v(S_1) = 1 - P_u(S_2) -P_v(S_1) > 1- \frac{1}{500}
$$
from the definition of $S_1'$ and $S_2'$. Thus by Lemma 6.8 of \cite{lovasz2006hit}, we have
$$
d_h(u,v) \geq \frac{1}{128 \log (3 + \alpha(u))} \geq \frac{1}{2^{12} \log \frac{nD}{p}} \ \ \mbox{or} \ \
|u - v| \geq \frac{1}{4\sqrt{n}} \max \{ F(u), F(v) \}.
$$
By Lemma 3.2 of \cite{lovasz2006hit}, the latter implies that
$$
|u - v| \geq \frac{1}{2^8 \sqrt{n}} \max \{ s(u), s(v) \}
$$
In either case, by Lemma 3.5 in \cite{lovasz2006fast}, for any point $x \in [u, v]$, we have
\begin{align*}
s(x) & \leq 2^{14} \log \frac{nD}{p} |u-v|\sqrt{n} \\
& \leq 2^{14} \log \frac{nD}{p} d_{\mathcal{K}\setminus W_2}(u,v) D \sqrt{n}
\end{align*}
where the second inequality follows from $u,v \in \mathcal{K}\setminus W_2$.
Recall the original partition $S_1', S_2'$ and $S_3'=\mathcal{K}\setminus \{S_1'\cup S_2'\}$ of $\mathcal{K}$. We will apply Theorem 2.1 of \cite{lovasz2006hit} with a different partition. Consider the partition of $ \mathcal{K} \setminus W_2$ defined as $\bar S_1 = S_1' \backslash W$, $\bar S_2 = S_2' \backslash W$ and $\bar S_3=\mathcal{K}\setminus \{W_2 \cup S_1'\cup S_2'\}$. These definitions imply that $\bar S_3 \subset S_3' \cup W$ so that \begin{equation}\label{S3rel}\pi_h(\bar S_3) \leq \pi_h(S_3')+\pi_h(W).\end{equation} Define for $x \in \mathcal{K} \setminus W_2$
$$
\bar h(x) = \frac{s(x)}{2^{16} D \sqrt{n} \log \frac{nD}{p}}
$$
It follows that for any $u \in S_1'\backslash W$ and $v \in S_2'\backslash W$ and $x \in [u,v]$, we have
$\bar h(x) \leq d_{\mathcal{K}\setminus W_2}(u,v)/3$, and since $\mathcal{K}\setminus W_2$ is a convex body, we have by Theorem 2.1 of \cite{lovasz2006hit} that
\begin{equation}\label{BigStep} \int_{\bar S_3} h(x) dx \geq \frac{\int_{\mathcal{K}\setminus W_2} \bar h(x) h(x) dx }{\int_{\mathcal{K}\setminus W_2} h(x) dx} \min\left\{\int_{\bar S_1} h(x) dx,\int_{\bar S_2} h(x) dx \right\} \end{equation}
Although $\mathbb{E}_h(s(x))$ is large, we need a lower bound on $\int_{\mathcal{K}\setminus W_2} s(x) h(x) dx/ \int_\mathcal{K} h(x)dx$. Since $s(x)$ can be large if $\mathcal{K}$ is unbounded we modify the standard bound next. Because the level set of measure $1/8$
contains a ball of radius $r$, we have
$$\begin{array}{rl}
\int_{\mathcal{K}\setminus W_2} s(x) h(x) dx/\int_\mathcal{K} h(x)dx &= \int_{\mathcal{K}\setminus W_2} \int_0^{s(x)} dt h(x) dx /\int_\mathcal{K} h(x)dx\\
& = \int_0^\infty \int_{\{ x \in \mathcal{K}\setminus W_2: \lambda(x,t) \geq 63/64 \}} h(x) dx dt /\int_\mathcal{K} h(x)dx\\
& \geq \frac{1}{16}\int_0^\infty \int_{\{x \in \mathcal{K}\setminus W_2: \lambda(x,t)\geq 3/4\}}h(x)dxdt/\int_\mathcal{K} h(x)dx\\
& \geq \frac{1}{16}\int_0^\infty \left(\int_{\{x \in \mathcal{K}: \lambda(x,t)\geq 3/4\}}h(x)dx - \int_{W_2}h(x)dx\right)_+dt/\int_\mathcal{K} h(x)dx \\
& \geq \frac{1}{16}\int_0^\infty \left(\frac{1}{2}-\frac{12t\sqrt{n}}{r} - \frac{p}{Cn^2}\right)_+dt \\
& \geq \frac{r}{2^{12}\sqrt{n}}
\end{array}$$
where the first and third inequality follows from page 998 in \cite{lovasz2006hit}, the second by definition of $W_2$ where we take $C$ and $n$ large enough. Therefore, dividing both sides of (\ref{BigStep}) by $\int_\mathcal{K} h(x) dx$, (\ref{S3rel}), $\pi_h(S_i') \geq \pi_h(S_i)/2$, $i=1,2$, and $p=\pi_h(S_1)\leq \pi_h(S_2)$, we have
\begin{align*}
\pi_h(S_3') + \pi_h(W) \geq \pi_h(\bar S_3) &\geq \frac{\int_{\mathcal{K}\setminus W_2} \bar h(x) h(x) dx }{\int_{\mathcal{K}\setminus W_2} h(x) dx} \min\{\pi_h(\bar S_1), \pi_h(\bar S_2)\}\\
& = \frac{\int_{\mathcal{K}\setminus W_2} s(x) h(x) dx/\int_\mathcal{K} h(x)dx}{2^{16} D \sqrt{n} \log \frac{nD}{p}} \min\{\pi_h(S_1' \backslash W), \pi_h(S_2' \backslash W)\}\\
& \geq \frac{1}{2^{28} n(D/r) \log \frac{nD}{p}}\left\{\min\{\pi_h(S_1'), \pi_h(S_2')\} - \pi_h(W) \right\}\\
& \geq \frac{1}{2^{28} n(D/r) \log \frac{nD}{p}}\left\{\frac{1}{2}\min\{\pi_h(S_1), \pi_h(S_2)\} - \pi_h(W) \right\}\\
& \geq \frac{1}{2^{28} n(D/r) \log \frac{nD}{p}}\left\{\frac{\pi_h(S_1)}{2} - p/4 \right\} \geq \frac{\pi_h(S_1)}{2^{32} n(D/r) \log \frac{nD}{p}}\\
\end{align*}
where we used that $\pi_h(W) \leq p/4$. Therefore,
$$ \int_{S_1} P_x(S_2)dx \geq \frac{\pi_h(S_3')}{2000}\geq \frac{\pi_h(S_1)}{Cn(D/r)\log(nD/p)} $$
\end{proof}
\section{Proofs of Section 5}
\begin{proof}[Proof of Lemma \ref{lma: warm-start}] Define
$$
Y(a) = \int_{\mathcal{K}} e^{-F(x)a} dx
$$
With this notation, we have that
$$
\left\| \frac{\mu_i}{\mu_{i+1}} \right\| = \frac{Y(2/T_i - 1/T_{i+1}) Y(1/T_{i+1})}{Y(1/T_i)^2}
$$
Define
$$
G(x,t) = g\left(\frac{x}{t} \right)^t.
$$
Then it holds that
\begin{align*}
G(\lambda(x,t)+(1-\lambda)(x',t')) & = g\left(\frac{\lambda x+ (1-\lambda) x'}{\lambda t + (1 - \lambda)t'} \right)^{\lambda t + (1 - \lambda)t'} \\
& = g\left(\frac{\lambda t}{\lambda t + (1 - \lambda)t'} \frac{x}{t} + \frac{(1-\lambda) t'}{\lambda t + (1 - \lambda)t'} \frac{x'}{t'} \right)^{\lambda t + (1 - \lambda)t'} \\
& \geq \exp(-\beta(\lambda t + (1-\lambda) t')) g\left( \frac{x}{t} \right)^{\lambda t} g\left( \frac{x'}{t'} \right)^{(1-\lambda) t'} \\
& = \exp(-\beta(\lambda t + (1-\lambda) t')) G(x,t)^\lambda G(x',t')^{(1-\lambda)}
\end{align*}
Because
$$
\int_{\mathcal{K}} G(x,t) dx = \int_{\mathcal{K}} g\left( \frac{x}{t} \right)^t dx = t^n \int_{\mathcal{K}} g(x)^t dx = t^n Y(t)
$$
through Pr\'{e}kopa-Leindler inequality, we have
$$
\left(\frac{a+b}{2}\right)^{2n} Y\left( \frac{a+b}{2} \right)^2 \geq \exp(-\beta(a+b))a^n Y(a) b^n Y(b)
$$
Take $a = 2/T_i - 1/T_{i+1}$ and $b = 1/T_{i+1}$. Then we have
\begin{align*}
\left\| \frac{\mu_i}{\mu_{i+1}} \right\| &= \frac{Y(2/T_i - 1/T_{i+1}) Y(1/T_{i+1})}{Y(1/T_i)^2} \\
& \leq \left( \frac{1/T_i^2}{(2/T_i - 1/T_{i+1})(1/T_{i+1})} \right)^n \exp(2\beta/T_i) \\
& = \left( 1+\frac{1}{n - 2\sqrt{n}} \right)^n \exp(2\beta/T_i) \leq e^{n/(n-2\sqrt{n})} \exp(2\beta/T_i) \leq 5 \exp(2\beta/T_i)
\end{align*}
\end{proof}
\begin{proof}[Proof of Theorem \ref{Thm:Warm}]
Let us prove \eqref{eq:keep_it_close2} by induction. Suppose at the end of epoch $i$, we have
$$
d_{\sf tv}( \hat{\sigma}^{(m)}_{i} , \pi_{g_i}) \leq \left(1+\frac{1}{n\log 1/\rho}\right)^i \gamma
$$
We identify $\hat{\sigma}^{(m)}_{i} = \hat{\sigma}^{(0)}_{i+1}$, $\sigma^{(0)}_{i+1} = \pi_{g_{i}}$. Hence, by Step 2 in the proof of Theorem \ref{Thm:Allerrors}, we have
\begin{align*}
d_{\sf tv}( \hat{\sigma}^{(m)}_{i+1} , \pi_{g_{i+1}})
&\leq d_{\sf tv}( \sigma^{(m)}_{i+1} , \pi_{g_{i+1}}) + d_{\sf tv}( \hat{\sigma}^{(m)}_{i+1} ,\sigma^{(m)}_{i+1}) \\
&\leq d_{\sf tv}( \sigma^{(m)}_{i+1} , \pi_{g_{i+1}}) + m\tilde{\epsilon} + d_{\sf tv}(\hat{\sigma}^{(0)}_{i+1},\sigma^{(0)}_{i+1} ) \\
& =d_{\sf tv}( \sigma^{(m)}_{i+1} , \pi_{g_{i+1}}) + m\tilde{\epsilon} + 2d_{\sf tv}( \hat{\sigma}^{(m)}_{i} , \pi_{g_{i}}) \\
&\leq \left(1+\frac{1}{n\log 1/\rho}\right)^i \gamma \cdot \frac{1}{4n \log 1/\rho} +\left(1+\frac{1}{n\log 1/\rho}\right)^i \gamma \cdot \frac{1}{4n \log 1/\rho} + \left(1+\frac{1}{n\log 1/\rho}\right)^i \gamma \\
&\leq \left(1+ \frac{1}{n\log 1/\rho}\right)^{i+1} \gamma
\end{align*}
by choosing $m = \mathcal{O}^*(n^3)$ and $$\tilde{\epsilon} =\frac{1}{m} \cdot \left(1+\frac{1}{n\log 1/\rho}\right)^i \gamma \cdot \frac{1}{4n\log 1/\rho}.$$
Thus the final epoch, $i=\sqrt{n}\log 1/\rho$, the error is at most $\left(1+\frac{1}{n\log 1/\rho}\right)^{\sqrt{n}\log 1/\rho} \gamma \leq e \gamma$.
\end{proof}
\begin{proof}[Proof of Theorem \ref{eq:opt_guarantee}]
In \cite{kalai2006simulated}, the authors proved the theorem for the case $f(x) = c\cdot x$ and claimed it can be extended to arbitrary convex functions. Here we give a proof for the sake of completeness. \cite{kalai2006simulated} proved above inequality with arbitrary convex set $K$. Let's see how to relate an arbitrary convex function $f(x)$ to a linear function by increasing the dimension by 1. Consider a convex set $\mathcal{K} \in \mathbb{R}^n$ and a continuous convex function $f: \mathbb{R}^n \rightarrow \mathbb{R}$. Consider the epigraph
$$
\tilde{\mathcal{K}}:=\{(x,y): x\in \mathcal{K}, y \geq f(x) \}.
$$
Define the linear function $\tilde{f}(x,y) = c \cdot (x,y)$, where $(x,y)\in \tilde{\mathcal{K}}$ and $c = ({\bf 0},1) \in \mathbb{R}^{n+1}$. We have
$$
\mathbb{E} \tilde{f}(X,Y) \geq \mathbb{E} f(X) \quad \text{the first one increase mass on large values}
$$
$$
\min_{(x,y)\in \tilde{\mathcal{K}}} \tilde{f}(x,y) = \min_{x \in \mathcal{K}} f(x)
$$
Thus
$$
\mathbb{E} f(X) - \min_{x\in\mathcal{K}} f(x) \leq \mathbb{E} \tilde{f}(X,Y) - \min_{(x,y)\in \tilde{\mathcal{K}}} \tilde{f}(x,y) \leq (n+1)T
$$
proof completed.
For the second claim, it is not hard to see that
$$
\mathbb{E}_F f(X) - \min_{x \in \mathcal{K}}f(x) = \mathbb{E}_F \left[f(X) - \min_{x \in \mathcal{K}}f(x)\right].
$$
Since adding a constant to the function does not have an effect on the density, we can assume without loss of generality that $\min_{x \in K}f(x) = 0$. Thus we have
\begin{align*}
\mathbb{E}_F f(X) &= \frac{\int_{\mathcal{K}} f(x) \exp\{-F(x)/T\} dx}{\int_{\mathcal{K}} \exp\{-F(x)/T\} dx} \leq \frac{\int_{\mathcal{K}} f(x) \exp\{-f(x)/T\} dx \cdot \exp(\rho/T) }{\int_{\mathcal{K}} \exp\{-f(x)/T\} dx \cdot \exp(-\rho/T)} \\& \leq \exp(2\rho/T) \cdot \mathbb{E} f(X) \leq (n+1)T \cdot \exp(2\rho/T)
\end{align*}
and
$$
\mathbb{E}_F f(X) - \min_{x\in\mathcal{K}} f(x) \leq (n+1)T \cdot \exp(2\rho/T).
$$
\end{proof}
\begin{proof}[Proof of Corollary \ref{cor:4.5}]
We choose $\rho =\epsilon/n$ and $T_K=\rho$. Given the final temperature, $K=\sqrt{n}\log(n/\epsilon)$. The optimization guarantee follows from Theorem~\ref{eq:opt_guarantee}. The number of queries is $\mathcal{O}^*(n^3)$ for one sample in one phase (Theorem~\ref{thm:mixing_LC}) times $\mathcal{O}^*(n)$ samples per phase for rounding (Section~\ref{sec:rounding}) times $K=\mathcal{O}^*(\sqrt{n})$ phases. The resulting distribution, however, is only $e\gamma$-close to the distribution with density proportional to $\exp\{-nF/\epsilon\}$ (by Theorem~\ref{Thm:Warm}). The guarantee of Theorem~\ref{eq:opt_guarantee} holds for the latter distribution, and we need to upper bound the effect of having a sample from an almost-desired distribution. Thankfully, $\gamma$ enters logarithmically in oracle complexity. Since $f$ is $L$-Lipschitz and domain is bounded, the range of function values over $\mathcal{K}$ is bounded by $B=\mathcal{O}(nLR)$. Then $\gamma$ can be chosen as $\epsilon/B$, which again only impacts oracle complexity by terms logarithmic in $n, L, R$.
\end{proof}
\end{appendix}
\bibliographystyle{apalike}
|
\section{Approximation} \label{sec:approximation}
In this section we construct a real valued Lipschitzian auxiliary function
from an integral varifold in a cylinder whose first variation is representable
by integration, see \ref{lemma:mini-lip-approx}. This auxiliary function
captures information on the height-excess measured in Lebesgue and Orlicz
spaces and the quadratic tilt-excess of the varifold. In conjunction with the
basic coercive estimate in \ref{lemma:coercive-estimate} and the interpolation
inequalities of Section \ref{sec:interpolation} it will be used in
\ref{lemma:app-coercive} to obtain a coercive estimate involving an
approximate height quantity.
The auxiliary function is constructed as ``upper envelope of the
modulus'' of an approximating Lipschitzian $\mathbf{Q}_Q ( \mathbf{R}^{n-m}
)$~valued function constructed in
\cite[3.15]{snulmenn.poincare}. Approximations by $\mathbf{Q}_Q (
\mathbf{R}^{n-m})$~valued functions are a powerful tool, originating from
Almgren \cite[3.1--3.12]{MR1777737}, whose handling is at times
complex. The fact that in the present setting we are able to
encapsulate their usage in the construction of the real valued
auxiliary function considerably simplifies our proof of decay rates
for the quadratic tilt-excess in \ref{theorem:quadratic-tilt-decay}.
\begin{definition}[see Almgren \protect{\cite[1.1\,(1)\,(3), 2.3\,(2)]{MR1777737}}] \label{miniremark:almgren1}
Suppose $Q$ is a positive integer and $Y$ is a finite dimensional
inner product space. Then
\begin{gather*}
{\textstyle \mathbf{Q}_Q (Y) = \left \{ \sum_{i=1}^Q \llbracket y_i \rrbracket
\with y_1, \ldots, y_Q \in Y \right \}}
\end{gather*}
is metrised by $\mathscr{G}$ such that, whenever $y_1,\ldots,y_Q \in Y$
and $\upsilon_1,\ldots,\upsilon_Q \in Y$, $\mathscr{G} \big ( \sum_{i=1}^Q
\llbracket y_i \rrbracket , \sum_{i=1}^Q \llbracket \upsilon_i \rrbracket \big)$
equals the infimum of the set of numbers
\begin{gather*}
{\textstyle \left ( \sum_{i=1}^Q \left | y_i - \upsilon_{\pi(i)} \right
|^2 \right)^{1/2}}
\end{gather*}
corresponding to all permutations $\pi$ of $\{ 1, \ldots, Q \}$.
\end{definition}
\begin{definition}[see Almgren \protect{\cite[1.1\,(9)\,(10)]{MR1777737}}]
\label{miniremark:almgren2}
Suppose ${m}$ and $Q$ are positive integers and $Y$ is a finite
dimensional inner product space.
A function $f : \mathbf{R}^{m} \to \mathbf{Q}_Q(Y)$ is called \emph{affine} if
and only if there exist affine functions $f_i : \mathbf{R}^{m} \to Y$
corresponding to $i = 1,\ldots,Q$ such that
\begin{gather*}
f (x) = {\textstyle\sum_{i=1}^Q} \llbracket f_i (x) \rrbracket
\quad \text{whenever $x \in \mathbf{R}^{m}$}
\end{gather*}
and in this case $\| f \| = \Lip f$. Moreover, if $a \in A \subset
\mathbf{R}^{m}$ and $f : A \to \mathbf{Q}_Q ( Y )$ then $f$ is \emph{affinely
approximable at $a$} if and only if $a \in \Int A$ and there exists an
affine function $g : \mathbf{R}^{m} \to \mathbf{Q}_Q ( Y)$ such that
\begin{gather*}
g(a) = f(a) \quad \text{and} \quad \lim_{x \to a} \mathscr{G} ( f(x),
g(x) ) / |x-a| = 0.
\end{gather*}
The function $g$ is unique and denoted by $\Aff f(a)$. The concept of
\emph{approximate affine approximability} is obtained through replacement
of the condition $a \in \Int A$ by $a \in A$ and replacement of $\lim$ by
$\aplim$. The corresponding affine function is denoted by $\ap \Aff f(a)$.
\end{definition}
\begin{remark}
\label{remark:apAf-apDf}
In comparison to Almgren \cite[1.1\,(10)]{MR1777737}, the requirement
``$g(a)=f(a)$'' has been added. Consequently, [approximate] affine
approximability implies [approximate] continuity. Moreover, supposing $Q =
1$ and denoting by $i : \mathbf{Q}_1 (Y) \to Y$ the canonical isometry, the
function $f$ is [approximately] affinely approximable at $a$ if and only if
$i \circ f$ is [approximately] differentiable at $a$, see \cite[3.1.1,
3.1.2]{MR41:1976}, and in this case
\begin{gather*}
i \circ \ap \Aff f(a) = i(f(a)) + \ap \Der (i \circ f)(a).
\end{gather*}
\end{remark}
\begin{miniremark}
\label{miniremark:Q-affine}
Suppose $Q$ is a positive integer, $Y$ is a finite dimensional inner product
space, $a \in \mathbf{R}^{m}$, and $f : \mathbf{R}^{m} \to \mathbf{Q}_Q (Y)$ is
affine. Then
\begin{gather*}
\Lip f = \limsup_{x \to a } |x-a|^{-1} \mathscr{G} (f(x),f(a))
= \aplimsup_{x \to a} |x-a|^{-1} \mathscr{G} (f(x),f(a));
\end{gather*}
in fact, in view of \cite[1.1\,(9)]{MR1777737} only the last equation needs
to be proven. For this purpose denote the approximate limit superior by
$\lambda$ and define the Lipschitzian function $g : \mathbf{R}^{m} \to \mathbf{R}$ by
$g(x) = \mathscr{G} (f(x),f(a)) - \lambda |x-a|$ for $x \in \mathbf{R}^{m}$. One
infers $\ap \Der g^+ (a) = 0$, whence it follows $\Der g^+ (a) =0$ by
\cite[3.1.5]{MR41:1976} with $C$, $B$, $f$, $\eta$, and $M$ replaced by
$\mathbf{R}^{m}$, $\mathbf{R}^{m}$, $g^+$, $1$, and $\Lip g$; therefore $\limsup_{x
\to a} |x-a|^{-1} \mathscr{G}(f(x),f(a)) \leq \lambda$. The reverse
inequality follows since $a \in \Int \dmn f$.
\end{miniremark}
\begin{miniremark}
\label{miniremark:Q-affine-approx}
Suppose $Q$ is a positive integer, $Y$ is a finite dimensional inner product
space, $a \in \mathbf{R}^{m}$, $f$ maps a subset of $\mathbf{R}^{m}$ into
$\mathbf{Q}_Q(Y)$, and $f$ is approximately affine approximable at~$a$. Then
\ref{miniremark:Q-affine} implies
\begin{gather*}
\| \ap \Aff f(a) \| = \aplimsup_{x \to a} |x-a|^{-1} \mathscr{G}(f(x),f(a)).
\end{gather*}
\end{miniremark}
\begin{lemma}
\label{lemma:q-sup-function}
Suppose $Q$ is a positive integer, $Y$ is a finite dimensional inner product
space, and $\sigma : \mathbf{Q}_Q ( Y ) \to \mathbf{R}$ satisfies
\begin{gather*}
\sigma (S) = \sup \{ |y| \with y \in \spt S \}
\quad \text{for $S \in \mathbf{Q}_Q (Y)$}.
\end{gather*}
Then $\Lip \sigma \leq 1$.
\end{lemma}
\begin{proof}
One may express $\sigma = p \circ \xi \circ g$, where $g : \mathbf{Q}_Q(Y) \to
\mathbf{Q}_Q(\mathbf{R})$ denotes the push forward induced by the norm on $Y$ mapping
$Y$ into $\mathbf{R}$, and $\xi : \mathbf{Q}_Q( \mathbf{R} ) \to \mathbf{R}^Q$ and $p : \mathbf{R}^Q
\to \mathbf{R}$ are characterised by
\begin{gather*}
\xi \big ( \tsum{i=1}{Q} \llbracket y_i \rrbracket \big ) = (y_1, \ldots, y_Q )
\quad \text{if $y_i \leq y_{i+1}$ for $i = 1, \ldots, Q-1$},
\\
p ( y_1, \ldots, y_Q ) = y_Q
\end{gather*}
whenever $(y_1, \ldots, y_Q ) \in \mathbf{R}^Q$. Clearly, $\Lip p \leq 1$.
Moreover, one readily verifies $\Lip g \leq 1$. Finally, $\Lip \xi \leq 1$
by Almgren \cite[1.1\,(4)]{MR1777737}.
\end{proof}
\begin{lemma}
\label{lemma:comp-Q-valued}
Suppose $Q$ is a positive integer, $Y$ is a finite dimensional inner product
space, $a \in \mathbf{R}^{m}$, $f$ maps a subset of $\mathbf{R}^{m}$ into
$\mathbf{Q}_Q(Y)$, and $\sigma : \mathbf{Q}_Q (Y) \to \mathbf{R}$ is Lipschitzian.
Then the following two statements hold.
\begin{enumerate}
\item \label{item:comp-Q-valued:classical} If $f$ is affinely approximable
at~$a$ and $\sigma \circ f$ is differentiable at~$a$, then $| \Der (\sigma
\circ f)(a)| \leq \Lip (\sigma) \| \Aff f(a) \|$.
\item \label{item:comp-Q-valued:approximate} If $f$ is approximately
affinely approximable at~$a$ and $\sigma \circ f$ is approximately
differentiable at~$a$, then $| \ap \Der ( \sigma \circ f ) (a) | \leq
\Lip(\sigma) \| \ap \Aff f(a) \|$.
\end{enumerate}
\end{lemma}
\begin{proof}
\eqref{item:comp-Q-valued:approximate} is a consequence
of~\ref{miniremark:Q-affine-approx} together with~\ref{remark:apAf-apDf}
and implies \eqref{item:comp-Q-valued:classical}.
\end{proof}
\begin{miniremark}
\label{remark:unitmeasure}
Notice that
\begin{gather*}
\sup \{ \unitmeasure{{m}} \with {m} \in \mathscr{P} \} < 6;
\end{gather*}
in fact, using $3 < \boldsymbol{\Gamma} ( 1/2 )^2 < 3.2$ and $({m}+2)
\unitmeasure{{m}+2} = 2 \boldsymbol{\Gamma} (\frac{1}{2})^2
\unitmeasure{{m}}$ for ${m} \in \mathscr{P}$ by \cite[3.2.13]{MR41:1976}, one
obtains $\unitmeasure{6} = \frac{1}{6} \boldsymbol{\Gamma}( \frac{1}{2} )^6
< \frac{8}{15} \boldsymbol{\Gamma} ( \frac{1}{2} )^4 = \unitmeasure{5} < 6$
and the supremum does not exceed $\unitmeasure{5}$.
\end{miniremark}
\begin{theorem}
\label{lemma:mini-lip-approx}
Suppose ${m}, {n}, Q \in \mathscr{P}$ and $1 <{m} < {n}$.
Then there exists a positive, finite number $\Gamma$ with the following
property.
If $0 < r < \infty$, $T = \im \mathbf{p}^\ast$, $V \in \IVar_{m} ( \mathbf{R}^{n}
\cap \oball{0}{4r} )$,
\begin{gather*}
(Q-1/2) \unitmeasure{{m}} r^{m}
\leq \| V \| ( \cylinder{T}{0}{r}{r} )
\leq (Q+1/2) \unitmeasure{{m}} r^{m},
\\
\| V \| ( \cylinder{T}{0}{r}{2r} \without \cylinder{T}{0}{r}{r/2} )
\leq (1/2) \unitmeasure{{m}} r^{m},
\\
\measureball{\| V \|}{\oball{0}{4r}} \leq Q \unitmeasure{{m}}
(5r)^{m}, \\
\eta = \| \delta V \| ( \oball 0{4r} )^{{m}/({m}-1)} + \tint{}{}
\| \project{S} - \project{T} \|^2 \ensuremath{\,\mathrm{d}} V (z,S),
\end{gather*}
$H$ consists of all $z \in \cylinder T0rr$ such that
\begin{gather*}
\measureball{\| V \|}{\cball{z}{s}}
\geq (40\isoperimetric{{m}}{m})^{-{m}} s^{m}
\quad \text{whenever $0 < s < 2r$},
\end{gather*}
and $\Phi$ is as in \ref{miniremark:Phi}, then there exists a Borel
subset~$X$ of $\mathbf{R}^{m} \cap \cball{0}{r}$ and a function $f : X \to \mathbf{R}$
with $\Lip f \leq 1$ satisfying the following five conditions whenever $1
\leq q \leq \infty$ and $A$ is a subset of~$X$:
\begin{enumerate}
\item \label{item:mini-lip-approx:bad-set} $\mathscr{L}^{m} ( \cball 0r
\without X ) \leq \Gamma \eta$.
\item \label{item:mini-lip-approx:2-height-control} $\eqLpnorm{\| V \|
\mathop{\llcorner} H}{q} {\perpproject{T}} \leq \Gamma \big (
\eqLpnorm{\mathscr{L}^{m} \mathop{\llcorner} X}{q}{f} + \eta^{1/q+1/{m}} \big
)$.
\item \label{item:mini-lip-approx:Phi-height-control} $\eqLpnorm{\| V \|
\mathop{\llcorner} H}{r^{-{m}} \Phi} { \perpproject{T} } \leq \Gamma \big (
\eqLpnorm{\mathscr{L}^{m} \mathop{\llcorner} X}{r^{-{m}} \Phi} f +
\eta^{1/{m}} \big )$.
\item \label{item:mini-lip-approx:inverse-2-height-control}
$\eqLpnorm{\mathscr{L}^{m} \mathop{\llcorner} A}{2}{f} \leq \eqLpnorm{\| V \|
\mathop{\llcorner} H \cap \mathbf{p}^{-1} [ A ]}{2}{\perpproject{T}}$.
\item \label{item:mini-lip-approx:tilt-control} $\eqLpnorm{\mathscr{L}^{m}
\mathop{\llcorner} X}{2}{\ap \Der f} \leq (2Q \eta)^{1/2}$.
\end{enumerate}
\end{theorem}
\begin{proof}
Notice that $(\isoperimetric{{m}} {m} )^{-{m}} \leq \unitmeasure
{m}$, see for instance~\cite[2.4]{snulmenn.isoperimetric}. Define $\beta
= {m}/({m}-1)$, and
\begin{gather*}
\Delta_1 = \varepsilon_{\text{\cite[3.15]{snulmenn.poincare}}}
\left( {n-m}, {m}, Q, 1, 5^{m} Q, \tfrac 12, \tfrac 12, \tfrac 12, \tfrac 14,
(40 \isoperimetric{{m}}{m})^{-{m}}/\unitmeasure{{m}} \right),
\\
\Delta_2 = ( \log 2 )^{1/\beta},
\quad
\Delta_3 = 1/ \Phi^{-1} ( 1/(6(Q+1))),
\quad
\Delta_4 = ( 20 \isoperimetric{{m}} {m} )^{-{m}} \Delta_1^\beta,
\\
\Delta_5 = \Delta_1^2 ( Q \unitmeasure{{m}} {n})^{-1}
( 60 \isoperimetric{{m}} {m} )^{-2{m}},
\\
\Delta_6 = \sup \{ 3 + 2Q + (12Q + 6) 5^{m}, 8 (Q+2) \},
\\
\Delta_7 = (1/2) \unitmeasure{{m}}
\lambda_{\text{\cite[3.15\,(4)]{snulmenn.poincare}}} ( {m}, 1/2, 1/4)^{m} 6^{-{m}},
\\
\Delta_8 = \sup \{ \Gamma_{\text{\cite[3.15\,(6)]{snulmenn.poincare}}} ( {m} ),
2 \unitmeasure{{m}}^{-1/{m}} \},
\quad
\Delta_9 = \Delta_6 {n} \besicovitch{{n}} \sup \{ 1,\Delta_1^{-2} \},
\\
\Delta_{10} = (12)^{{m}+1} Q \sup \{ Q, \Delta_8 \Delta_9^{1+1/{m}} \},
\quad
\Delta_{11} = \inf \{ 1, \Delta_4, \Delta_5, \Delta_9^{-1} \Delta_7 \},
\\
\Delta_{12} = 2 \Delta_{10} \Delta_2^{-1},
\quad
\Delta_{13} = \sup \{ 6 (Q+1) \Delta_{11}^{-1-1/{m}}, \Delta_3 \Delta_{11}^{-1/{m}} \},
\\
\Gamma = \sup \{ \Delta_9, \Delta_{10}, \Delta_{12}, \Delta_{13} \}.
\end{gather*}
Notice that $\Delta_2 < 1 \leq \Delta_9$.
Suppose $r$, $T$, $V$, $\eta$, $H$, and $\Phi$ are related to ${m}$,
${n}$, and $Q$ as in the body of the lemma. Since the statement of the
lemma is invariant by replacing $V$, $f$ with
$(\boldsymbol{\mu}_{(1/r)})_{\#} V$, $r^{-1} f \circ \boldsymbol{\mu}_{r}$,
we can assume $r=1$.
One may also assume $\eta \leq \Delta_{11}$ since otherwise
\begin{gather*}
\measureball{\mathscr{L}^{m}}{\cball 01}
\leq 6 \leq \Delta_{13} \Delta_{11} \leq \Gamma \eta,
\\
\eqLpnorm{\| V \| \mathop{\llcorner} \cylinder T011}{q}{ \perpproject{T} }
\leq 6 (Q+1) \leq \Delta_{13} \Delta_{11}^{1+1/{m}} \leq \Gamma \eta^{1/q+1/{m}},
\\
\eqLpnorm{\| V \| \mathop{\llcorner} \cylinder T011}{\Phi}{ \perpproject{T} }
\leq \Delta_3 \leq \Delta_{13} \Delta_{11}^{1/{m}} \leq \Gamma \eta^{1/{m}}
\end{gather*}
by \ref{remark:unitmeasure} and \ref{miniremark:Phi}, hence one may take
$X = \varnothing$ and $f = \varnothing$.
One applies \cite[3.15]{snulmenn.poincare} with
\begin{gather*}
\text{$m$, $n$, $L$, $M$, $\delta_1$, $\delta_2$, $\delta_3$,
$\delta_4$, $\delta_5$, $a$, $h$, $\mu$, and $\varepsilon_1$}
\quad \text{replaced by} \\
\text{${n-m}$, ${m}$, $1$, $5^{m} Q$, $\tfrac 12$, $\tfrac 12$, $\tfrac 12$, $\tfrac 14$,
$(40 \isoperimetric{{m}}{m})^{-{m}}/\unitmeasure{{m}}$, $0$,
$r$, $\| V \|$, and $\Delta_1$}
\end{gather*}
to obtain $\bar B$, $\bar{f}$ and $\bar{H}$ named $B$, $f$ and $H$ there.
First, \emph{it will be shown that $H = \bar{H}$}; in fact, noting $\eta
\leq \inf \{ \Delta_4, \Delta_5 \}$, one estimates
\begin{gather*}
\measureball{\| \delta V \|}{ \oball{z}{2} } \leq \eta^{1/\beta} \leq
\Delta_4^{1/\beta} \leq \Delta_1 \| V \|( \oball{z}{2} )^{1/\beta}, \\
\begin{aligned}
\tint{\oball{z}{2} \times \grass{{n}}{{m}}}{} | \project{S}-
\project{T}| \ensuremath{\,\mathrm{d}} V(z,S) & \leq \| V \| ( \oball 04)^{1/2}
{n}^{1/2} \eta^{1/2} \\
& \leq ( Q \unitmeasure{{m}})^{1/2} 3^{m} {n}^{1/2}
\Delta_5^{1/2} \leq \Delta_1 \measureball{\| V \|}{\oball z2}.
\end{aligned}
\end{gather*}
whenever $z \in \cylinder T011$ and $\measureball{\| V \|}{ \oball z2 }
\geq ( 20 \isoperimetric{{m}} {m} )^{-{m}}$.
Choose a Borel subset $X$ of $\dmn \bar{f}$ with $\mathscr{L}^{m} ( (\dmn
\bar f) \without X ) = 0$ and define $f : X \to \mathbf{R}$ by
\begin{gather*}
f(x) = \sup \{ |y| \with y \in \spt \bar f(x) \} \quad \text{whenever
$x \in X$}.
\end{gather*}
Clearly, $\Lip f \leq \Lip \bar f \leq 1$ by \ref{lemma:q-sup-function} and
one infers
\begin{gather*}
| \ap \Der f(x) | \leq \| \ap \Aff \bar f (x) \|
\quad \text{for $\mathscr{L}^{m}$ almost all $x \in X$} ;
\end{gather*}
in fact, $\bar f$ is approximately affinely approximable at $\mathscr{L}^{m}$
almost all $x \in X$ by~\cite[3.15\,(7a)]{snulmenn.poincare} and $f$ is
approximately differentiable by~\cite[2.8.18, 2.9.11, 3.1.8]{MR41:1976} at
$\mathscr{L}^{m}$ almost all $x \in X$ so the assertion follows from
\ref{lemma:q-sup-function} and
\ref{lemma:comp-Q-valued}\,\eqref{item:comp-Q-valued:approximate}.
Next, \emph{it will be proven that $\mathscr{L}^{m} ( \cball 01 \without
X ) \leq \Delta_9 \eta$}. For this purpose define sets $B_1$ and $B_2$
consisting of those $z \in \cylinder T011$ satisfying
\begin{gather*}
\text{$
\measureball{\| \delta V \|} { \cball zs } > \Delta_1 \| V \| ( \cball zs )^{1/\beta}
$ for some $0<s<2 $},
\\
\text{$
\tint{\cball zs \times \grass {n} {m}}{} \| \project{S} - \project{T} \|^2 \ensuremath{\,\mathrm{d}} V(z,S)
> \Delta_1^2 {n}^{-1} \measureball{\| V \|}{ \cball zs }
$ for some $0 < s < 2$}
\end{gather*}
respectively. To estimate $\| V \| ( B_1 )$ we employ the
Besicovitch-Federer covering theorem which provides disjointed families
$F_1, \ldots, F_{\besicovitch{{n}}}$ of closed balls such that
\begin{gather*}
{\textstyle B_1 \subset \bigcup \bigcup \{ F_i \with i = 1, \ldots, \besicovitch {n} \}
\subset \oball 04},
\\
\| V \| (C) < \Delta_1^{-\beta} \| \delta V \| ( C )^\beta
\quad \text{whenever $C \in F_i$ and $i = 1, \ldots, \besicovitch{{n}}$},
\end{gather*}
and we obtain
\begin{align*}
\| V \| ( B_1 ) & \leq \Delta_1^{-\beta}
\tsum{i=1}{\besicovitch {n}} \tsum{C \in F_i}{} \| \delta V \| ( C )^\beta
\\
& \leq \Delta_1^{-\beta} \tsum{i=1}{ \besicovitch {n} } \big(
\tsum{C \in F_i}{} \| \delta V \| ( C ) \big)^\beta
\leq \Delta_1^{-\beta} \besicovitch{{n}} \| \delta V \| ( \oball 04 )^\beta.
\end{align*}
In a similar fashion we find another disjointed families $F_1, \ldots,
F_{\besicovitch {n}}$ of closed balls such that
\begin{gather*}
{\textstyle B_2 \subset \bigcup \bigcup \{ F_i \with i = 1, \ldots, \besicovitch {n} \}
\subset \oball 04},
\\
\| V \| (C) < \Delta_1^{-2} {n}
\tint{C \times \grass {n} {m}}{} \| \project{S} - \project{T} \|^2 \ensuremath{\,\mathrm{d}} V(z,S)
\quad \text{for $C \in F_i$, $i = 1, \ldots, \besicovitch {n}$},
\end{gather*}
and in consequence
\begin{align*}
\| V \| ( B_2 ) & \leq \Delta_1^{-2} {n} \tsum{i = 1}{\besicovitch {n}}
\tsum{C \in F_i}{} \tint{C \times \grass {n} {m}}{} \| \project{S} - \project{T} \|^2 \ensuremath{\,\mathrm{d}} V(z,S)
\\
& \leq \Delta_1^{-2} {n} \besicovitch {n} \tint{}{} \| \project{S} - \project{T} \|^2 \ensuremath{\,\mathrm{d}} V(z,S).
\end{align*}
Verifying $\bar B \subset B_1 \cup B_2$ by means of H{\"o}lder's inequality,
the asserted estimate follows from \cite[3.15\,(3)]{snulmenn.poincare}.
Since in particular $\mathscr{L}^{m} ( \cball 01 \without X ) \leq
\Delta_9 \Delta_{11} \leq \Delta_7$, one applies
\cite[3.15\,(6)]{snulmenn.poincare} with $S$ replaced by $Q \llbracket 0
\rrbracket$ to estimate, concerning
\eqref{item:mini-lip-approx:2-height-control},
\begin{align*}
& \eqLpnorm{\| V \| \mathop{\llcorner} H}{q}{ \perpproject T} \\
& \qquad \leq (12)^{{m}+1}
Q \big ( Q^{1/2} \eqLpnorm{\mathscr{L}^{m} \mathop{\llcorner} X} qf
+ \Delta_8 \mathscr{L}^{m} ( \cball 01 \without X )^{1/q+1/{m}} \big ) \\
& \qquad \leq \Delta_{10} \big ( \eqLpnorm{\mathscr{L}^{m} \mathop{\llcorner} X} qf
+ \eta^{1/q+1/{m}} \big )
\quad \text{for $1 \le q \le \infty$.}
\end{align*}
Consequently, concerning \eqref{item:mini-lip-approx:Phi-height-control},
one notes that $\eta \leq 1$ and $\Phi ( \Delta_2 ) = 1$ and estimates
\begin{align*}
\tint{H}{} \Phi \circ | \gamma^{-1} \perpproject T | \ensuremath{\,\mathrm{d}} \| V \|
& = \tsum{i=1}{\infty} i!^{-1} \gamma^{-i\beta} \tint{H}{} | \perpproject{T} |^{\beta i} \ensuremath{\,\mathrm{d}} \| V \|
\\
& \leq \tfrac 12 \tsum{i=1}{\infty} i!^{-1} ( 2 \Delta_{10}/ \gamma )^{\beta i}
\big ( \tint{X}{} |f|^{\beta i} \ensuremath{\,\mathrm{d}} \mathscr{L}^{m} + \eta^{1+\beta i/{m}} \big )
\\
& = \tfrac 12 \tint{X}{} \Phi \circ | 2 \Delta_{10} \gamma^{-1} f | \ensuremath{\,\mathrm{d}} \mathscr{L}^{m}
+ \tfrac 12 \eta \Phi ( 2 \Delta_{10} \eta^{1/{m}} \gamma^{-1} ) \leq 1
\end{align*}
whenever $2 \Delta_{10} \Delta_2^{-1} \big ( \eqLpnorm{\mathscr{L}^{m}
\mathop{\llcorner} X}{\Phi}{f} + \eta^{1/{m}} \big ) < \gamma < \infty$, hence
\begin{gather*}
\eqLpnorm{\| V \| \mathop{\llcorner} H}{\Phi}{\perpproject{T}}
\leq \Delta_{12} \big ( \eqLpnorm{\mathscr{L}^{m} \mathop{\llcorner} X}{\Phi}{f}
+ \eta^{1/{m}} \big ).
\end{gather*}
To prove \eqref{item:mini-lip-approx:inverse-2-height-control} and
\eqref{item:mini-lip-approx:tilt-control}, recall
\begin{gather*}
H \cap \mathbf{p}^{-1} [ \dmn \bar f ]
= \{ z \with \mathbf{q} (z) \in \spt \bar f ( \mathbf{p} (z)) \}
\subset \{ z \with \boldsymbol{\Theta}^{m} ( \| V \|, z) \in \mathscr{P} \}
\end{gather*}
from \cite[3.15\,(2)\,(4)]{snulmenn.poincare} and observe: \emph{If $A$ is
a subset of $X$, $g$ is an $\mathscr{L}^{m} \mathop{\llcorner} A$ measurable real
valued function and $h$ is an $\| V \| \mathop{\llcorner} H \cap \mathbf{p}^{-1} [ A
]$ measurable real valued function such that
\begin{gather*}
\mathscr{L}^{m} ( A \cap \{x \with g(x)> t \} \without \mathbf{p} [ H
\cap \{ z \with h(z) > t \} ] ) = 0 \quad \text{for $0 < t <
\infty$},
\end{gather*}
then $\eqLpnorm{\mathscr{L}^{m} \mathop{\llcorner} A} q g \leq \eqLpnorm{\| V \|
\mathop{\llcorner} H \cap \mathbf{p}^{-1} [ A ]} qh$;} in fact
\begin{align*}
\mathscr{L}^{m} ( A \cap \{ x \with g(x) > t \} )
& \leq \mathscr{H}^{m} ( H \cap \mathbf{p}^{-1} [ A ] \cap \{ z \with h(z) > t \} )
\\
& \leq \| V \| ( H \cap \mathbf{p}^{-1} [ A ] \cap \{ z \with h(z) > t \} )
\end{align*}
by \cite[2.10.35]{MR41:1976} and Allard \cite[3.5\,(1b)]{MR0307015}.
One applies this observation with $g$ and $h$ replaced by $f$ and $\big |
\perpproject T | U \big |$ to deduce
\eqref{item:mini-lip-approx:inverse-2-height-control}. Recalling $|\ap
\Der f(x)| \le \|\ap \Aff f(x) \|$ for $\mathscr{L}^{m}$ almost all $x \in X$ together
with~\cite[3.15\,(7d)]{snulmenn.poincare}, one applies the observation once
more, with
\begin{gather*}
\text{$g(x)$ and $h(z)$}
\quad \text{replaced by} \quad
\text{$| \ap \Der f(x) |$ and $(2Q)^{1/2} \project{\| \Tan^{m} ( \| V \|, z )} - \project{T} \|$}
\end{gather*}
to infer \eqref{item:mini-lip-approx:tilt-control}.
\end{proof}
\section{A coercive estimate}
\label{sed:coercive}
In this section we provide, in \ref{lemma:coercive-estimate}, the first main
ingredient for the proof of the decay rates almost everywhere of the quadratic
tilt-excess of two dimensional integral varifolds whose first variation is
representable by integration, namely a coercive estimate. In this estimate the
quadratic tilt-excess is controlled by the variation measure of the first
variation and the height-excess. In order to be effective for the present
purpose, two aspects are crucial. Firstly, in the height-excess only the set of
points where the density ratio is bounded from below are taken into
account. Secondly, the height-excess term which is multiplied by a first
variation term is measured in the Orlicz space seminorm naturally corresponding
to square summable weak derivatives in two dimensions.
In the basic form of such coercive estimate all quantities are measured as
square integrals, see Allard~\cite[8.13]{MR0307015}. In \cite[5.5]{MR485012}
Brakke devised an interpolation procedure to obtain estimates in which the
variation is measured by its variation measure. This was further refined
in~\cite[4.14]{snulmenn.decay} by allowing the height-excess to be measured in
different Lebesgue spaces and in \cite[4.10]{snulmenn.decay} by restricting the
height-excess to the set of points where the density ratio is bounded from below
using a possibly discontinuous ``cut-off'' function, see
\ref{remark:tv-cutoff}. In \ref{lemma:coercive-estimate} we additionally refine
Brakke's interpolation procedure to include the relevant Orlicz space norm.
\begin{definition}
\label{def:orlicz}
If $\Phi : \{ t \with 0 \leq t < \infty \} \to \{ t \with 0 \leq t < \infty
\}$ is a nondecreasing convex function with $\Phi(0)=0$ and $\lim_{t \to
\infty} \Phi (t) = \infty$, $\mu$ measures $X$, and $Z$ is a Banach
space, then one defines the seminorm $\mu_{(\Phi)}$ on $\mathbf{A} ( \mu,
Z )$ by
\begin{gather*}
\Lpnorm{\mu}{\Phi}{f} = \inf \big \{ \lambda \with
\text{$0 < \lambda \leq \infty$, $\orlicz{\Phi}{\lambda^{-1}f}{\mu} \leq 1$} \big \}
\quad \text{for $f \in \mathbf{A} ( \mu, Z)$}.
\end{gather*}
\end{definition}
\begin{remark}
\label{remark:orlicz_eq}
Notice that $\Lpnorm{\mu}{\Phi}{f} = 0$ if and only if $f (x) = 0$ for
$\mu$ almost all $x$. Moreover, if $\Lpnorm{\mu}{\Phi}{f} > 0$ then
$\orlicz{\Phi}{\lambda^{-1}f}{\mu} \leq 1$ for $\lambda =
\Lpnorm{\mu}{\Phi}{f}$ and equality holds if $\tint{}{} \Phi \circ |
s^{-1} f | \ensuremath{\,\mathrm{d}} \mu < \infty$ for some $0 < s < \Lpnorm{\mu}{\Phi}{f}$.
\end{remark}
\begin{remark}
The functions $\Phi$ and $\mu_{(\Phi)}$ are a ``Young's function'' and
its corresponding ``Luxemburg norm'' in the terminology of \cite[Chapter
4, 8.1, 8.6]{MR928802}.
\end{remark}
\begin{remark}
\label{remark:basic_orlicz}
Suppose $\Phi$, $\mu$, $X$ and $Z$ are as in \ref{def:orlicz}. Then the
following basic properties hold.
\begin{enumerate}
\item
\label{item:basic_orlicz:mult}
If $0 < c < \infty$ and $f \in \mathbf{A} (\mu,Z)$, then
$\eqLpnorm{c\mu}{\Phi}{f} = \Lpnorm{\mu}{c \Phi}{f}$.
\item
\label{item:basic_orlicz:estimate}
If $0 < \varepsilon \leq 1$ and $f \in \mathbf{A} ( \mu,Z )$, then
$\varepsilon \Lpnorm{\mu}{\Phi}{f} \leq \Lpnorm{\mu}{\varepsilon
\Phi}{f}$.
\item
\label{item:basic_orlicz:push}
If $u : X \to Y$, $f : Y \to Z$, and $f \circ u \in \mathbf{A} (\mu,
Z)$, then $f \in \mathbf{A} ( u_\# \mu, Z)$ and $\eqLpnorm{u_\#
\mu}{\Phi}{f} = \Lpnorm{\mu}{\Phi}{f \circ u}$, see \cite[2.1.2,
2.4.18\,(1)]{MR41:1976}.
\end{enumerate}
\end{remark}
\begin{miniremark}
\label{miniremark:kappa}
Suppose $2 \leq {m} \in \mathscr{P}$ and $\kappa : \{ t \with 0 \leq t < \infty
\} \to \{ t \with 0 \leq t < \infty \}$ satisfies
\begin{gather*}
\kappa (0) = 0,
\qquad
\kappa(t) = t \big ( 1 + (\log(1 + 1/t))^{1-1/{m}} \big )
\quad
\text{for $0 < t < \infty$}.
\end{gather*}
Then one verifies that $\kappa$ is continuous increasing and concave, in
particular $\kappa ( \tau t ) \leq \tau \kappa (t)$ for $1 \leq \tau <
\infty$ and $0 \leq t < \infty$.
\end{miniremark}
\begin{miniremark}
\label{miniremark:Phi}
Suppose $2 \leq {m} \in \mathscr{P}$ and $\Phi : \{ t \with 0 \leq t < \infty
\} \to \{ 0 \leq t < \infty \}$ is defined by
\begin{gather*}
\Phi (t) = \exp \big ( t^{{m}/({m}-1)} \big ) -1 \quad \text{for
$0 \leq t < \infty$}.
\end{gather*}
Then $\Phi$ satisfies the conditions of \ref{def:orlicz}, and $\Phi$ maps
$\{ t \with 0 \leq t < \infty \}$ univalently onto $\{ t \with 0 \leq t <
\infty \}$ with
\begin{gather*}
\Phi^{-1} (t) = ( \log ( 1 + t ) )^{1-1/{m}} \quad \text{for $0
\leq t < \infty$}.
\end{gather*}
Therefore $\Lpnorm{\mu}{\Phi}{1} = 1 / \Phi^{-1}(1/\mu(X))$ whenever
$\mu$ measures $X$ and $0 < \mu (X) < \infty$ by \ref{remark:orlicz_eq}.
Notice, if $0 \leq \alpha < \infty$ then
\begin{gather*}
\inf \{ \alpha t + 1/\Phi(t) \with 0 < t < \infty \}
\leq \kappa ( \alpha ),
\end{gather*}
where $\kappa$ is as in \ref{miniremark:kappa}; in fact, consider $t =
\Phi^{-1} ( 1 / \alpha )$ if $\alpha > 0$.
\end{miniremark}
\begin{theorem}
\label{lemma:coercive-estimate}
Suppose ${m}$, ${n}$, $p$, $U$, and $V$ are as in
\ref{miniremark:situation_general}, $p = 1< {m}$, $\kappa$ and $\Phi$ are
related to ${m}$ as in~\ref{miniremark:kappa} and~\ref{miniremark:Phi},
$C$ and $K$ are compact subsets of $U$, $C \subset K$, $0 < r < \infty$, $H$
is the set of all $z \in \spt \| V \|$ such that
\begin{gather*}
\measureball{\| V \|}{\cball zs} \geq (40 \isoperimetric{{m}} {m} )^{-{m}} s^{m}
\quad \text{whenever $0 < s < \infty$, $\cball zs \subset K$},
\end{gather*}
$c \in \mathbf{R}^{n}$, $T \in \grass{{n}}{{m}}$, and $h : U \to \mathbf{R}$
satisfies $h(z) = \dist (z-c,T)$ for $z \in U$.
Then there holds
\begin{multline*}
r^{-{m}} \tint{\{ z \with \oball{z}{r} \subset C\} \times \grass{{n}}{{m}}}{}
\| \project{S} - \project{T} \|^2 \ensuremath{\,\mathrm{d}} V (z,S)
\leq \Gamma \big ( r^{-{m}} \| \delta V \| ( K)^{{m}/({m}-1)}
\\
+ \kappa ( r^{-{m}} \| \delta V \| ( K )
\eqLpnorm{\| V \| \mathop{\llcorner} C \cap H}{r^{-{m}} \Phi}{h} )
+ r^{-{m}-2} \tint{C \cap H}{} |h|^2 \ensuremath{\,\mathrm{d}} \| V \| \big ),
\end{multline*}
where $\Gamma$ is a positive, finite number depending only on ${m}$.
\end{theorem}
\begin{proof}
Assume $c = 0$, hence $h = \big |\perpproject{T}|U \big |$, and notice that
in view of \ref{remark:basic_orlicz}, \cite[3.2(2) and 4.12(1)]{MR0307015},
one may employ homotheties to reduce the problem to the case
$r=1$. Abbreviate $\mu = \eqLpnorm{\| V \| \mathop{\llcorner} C \cap H}{\Phi}{h}$ and
denote $(\| V \|, {m} )$ approximate differentials by ``$\ap \Der $''.
Select $\phi \in \mathscr{D} (U,\mathbf{R})$ with
\begin{gather*}
0 \leq \phi \leq 1, \quad \spt \phi \subset C, \quad \{ z \with \oball
z1 \subset C \} \subset \{ z \with \phi (z) = 1 \}, \quad | \Der \phi |
\leq 2.
\end{gather*}
Using \cite[4.7]{snulmenn.decay} with $\delta = \frac{1}{40}$, one obtains
a Borel function $f : U \to \{ t \with 0 \leq t \leq 1 \}$ with $f| U
\without K = 0$ such that the varifolds $V_1, V_2 \in \RVar_{m} ( U )$
defined by
\begin{gather*}
V_1 (A) = \tint{A}{\ast} f(z) \ensuremath{\,\mathrm{d}} V(z,S) \quad \text{for $A \subset U
\times \grass{{n}}{{m}}$}
\end{gather*}
and $V_2 = V-V_1$ satisfy
\begin{gather*}
\text{$f(z) = 1$ and $\ap \Der f(z) = 0$ for $\| V \|$ almost all $z \in U
\without H$}, \\
\tint{}{} \phi(z)^2 \| \project{S} - \project{T} \|^2 \ensuremath{\,\mathrm{d}} V_1 (z,S) \leq
4 \| V_1 \| (K) \leq \Delta \| \delta V \| (K)^{{m}/({m}-1)}, \\
\| \delta V_2 \| \leq (1-f) \| \delta V \| + | \ap \Der f | \| V \|, \quad
\| V \| ( | \ap \Der f| ) \leq (400)^{m} \| \delta V \| ( K ),
\end{gather*}
where $\Delta = 4
(400)^{{m}^2/({m}-1)}(\isoperimetric{{m}}{m})^{{m}/({m}-1)}$;
compare \cite[p.~24, l.~14--20]{snulmenn.decay}. In~particular, one infers
\begin{gather*}
\| V_2 \| \leq \| V \| \mathop{\llcorner} H, \quad \| \delta V_2 \| ( K ) \leq
(800)^{m} \| \delta V \| ( K ).
\end{gather*}
Defining $g = \phi^2 ( \perpproject{T}|U)$, one derives
\begin{gather*}
\tint{}{} \phi (z)^2 \| \project{S} - \project{T} \|^2 \ensuremath{\,\mathrm{d}} V_2 (z,S)
\leq \sup \big \{ 16 \tint{}{} |\Der \phi|^2 |h|^2 \ensuremath{\,\mathrm{d}} \| V_2 \|, 2 | (
\delta V_2 ) ( g ) | \big \}
\end{gather*}
as in Brakke~\cite[5.5, p.~139, l.~1--14]{MR485012}, hence
\begin{multline*}
\tint{\{ z \with \oball z1 \subset C \} \times \grass {n} {m}}{}
\| \project{S} - \project{T} \|^2 \ensuremath{\,\mathrm{d}} V(z,S) \\
\leq \Delta \| \delta V \| (K)^{{m}/({m}-1)} + 2 | ( \delta V_2 )
(g) | + 64 \tint{C \cap H}{} |h|^2 \ensuremath{\,\mathrm{d}} \| V \|.
\end{multline*}
If $\mu = 0$, then $g(z) = 0$ for $\| V \|$ almost all $z \in H$, hence
$\Der g(z)|\Tan^{m} ( \| V \|, z ) = 0$ for $\| V \|$ almost all $z \in H$
by \cite[2.10.19\,(4), 3.2.16]{MR41:1976} and $( \delta V_2 ) (g) = 0$.
Therefore one may assume $\mu > 0$.
In order to estimate $| ( \delta V_2 ) ( g ) |$, suppose $0 < t < \infty$,
define $\eta : \{ s \with 0 \leq s < \infty \} \to \mathbf{R}$ by
\begin{gather*}
\eta (0) = 1, \qquad \eta (s) = \inf \{ 1, t/s \} \quad \text{for $0
< s < \infty$}.
\end{gather*}
Moreover, let $Z = \{ z \with t < h(z) \}$ and define Lipschitzian maps by
\begin{gather*}
g_1 = \phi^2 ( \eta \circ h ) \perpproject{T}|U, \quad g_2 =
g-g_1.
\end{gather*}
Since $g_2 | U \without Z = 0$, one notices that
\begin{gather*}
\ap \Der g_2 (z) = 0 \quad \text{for $\| V \|$ almost all $z \in U
\without Z$}
\end{gather*}
by \cite[2.10.19\,(4)]{MR41:1976}. Additionally, one computes
\begin{gather*}
| g_1 (z) | \leq t \quad \text{for $z \in U$}, \qquad \| \Der
g_2 (z) \| \leq 2 \phi(z)^2 + | \Der \phi(z)|^2 h(z)^2 \quad
\text{for $z \in Z$},
\end{gather*}
see the case $r=1$ of \cite[4.10, p.~24, l.~26 -- p.~25,
l.~12]{snulmenn.decay}. It follows that
\begin{gather*}
\| \Der g_2 (z) \| \leq 2 \phi (z)^2 \Phi (t/\mu)^{-1} \Phi (
h(z)/\mu) + | \Der \phi(z)|^2 h(z)^2 \quad \text{for $z \in Z$}.
\end{gather*}
Therefore one estimates, using \cite[4.5\,(4)]{snulmenn.decay} and
\ref{remark:orlicz_eq},
\begin{align*}
| ( \delta V_2 ) (g) | & \leq t \| \delta V_2 \| ( K ) + 2 \Phi
(t/\mu)^{-1} \tint{C}{} \Phi \circ | \mu^{-1} h | \ensuremath{\,\mathrm{d}} \| V_2 \| + 4
\tint{C}{} |h|^2 \ensuremath{\,\mathrm{d}} \| V_2 \| \\
& \leq (800)^{m} t \| \delta V \| (K) + 2 \Phi(t/\mu)^{-1} + 4
\tint{C \cap H}{} |h|^2 \ensuremath{\,\mathrm{d}} \| V \|,
\end{align*}
hence, noting \ref{miniremark:Phi} with $\alpha = \|\delta V\|(K) \mu$, one
may take $\Gamma = \Delta + (1600)^{m}$.
\end{proof}
\begin{remark} \label{remark:tv-cutoff}
Notice that the function $f$ furnished by \cite[4.7]{snulmenn.decay} will
necessarily be discontinuous in some cases, see
\cite[4.8]{snulmenn.decay}. However, inspecting the proof of
\cite[4.7]{snulmenn.decay} and using \cite[8.7]{snulmenn:tv.v2}, one is at
least assured that $f$ is generalised weakly differentiable in the sense
of \cite[8.3]{snulmenn:tv.v2} with
\begin{gather*}
\derivative{V}{f} (z) = ( \| V \|, {m} ) \ap \Der f(z) \circ \project {
\Tan^{m} ( \| V \|, z ) } \quad \text{for $\| V \|$ almost all $z$}.
\end{gather*}
\end{remark}
\section{Differentiation results}
\label{sec:differentiation}
In the present section a differentiation theorem for measures relative to
varifolds is provided. It slightly generalises
\cite[3.1]{snulmenn.isoperimetric} so as to become applicable in the proof of
the sharp decay rate almost everywhere for the quadratic tilt-excess of
certain two-dimensional varifolds in Section~\ref{sec:quadratic_tilt_decay}.
Additionally, a corollary for use in the study of one-dimensional varifolds in
Section~\ref{sec:one_dimensional_decay} is noted.
Results of the type considered here occur for instance in Calder{\'o}n and
Zygmund \cite[Theorem~10, p.~189]{MR0136849} in which paper a differentiation
theory of higher order in Lebesgue spaces is developed. The classical Rademacher
theorem is contained in that theory as special case, see Calder{\'o}n and
Zygmund \cite[Theorem~12, p.~204]{MR0136849}.
\begin{miniremark}
\label{miniremark:situation_general}
Suppose ${m}, {n} \in \mathscr{P}$, ${m} \leq {n}$, $1 \leq p \leq \infty$,
$U$ is an open subset of $\mathbf{R}^{n}$, $V \in \Var_{m} ( U)$, $\| \delta V
\|$ is a Radon measure, $\boldsymbol{\Theta}^{m} ( \| V \|, z) \geq 1$ for $\| V \|$
almost all $z$. If $p > 1$, then suppose additionally that $\mathbf{h} ( V,
\cdot ) \in \Lploc{p} ( \| V \|, \mathbf{R}^{n})$ and
\begin{gather*}
\delta V (\theta) = - \tint{}{} \mathbf{h} (V,z) \bullet \theta(z) \ensuremath{\,\mathrm{d}}
\| V \| z \quad \text{for $\theta \in \mathscr{D} ( U, \mathbf{R}^{n})$}.
\end{gather*}
Therefore $V \in \RVar_{m} (U)$ by Allard \cite[5.5\,(1)]{MR0307015}.
If~$p = 1$ let $\psi = \| \delta V \|$. If $1 < p < \infty$ define a Radon
measure $\psi$ over $U$ by the requirement $\psi (B) = \tint{B}{} |
\mathbf{h} ( V, z ) |^p \ensuremath{\,\mathrm{d}} \| V \| z$ whenever $B$ is Borel subset of~$U$.
\end{miniremark}
\begin{theorem}
\label{thm:O_o}
Suppose ${m}$, ${n}$, $p$, $U$ and $V$ are as in
\ref{miniremark:situation_general}, $1 \le p \le {m}$, $\omega$ is a
modulus of continuity, $\mu$ measures $U$ with $\mu(U \without \spt \|V\|) =
0$, and $Z$ is a $\|V\|$ measurable set with $\mu(Z) = 0$. In case $p <
{m}$ suppose additionally that there exist $1 < q \leq \infty$ and $f \in
\Lploc{q} (\|V\|)$ such that
\begin{gather*}
\liminf_{r \to 0+} r^{(1-1/q){m} p/(p-{m})} \omega (r) > 0, \\
\mu (B) = \tint{B}{} f \ensuremath{\,\mathrm{d}} \|V\| \quad \text{whenever $B$ is Borel subset of $U$.}
\end{gather*}
Then for $\mathscr{H}^{m} $ almost all $z \in Z$
\begin{gather*}
\limsup_{r \to 0+} r^{-{m}} \omega(r)^{-1} \measureball{\mu}{\cball
zr} \quad \text{equals either $0$ or $\infty$}.
\end{gather*}
\end{theorem}
\begin{proof}
\setcounter{equation}{0}
For $i \in \mathscr{P}$ let $G_i$ denote the set of all $z \in \spt \| V \|$ such
that either $\oball z{1/i} \not \subset U$ or
\begin{gather*}
\measureball{\| \delta V \|}{\cball zr } > (2 \isoperimetric{{m}})^{-1} \|V\|(\cball zr)^{1-1/{m}}
\quad
\text{for some $0 < r < 1/i$}.
\end{gather*}
Notice that $G_{i+1} \subset G_i$ and that $G_i$ is relatively open in $\spt
\| V \|$ by an argument analogous to \cite[2.9.14]{MR41:1976}.
We start with some preliminary observations. Consider the \emph{case $p <
{m}$}. If $q < \infty$, define $\nu$ to be the Radon measure over $U$
characterised by the requirement $\nu (B) = \tint{B}{} f^q \ensuremath{\,\mathrm{d}} \| V \|$
whenever $B$ is a Borel subset of $U$. Using H{\"o}lder's inequality, one
may employ \cite[2.9, 2.10]{snulmenn.isoperimetric} with $m$, $n$, $\mu$,
$s$, $\varepsilon$, and $\Gamma$ replaced by ${n-m}$, ${m}$, $\| V \|$,
${m}$, $(2 \isoperimetric{{m}} )^{p/(p-{m})}$, and $5 {m} \gamma (
{m} )$ to see that for $\mathscr{H}^{m}$ almost all $z \in U$ there exists an
$i \in \mathscr{P}$ such that
\begin{gather*}
\boldsymbol{\Theta}^{{m}^2/({m}-p)} ( \| V \| \mathop{\llcorner} G_i, z ) = 0 .
\end{gather*}
This implies, by another use of H{\"o}lder's inequality, that for $\mathscr{H}^{m}$
almost all $z \in U$ there exists an $i \in \mathscr{P}$ such that
\begin{gather*}
\boldsymbol{\Theta}^{{m}(1+(1-1/q)p/({m}-p))} ( \mu \mathop{\llcorner} G_i, z) = 0 ,
\end{gather*}
since, if $q < \infty$, then $\boldsymbol{\Theta}^{\ast {m}} ( \nu, z ) < \infty$ for
$\mathscr{H}^{m}$ almost all $z \in U$ due
to~\cite[2.10.19\,(3)]{MR41:1976}. Observe that if $z \in \spt \|V\|
\without \bigcup_{i=1}^{\infty} G_i$, then, according to
\cite[2.5]{snulmenn.isoperimetric}, $\boldsymbol{\Theta}_*^{m}(\|V\|,z) > 0$;
hence, $\spt \| V \| \cap \{ z \with \boldsymbol{\Theta}^{m} ( \| V \|, z ) = 0
\} \subset \bigcap_{i=1}^\infty G_i$. Recalling that $G_i$ are relatively
open in $\spt \|V\|$ and $\mu(U \without \spt \|V\|) = 0$, one infers that
\begin{gather*}
\boldsymbol{\Theta}^{m}( \| V \|, z ) = 0
\quad \text{implies} \quad
\boldsymbol{\Theta}^{{m}(1+(1-1/q)p/({m}-p))} ( \mu, z) = 0
\end{gather*}
for $\mathscr{H}^{m}$ almost all $z \in U$. In \emph{case $p = {m}$},
one notices that
\begin{gather*}
\boldsymbol{\Theta}^{m}_\ast ( \| V \|, z ) \geq 1/2
\quad \text{whenever $z \in \spt \| V \|$},
\\
{
\textstyle \spt \| V \| \cap
\big \{ z \with \| \delta V \| ( \{ z \} ) < (2\isoperimetric{{m}})^{-1} \big \}
\subset U \cap \bigcup_{i=1}^\infty \big\{ c \with G_i \cap \cball{c}{1/i} = \varnothing \big\} ,
}
\\
\spt \| V \| \cap \{ z \with \| \delta V \| ( \{ z \} ) \ge (2\isoperimetric{{m}})^{-1}\} \text{ is countable}
\end{gather*}
by \cite[4.8\,(4), 7.6]{snulmenn:tv.v2} and
\cite[2.5]{snulmenn.isoperimetric}
Therefore, one may assume in both cases that $Z \subset \{ z \with
\boldsymbol{\Theta}^{\ast {m}} ( \| V \|, z ) > 0 \}$. As this implies that $Z$ is
a~union of countably many $\mathscr{H}^{m}$~measurable sets of finite
$\mathscr{H}^{m}$~measure by~\cite[2.10.19\,(3)]{MR41:1976}, one may also
assume that $Z$ is compact.
Define sets
\begin{gather*}
Z_j = Z \cap \{ z \with \measureball{\mu}{\cball zr} \leq j r^{{m}} \omega(r+) \text{ for all } 0 < r < 1/j \}
\end{gather*}
whenever $j \in \mathscr{P}$ and $1/j < \dist ( Z, \mathbf{R}^{n} \without U)$ whose
union contains
\begin{gather*}
Z \cap \Big\{ z \with \limsup_{r \to 0+} r^{-{m}} \omega(r)^{-1} \measureball{\mu}{\cball zr} < \infty \Big\}
\end{gather*}
and observe that the sets $Z_j$ are compact, see~\cite[2.9.14]{MR41:1976}.
Hence, it is sufficient to prove for each $j \in \mathscr{P}$ with $1/j < \dist ( Z,
\mathbf{R}^{n} \without U )$ that
\begin{gather*}
\lim_{r \to 0+} r^{-{m}} \omega(r)^{-1} \measureball{\mu}{\cball cr}
= 0 \quad \text{for $\mathscr{H}^{m} $ almost all $c \in Z_j$}.
\end{gather*}
In fact, this equality will be proven whenever $c \in Z_j$ and $j \in \mathscr{P}$
with $1/j < \dist (Z, \mathbf{R}^{n} \without U )$ satisfy for some $i \in \mathscr{P}$
with $i \geq 2j$ that $\boldsymbol{\Theta}^{m} ( \|V\| \mathop{\llcorner} U \without Z_j, c )
= 0$ and
\begin{gather*}
\boldsymbol{\Theta}^{{m}^2/({m}-p)} ( \|V\| \mathop{\llcorner} G_i, c ) = 0
\quad
\text{if $p < {m} $},
\qquad G_i \cap \cball{c}{1/i} = \varnothing
\quad \text{if $p = {m}$}
\end{gather*}
as $\mathscr{H}^{m}$ almost all $c \in Z_j$ do according to
\cite[2.10.19\,(4)]{MR41:1976} and the second paragraph of this proof.
For such $c$, $j$, and $i$, suppose $0 < \varepsilon \leq ( 6 {m}
\isoperimetric{{m}} )^{-1}$, $0 < r < 1/i$, and
\begin{gather*}
\| V \| ( \cball{c}{2r} \without Z_j ) < \varepsilon^{m} r^{m}.
\end{gather*}
Whenever $\zeta \in \cball cr \cap ( \spt \|V\| ) \without ( G_i \cup
Z_j)$ and $s = \dist (\zeta,Z_j)$ there exists $z \in Z_j$ with $s =
|\zeta-z|$ and one infers, using \cite[2.5]{snulmenn.isoperimetric} with
$n$, $m$, $U$, $\mu$, $\varepsilon$, and $\varrho$ replaced by ${m}$,
${n-m}$, $\oball{\zeta}{1/i}$, $\| V \|$, $(2 \isoperimetric{{m}}
)^{-1}$, and $s/2$, that
\begin{gather*}
s \leq |\zeta-c| \leq r < 1/i \leq 1/(2j),
\quad
\cball \zeta{s/2} \subset \cball z{3s/2} \cap \cball c{2r} \without Z_j,
\\
( 4 {m} \isoperimetric{{m}} )^{-{m}} s^{m}
\leq \measureball{\| V \|}{\cball{\zeta}{s/2}}
\leq \| V \| ( \cball{c}{2r} \without Z_j )
< \varepsilon^{m} r^{m},
\quad
3s/2 < r,
\\
\measureball{\mu}{\cball \zeta{s/2}}
\leq \measureball{\mu}{\cball z{3s/2}}
\leq j (3s/2)^{{m}} \omega((3s/2)+)
\leq \gamma \omega (r) \measureball{\|V\|}{\cball \zeta{s/2}},
\end{gather*}
where $\gamma = j ( 6 {m} \isoperimetric{{m} } )^{{m}}$. Therefore the
Besicovitch-Federer covering theorem yields the existence of~countable
disjointed families $F_1, \ldots, F_{\besicovitch{{n}}}$ of closed balls
such that
\begin{gather*}
\cball cr \cap ( \spt \|V\| ) \without ( G_i \cup Z_j )
\subset {\textstyle\bigcup\bigcup \{ F_k \with k = 1, \ldots,
\besicovitch{{n}} \}} \subset \cball c{2r} \without Z_j, \\
\mu (S) \leq \gamma \omega(r) \| V \| (S) \quad \text{whenever
$S \in F_k$ and $k = 1, \ldots, \besicovitch{{n}}$} .
\end{gather*}
Recalling $\mu(Z) = 0$, the conclusion follows from
\begin{align*}
\mu ( \cball cr \without G_i )
& = \mu ( \cball cr \cap ( \spt \|V\| ) \without ( G_i \cup Z_j ) )
\\
& \leq \tsum{k=1}{\besicovitch{{n}}} \tsum{S \in F_k}{} \mu (S)
\leq \tsum{k=1}{\besicovitch{{n}}} \gamma \omega (r) \tsum{S \in F_k}{} \| V \| (S)
\\
& \le \besicovitch{{n} } \gamma \omega (r) \|V\| ( \cball c{2r} \without Z_j)
\leq \besicovitch{{n}} \gamma \varepsilon^{m} r^{m} \omega (r) ,
\end{align*}
since $\varepsilon$ can be chosen arbitrary small.
\end{proof}
\begin{remark}
The preceding theorem is a slight generalisation of
\cite[3.1]{snulmenn.isoperimetric} which treated the case that $\omega
(r)$ equals a positive power of $r$; see
\cite[3.2--3.4]{snulmenn.isoperimetric} for comments on earlier
developments and the sharpness of certain hypotheses. The case $q=1$ is
excluded since in this case there is no modulus of continuity satisfying
the condition on the limit inferior.
\end{remark}
\begin{corollary}
\label{thm:differentiation}
Suppose ${m}$, ${n}$, $p$, $U$, and $V$ are as in
\ref{miniremark:situation_general}, $p={m}$, $g$ maps $\| V \|$ almost all
of $U$ into $\{ t \with 0 \leq t \leq \infty \}$, $\dmn g \subset \spt
\|V\|$, and $0 < q < \infty$.
Then, for $\| V \|$ almost all $z$, either
\begin{enumerate}
\item $\limsup_{\zeta \to z} |\zeta-z|^{-q} g(\zeta) = \infty$, or
\item $g(z) = 0$ and $\lim_{\zeta \to z} |\zeta-z|^{-q} g(\zeta) =
0$.\footnote{Our usage of limits is affected by the requirement
\begin{gather*}
\limsup_{z \to c} g( z) = \lim_{\varepsilon \to 0+} \sup \{
g(z) \with z \in \cball c\varepsilon \cap \dmn g \},
\end{gather*}
in particular $\limsup_{z \to c} g ( z ) = - \infty$ if $c \notin
\Clos \dmn g$.}
\end{enumerate}
\end{corollary}
\begin{proof}
Define $C = \{ z \with g(z) = 0 \}$ and $Z = U \cap \{ \limsup_{\zeta \to
z} |\zeta-z|^{-q} g(\zeta) < \infty \}$. Then $Z$ is a Borel subset of $U$
and $Z \cap ( \dmn g ) \without C$ is countable; in fact, $Z$ is the union
of the relatively closed subsets of $U$ defined by
\begin{gather*}
Z_i = U \cap \{ z \with \text{$g( \zeta ) \leq i | \zeta-z |^q$
whenever $\zeta \in \dmn g$ and $0 < | \zeta-z | < 1/i$} \}
\end{gather*}
for $i \in \mathscr{P}$ and $Z_i \cap ( \dmn g ) \without C$ is contained in the
set of isolated points of $Z_i$. Therefore $\| V \| ( Z \without C ) = 0$
and $C \cap Z$ is $\| V \|$ measurable. Hence, applying \ref{thm:O_o} with
$\omega (r)$, $\mu (B)$, and $Z$ replaced by $r$, $( \sup ( \{ 0 \} \cup g
[ B ]))^{({m}+1)/q}$, and $C \cap Z$, one derives
\begin{gather*}
\limsup_{\zeta \to z} | \zeta-z |^{-q} g ( \zeta ) = 0 \quad \text{for
$\| V \|$ almost all $z \in Z$}.
\end{gather*}
As $\dmn g$ does not contain isolated points, the conclusion follows.
\end{proof}
\begin{remark}
Clearly, the special case \cite[3.1]{snulmenn.isoperimetric} of
\ref{thm:O_o} would also be sufficient for the proof of the corollary.
\end{remark}
\section{Embedding results}
\label{sec:interpolation}
In the present section we formulate for convenient reference two embedding
results for Sobolev functions in Euclidean space which measure the lower order
term only on a set of suitably large Lebesgue measure.
\begin{lemma} \label{lemma:int-ineq}
Suppose $2 \leq {m} \in \mathscr{P}$, $a \in \mathbf{R}^{m}$, $0 < r < \infty$,
$0 < \varepsilon \le (\unitmeasure{{m}}/2)^{1/{m}}$, $A$ is an
$\mathscr{L}^{m}$ measurable subset of $\oball ar$, $\mathscr{L}^{m}(\oball ar
\without A) \le (\varepsilon r)^{m}$, and $f \in \Sob{}{1}{1}(\oball
ar)$.
Then there holds
\begin{multline*}
r^{-1} \eqLpnorm{\mathscr{L}^{m} \mathop{\llcorner} \oball
ar}{{m}}{f} \\
\leq \Gamma \big ( \varepsilon^{1/2}
\eqLpnorm{\mathscr{L}^{m} \mathop{\llcorner} \oball ar}{{m}}{\operatorname{\mathbf{D}}
f} + \varepsilon^{-1/2} r^{-1} \eqLpnorm{\mathscr{L}^{m}
\mathop{\llcorner} A}{{m}}{f} \big ),
\end{multline*}
where $\Gamma$ is a positive, finite number depending only on ${m}$.
\end{lemma}
\begin{proof}
By H{\"o}lder's inequality it is sufficient to prove the statement
that results from replacing $r^{-1} \eqLpnorm{\mathscr{L}^{m}
\mathop{\llcorner} \oball ar}{{m}}{f}$ by $r^{-1/2}
\eqLpnorm{\mathscr{L}^{m} \mathop{\llcorner} \oball ar}{2{m}}{f}$. The
latter is a special case of \cite[6.3]{snulmenn.decay} taking $\zeta =
\frac 23 {m}$, $\xi = {m}$, $s = {m}$ and $\lambda =
(\varepsilon r)^{m}$.
\end{proof}
\begin{theorem}
\label{corollary:orlicz-poincare}
Suppose $2 \leq {m} \in \mathscr{P}$, $\Phi$ is related to ${m}$ as in
\ref{miniremark:Phi}, $a \in \mathbf{R}^{m}$, $0 < r < \infty$, $A$ is an
$\mathscr{L}^{m}$ measurable subset of $\oball ar$ with $\mathscr{L}^{m}(A) \ge
\frac 12 \unitmeasure{{m}} r^{m}$, and $f \in \Sob{}{1}{1}(\oball
ar)$,
Then there holds
\begin{gather*}
\eqLpnorm{\mathscr{L}^{m} \mathop{\llcorner} \oball ar}{r^{-m}\Phi}{f} \le
\Gamma \big ( \eqLpnorm{\mathscr{L}^{m} \mathop{\llcorner} \oball
ar}{{m}}{\operatorname{\mathbf{D}} f} + r^{-1} \eqLpnorm{\mathscr{L}^{m}
\mathop{\llcorner} A}{{m}}{f} \big ),
\end{gather*}
where $\Gamma$ is a positive, finite number depending only on ${m}$.
\end{theorem}
\begin{proof}
The problem may be reduced firstly to the case $A = \oball ar$ by
\ref{lemma:int-ineq} and secondly to the case $a = 0$ and $r = 1$ using
translations and homotheties. The remaining case is a special case of
\cite[8.27]{MR2424078}.
\end{proof}
\subsection*{Outline of the proofs}
\subsubsection*{Theorem~A}
In order to explain the proof of Theorem~A, it is instructive to recall the
strategy of proof of similar results in \cite[5.2]{snulmenn.c2}. The proof rests
on an idea of Sch{\"a}tzle underlying \cite[Theorem 3.1]{MR2064971}. He realised
that in the presence of a coercive estimate, see Brakke \cite[5.5]{MR485012},
second order rectifiability of a varifold satisfying the general hypotheses with
$p = 2$ implies second order behaviour of the quadratic tilt-excess and the
quadratic height-excess. Having second order rectifiability at one's disposal
from \cite[4.8]{snulmenn.c2}, this procedure has been employed in
\cite[5.2\,(2)]{snulmenn.c2} with a refined coercive estimate. Specifically,
Brakke's estimate \cite[5.5]{MR485012} was improved by introducing height-excess
quantities measured in $\Lp q ( \| V \| )$ with $q = \frac{2 {m}}{{m}-2}$ if
${m} > 2$, see \cite[4.10, 4.11]{snulmenn.decay}. Employing an approximation
by $\mathbf{Q}_Q ( \mathbf{R}^{n-m})$ valued functions and interpolation in \cite[5.7,
6.4]{snulmenn.decay} yielded~a coercive estimate involving approximate
height-excess quantities, that is height-excess that excludes an arbitrary set
of small $\| V \|$ measure, see \cite[9.5]{snulmenn.decay}. Finally, in all
cases in which this method implied a positive answer to Question~3 for $q=2$ and
$\alpha < 1$, the differentiation theory from \cite[3.7]{snulmenn.isoperimetric}
was employed in conjunction again with the second order rectifiability to
establish that the limit in question is actually equal zero $\| V \|$ almost
everywhere.
The results obtained by the above method in case ${m} =2$ were not sharp as
the coercive estimate \cite[4.10]{snulmenn.decay} did only use Lebesgue spaces
so that subsequently the non-sharp embedding of weakly differentiable functions
with square-integrable weak derivative on $\oball ar$, for $a \in \mathbf{R}^2$ and $0
< r < \infty$, into $\Lp q ( \mathscr{L}^2 \mathop{\llcorner} \oball ar)$ for $q <
\infty$ needed to be employed. To be able to employ the sharp embedding into
Orlicz spaces, see \cite[8.27, 8.28]{MR2424078}, we therefore modify the above
procedure. In~particular, we obtain a coercive estimate involving the
appropriate Orlicz space in \ref{lemma:coercive-estimate} which takes the role
of \cite[4.10]{snulmenn.decay}. Moreover, to be able to proceed after obtaining
a weaker form of Theorem~A which results from replacing ``$\lim$'' and ``$=0$''
by ``$\limsup$'' and ``$<\infty$'' in its statement, the differentiation result
\cite[3.7]{snulmenn.isoperimetric} is adapted in \ref{thm:O_o} to include rates
which are not powers. It should also be noted that as the current proof does not
aim at decay \emph{estimates}, i.e.,~estimates for positive radii, such as
\cite[10.2]{snulmenn.decay} but only at decay \emph{rates}, i.e.,~the behaviour
as the radius approaches zero, we are able to encapsulate the usage of the
approximation by $\mathbf{Q}_Q ( \mathbf{R}^{n-m} )$ valued functions in the
construction of a real valued auxiliary function in
\ref{lemma:mini-lip-approx}. In comparison to \cite[9.5]{snulmenn.decay} where
the coercive estimate involving the approximate height-excess quantities was
derived as corollary to the more elaborate estimates in
\cite[9.4\,(9)]{snulmenn.decay} aiming at the proof of
\cite[10.2]{snulmenn.decay}, this greatly simplifies our derivation of the
corresponding estimates in \ref{theorem:quadratic-tilt-decay}.
\subsubsection*{Theorems B and C}
The qualitative construction principle for both theorems is that of Brakke
\cite[6.1]{MR485012}. The basis for obtaining quantitative information is
provided by the following variant of classical propositions, see
\ref{example:yet_another_cantor_set}.
\emph{If ${m}$ is a positive integer, $0 < \lambda < 1$, and $\omega$ is a
modulus of continuity, then there exist a countable disjointed family $G$ of
cubes contained in the unit cube $C$ of $\mathbf{R}^{m}$, $B \subset \mathbf{R} \cap \{
r \with r > 0 \}$ with $\inf B = 0$, and $\varepsilon > 0$ such that
$\mathscr{L}^{m}$ almost all $a \in C \without \bigcup G$ have the following
property: If $0 < r \leq \varepsilon$, then there exists a subfamily $H$ of
$G$ with $\bigcup H \subset \cball ar$,
\begin{gather*}
\mathscr{L}^{m} ( \textstyle{\bigcup H} ) \geq \omega (r) r^{m}
\quad \text{and} \quad
\text{if $r \in B$ then $\card H = 1$.}
\end{gather*}} Moreover, $B$ may be required to equal $\{ r \with 0 < r \leq
\varepsilon \}$ if and only if $\omega$ satisfies the Dini condition, see
\ref{example:cantor_set} and \ref{remark:topsoe}. Denoting by $F$ the
collection of balls with centres and ``radii'' equal to those of the cubes in
$G$ and fixing an isometric injection of $\mathbf{R}^{m}$ into $\mathbf{R}^{{m}+1}$, we
associate to each member of $F$ an ${m}$~dimensional submanifold
of~$\mathbf{R}^{{m}+1}$ of class~$1$ involving a~piece of a~catenoid, see
Figures~\ref{F:catenoid} and~\ref{F:bent-catenoid} on page~\pageref{F:catenoid}
and~\pageref{F:bent-catenoid}. The varifolds whose existence is asserted in
Theorems~B and~C then consist of the submanifolds corresponding to the members
of $F$ together with the image of $\mathbf{R}^{m} \without \bigcup F$ under the
injection [taken with multiplicity two in case of Theorem~C].
\subsubsection*{Theorem D} The key to the proof of Theorem~D is an a priori
estimate for ``weak subsolutions to Poisson's equation for the
Laplace-Beltrami operator on $V$'', see \ref{lemma:sup-L1-est} and
\ref{remark:laplace-beltrami}. This estimate is adapted from Allard and
Almgren, see~\cite[5\,(6)]{MR0425741}, and implies in particular a positive
answer to Question~1 if ${m} = 1$ and $q = \infty$, see
\ref{remark:sup-L1-est}. Since a positive answer to Question~3 for ${m} = 1$
and $q = 2$ is already known from \cite[5.2\,(2)]{snulmenn.c2}, part
\eqref{item:1d:diff} is now a consequence of suitable differentiation result,
see \ref{thm:differentiation}, which in turn is based on
\cite[3.1]{snulmenn.isoperimetric}. Part \eqref{item:1d:representation} then
follows by a suitable approximation by $\mathbf{Q}_Q ( \mathbf{R}^{{n}-1} )$ valued
functions, see \cite[3.15]{snulmenn.poincare}, and a selection theorem for
such function, see Almgren \cite[1.10]{MR1777737}.
\subsubsection*{Curvature varifolds} Our treatment of curvature varifolds
makes use of generalised weakly differentiable functions introduced in
\cite[8.3]{snulmenn:tv.v2}, in particular their differentiability properties,
see \cite[11.2, 11.4]{snulmenn:tv.v2}. This approach rests on the
characterisation of curvature varifolds in terms of such functions,
see~\cite[15.6]{snulmenn:tv.v2}. Otherwise, only some more elementary facts
are cited from~\cite{snulmenn:tv.v2}.
\subsection*{An open problem}
The cases of Question~3 remaining open are related to the following question
concerning the decay behaviour of tilt-excess quantities measures in the Lorentz
space with exponent $(2,\infty)$, see~\ref{remark:m=2,p=1}.
\begin{citing} [Question~4]
Suppose ${m} = 2$ and $p = 1$.
Do the general hypotheses imply that
\begin{gather*}
\limsup_{r \to 0+} r^{-1} E(c,r) < \infty
\quad
\text{for $\| V \|$ almost all $c$},
\end{gather*}
where $E(c,r) = \sup \{ t r^{-{m}/2} \| V \| ( \cball cr \cap \{ z \with |
\tau (z)-\tau(c) | > t \} )^{1/2} \with 0 < t < \infty \}$?
\end{citing}
If the answer should turn out to be in the affirmative, it would imply Theorem~A
as a corollary as well as a positive answer to Question~3 in the remaining cases
$\alpha = 2/q$ for $2 < q < \infty$ both by interpolation based on the
boundedness of $\tau$. For the model case of the Laplace operator the procedure
used to prove Theorem~A could be adapted since coercive estimates in Lorentz
spaces are available for the Laplace operator; such estimates are reviewed for
instance in Dolzmann and M{\"u}ller~\cite[\S 3]{MR1354111}.
\subsection*{Organisation of the paper}
In Section \ref{sec:notation} the notation is introduced. Sections
\ref{sec:lebesgue}--\ref{sec:differentiation} are of preparatory nature
supplying the density properties of the Lebesgue measure, properties of
Cartesian products of varifolds, and some differentiation theory for functions
on varifolds respectively. Sections
\ref{sec:qte-ex}--\ref{sec:quadratic_tilt_decay} treat the quadratic tilt-excess
whereas Sections \ref{sec:super_quadratic_tilt}--\ref{sec:sqte-decay} cover the
super-quadratic case. This includes, in Section \ref{sec:super_quadratic_tilt},
the example concerning the integrability of the second fundamental form of
curvature varifolds. Finally, one dimensional varifolds are considered in
Section \ref{sec:one_dimensional_decay}.
\subsection*{Acknowledgement}
The first author was partially supported by the Foundation for Polish Science
and partially by NCN Grant no. 2013/10/M/ST1/00416 while he was on leave from
Institute of Mathematics of the University of Warsaw.
It is a~pleasure to the authors to express their gratitude to Prof.~Guy David
for his valuable input to this paper and to thank Prof.~David Preiss for
informing them about the studies of Mejlbro and Tops{\o}e in \cite{MR0583557}
and Tops{\o}e in \cite{MR577961}.
\emph{AEI publication number:} AEI-2015-007
\subsection*{Known results}
In order to more formally describe the results of the present paper in the
context of known results, consider the following set of hypotheses; the notation
is explained in Section \ref{sec:notation}. Additionally, the terms ``[twice]
weakly differentiable'' are employed with their usual meaning, see
e.g.~\cite[p.~9--10]{snulmenn.decay}.
\begin{citing} [General hypotheses]
Suppose ${m}$ and ${n}$ are positive integers, ${m} < {n}$, $1
\leq p \leq \infty$, $U$ is an open subset of $\mathbf{R}^{n}$, $V$ is an
${m}$ dimensional integral varifold in $U$ whose first variation $\delta
V$ is representable by integration\footnote{That is, in the terminology of
Simon \cite[39.2]{MR756417}, $V$ is of locally bounded first variation.},
the generalised mean curvature vector $\mathbf{h} (V,\cdot)$ of $V$
belongs to $\Lploc{p} (\| V \|, \mathbf{R}^{n} )$, and if $p > 1$ then
\begin{gather*}
( \delta V ) ( \theta ) = - \tint{}{} \mathbf{h} (V,z) \bullet \theta
(z) \ensuremath{\,\mathrm{d}} \| V \| z \quad \text{for $\theta \in \mathscr{D} ( U,
\mathbf{R}^{n} )$}.
\end{gather*}
Let $Z = U \cap \{ z \with \Tan^{m} ( \| V \|, z ) \in \grass {n}
{m} \}$ and define $\tau : Z \to \Hom ( \mathbf{R}^{n}, \mathbf{R}^{n} )$ by
$\tau (z) = \project{\Tan^{m} ( \| V \|, z)}$ for $z \in Z$.
\end{citing}
The set $Z$ consists of all points such that the closed cone of $(\| V \|,
{m} )$ approximate tangent vectors forms an ${m}$ dimensional plane; for
these points $z$, $\tau (z)$ denotes the orthogonal projection of $\mathbf{R}^{n}$
onto this plane.
The study of regularity properties is usually preceded by a more basic study
of the density ratio by means of the monotonicity identity, see Allard
\cite[\S 5]{MR0307015}, Simon \cite[\S 17]{MR756417} and \cite[\S
4]{snulmenn:tv.v2}, and its consequence, the isoperimetric inequality, see
Allard \cite[\S 7]{MR0307015} and Michael and Simon \cite[\S 2]{MR0344978}. In
particular, if $p > {m}$ then $\boldsymbol{\Theta}^{m} ( \| V \|, \cdot )$ is an
upper semicontinuous real valued function whose domain is $U$, see Simon
\cite[17.8]{MR756417}. To which extent these properties persist if $p = {m}$
is unclear; the cases ${m} = 1$ and ${m} = 2$ are treated in
\cite[4.8]{snulmenn:tv.v2} and Kuwert and Sch{\"a}tzle
\cite[Appendix]{MR2119722} respectively. From Allard \cite[3.5\,(1),
8.3]{MR0307015} one is at least assured that $\mathscr{H}^{m} \mathop{\llcorner} \spt
\| V \| \leq \| V \|$ if $p = {m}$. If $p < {m}$, one easily constructs
examples with $\spt \| V \| = U$, see \cite[14.1]{snulmenn:tv.v2}. Yet, there
are precise local estimates available on the size of the set of points where
the density ratio is small, see \cite[\S 2]{snulmenn.isoperimetric}.
In order to put the study of regularity properties into perspective, it is
instructive to consider as well the behaviour of the Laplace operator.
\begin{citing} [Model case]
Suppose ${m}$ and ${n}$ are positive integers, ${m} < {n}$, $1
\leq p \leq \infty$, $u \in \Lploc 1 ( \mathscr{L}^{m}, \mathbf{R}^{n-m} )$,
the distributional Laplacian $T \in \mathscr{D}' ( \mathbf{R}^{m}, \mathbf{R}^{n-m}
)$ of $u$, defined by $T ( \phi ) = \tint{}{} u \bullet \Lap \phi \ensuremath{\,\mathrm{d}}
\mathscr{L}^{m}$ for $\phi \in \mathscr{D} ( \mathbf{R}^{m}, \mathbf{R}^{n-m}
)$, is representable by integration, and if $p > 1$ then there exists $f
\in \Lploc{p} ( \| V \|, \mathbf{R}^{n-m} )$ satisfying
\begin{gather*}
T ( \phi ) = \tint{}{} f(x) \bullet \phi(x) \ensuremath{\,\mathrm{d}} \mathscr{L}^{m}
x \quad \text{for $\phi \in \mathscr{D} ( U, \mathbf{R}^{n-m} )$}.
\end{gather*}
\end{citing}
If $1 < p < \infty$ then $u$ is twice weakly differentiable and the
distributional partial derivatives up to second order of $u$ correspond to
functions in $\Lploc p ( \mathscr{L}^{m}, \mathbf{R}^{n-m} )$; this is
a~consequence of the usual a priori estimate based on the Calder{\'o}n Zygmund
inequality, see e.g.~\cite[3.5]{snulmenn.c2}, and convolution.
This implies differentiability results in Lebesgue spaces for the weak
derivative, see for instance Calder{\'o}n and Zygmund \cite[Theorem~12,
p.~204]{MR0136849}.
For an integral varifold $V$, the existence of a notion of first order
derivative, i.e.,~that $\tau$ is defined $\| V \|$ almost everywhere, is a
simple consequence of its rectifiability, see Allard
\cite[3.5\,(1)]{MR0307015}. However, this derivative behaves rather like an
approximate derivative than a weak derivative as is exemplified by the fact
that a Poincar{\'e} inequality only holds under additional hypotheses on the
first variation, see \cite[p.~372]{snulmenn.poincare}. More information on the
validity of Sobolev Poincar{\'e} type inequalities may be retrieved from
Hutchinson \cite{MR1066398}, \cite[\S4]{snulmenn.poincare} and
\cite[\S10]{snulmenn:tv.v2}.
As the Grassmann manifold is compact, $\tau$ belongs to $\Lploc q ( \| V \|,
\Hom ( \mathbf{R}^{n}, \mathbf{R}^{n} ) )$ for $1 \leq q \leq \infty$. Yet, it is
important to understand for which $q$ effective coercive estimates are
available. Classically, this is the case for $q = 2$, see Allard
\cite[8.13]{MR0307015} and its refinements Brakke \cite[5.5]{MR485012} and
\cite[4.10, 4.14]{snulmenn.decay}.
\begin{citing} [Question 1]
Suppose $2 < q < \infty$ and $1 < p < \infty$.
Do the general hypotheses imply that for $\|V\|$ almost all $c$ there
exists $1 \leq \gamma < \infty$ such that there holds
\begin{multline*}
\big ( r^{-{m}} \tint{\cball c{r/2}}{} | \tau(z)-\tau(c) |^q \ensuremath{\,\mathrm{d}} \|
V \| z \big )^{1/q} \\
\leq \gamma \Big ( \big ( r^{-{m}} \tint{\cball cr}{} |\tau
(z)-\tau(c)|^2 \ensuremath{\,\mathrm{d}} \| V \| z \big )^{1/2} + \big ( r^{p-{m}}
\tint{\cball cr}{} | \mathbf{h} (V,z)|^p \ensuremath{\,\mathrm{d}} \| V \| z \big )^{1/p}
\Big )
\end{multline*}
whenever $0 < r \leq \gamma^{-1}$? A similar question may be phrased for
$q = \infty$ or $p \in \{ 1, \infty \}$.
\end{citing}
In case ${m} \geq 2$ and $q = \infty$ such estimates are known to fail by
Brakke \cite[6.1]{MR485012}. In view of \ref{remark:holes-tilt-large} the
Question 1 is related to the possible size of the exceptional sets occurring
in approximations by $\mathbf{Q}_Q ( \mathbf{R}^{n-m} )$ valued functions. It is also
related to the second ``main analytic estimate'' -- so termed in Almgren's
announcement \cite[p.~6]{MR574247} -- of Almgren's regularity proof for area
minimising currents of arbitrary codimension, see Almgren \cite[3.29,
3.30]{MR1777737} and De~Lellis and Spadaro \cite[Theorem~7.1]{MR3283929}.
Passing from first order to second order quantities, the analogous property to
weak differentiability of the weak derivative of $u$ would be generalised $V$
weak differentiability of~$\tau$, or equivalently, $V$ being a curvature
varifold in the sense of Hutchinson, see \ref{def:curvature_varifold} and
\ref{remark:curvature_varifold}. Considering three half lines emanating from
the origin in $\mathbf{R}^2$ at equal angles shows that even a stationary integral
varifold need not to be a curvature varifold, see Mantegazza \cite[3.4,
3.11]{MR1412686}. In view of \cite[4.8]{snulmenn.c2}, one may however still
define a notion of approximate second fundamental form, $\ap \mathbf{b}
(V,\cdot)$, of a varifold satisfying the general hypotheses such that
\begin{gather*}
\ap \mathbf{b} (V,z) = \mathbf{b} (M,z) \quad \text{for $\| V \|$ almost
all $z \in U \cap M$}
\end{gather*}
whenever $M$ is an ${m}$ dimensional submanifold of $\mathbf{R}^{n}$ of class
$2$. Since the corresponding approximate mean curvature vector is $\|V \|$
almost equal to $\mathbf{h} (V,\cdot)$ by \cite[4.8]{snulmenn.c2}, we assume
${m} \geq 2$ in the following question.
\begin{citing} [Question~2]
Suppose ${m} \geq 2$, $p = \infty$, and $0 < q < \infty$.
Do the general hypotheses imply that for $\| V \|$ almost all $c$
\begin{gather*}
\tint{\cball cr}{} \| \ap \mathbf{b} (V,z) \|^q \ensuremath{\,\mathrm{d}} \| V \| z < \infty
\quad \text{for some $r > 0$}?
\end{gather*}
\end{citing}
The existence proof in \cite[4.8]{snulmenn.c2} does not provide any integral
estimates of the resulting quantity and, in fact, no positive results are
known. Considering the scaling behaviour of the above integral, the example in
Brakke~\cite[6.1]{MR485012} shows that the answer is in the negative whenever $q
\geq 2$.
In case $V$ happens to be a curvature varifold, one may deduce
differentiability results for $\tau$ in Lebesgue spaces from general facts
about generalised $V$ weakly differentiable functions, see \cite[11.4,
15.9--15.12]{snulmenn:tv.v2}. The next question concerns to which extent these
properties persist for non-curvature varifolds.
\begin{citing} [Question 3]
Suppose $0 < \alpha \leq 1$ and $2 \leq q < \infty$.
Do the general hypotheses imply that
\begin{gather*}
\limsup_{r \to 0+} r^{-\alpha} \big ( r^{-{m}} \tint{\cball cr}{} |
\tau (z)- \tau (c) |^q \ensuremath{\,\mathrm{d}} \| V \| z \big )^{1/q} < \infty
\end{gather*}
for $\| V \|$ almost all $c$? A similar question may be phrased for $q =
\infty$.
\end{citing}
This may be seen as a pointwise H{\"o}lder condition with exponent $\alpha$ on
$\tau$ at~$c$ measured in a Lebesgue space with exponent $q$. If $p > {m}$
and $p \geq 2$, proving uniform estimates for $$r^{-\alpha} \big ( r^{-{m}}
\tint{\cball cr}{} | \tau (z) - \tau (c) |^2 \ensuremath{\,\mathrm{d}} \| V \| z \big )^{1/2}$$ for
all $c$ in some relatively open subset of $\spt \| V \|$ and all $0 <r \leq
\varepsilon$ for some $\varepsilon > 0$ is -- via H{\"o}lder continuity of
$\tau$ -- the key to Allard's regularity theorem, see \cite[\S 8]{MR0307015}.
If ${m} \geq 2$ and $p = \infty$, then $\tau$ may be discontinuous at points
in a set of positive $\| V \|$ measure and $\spt \| V \|$ may fail to be
associated to an $\mathbf{R}^{n-m}$ valued or even $\mathbf{Q}_Q ( \mathbf{R}^{n-m} )$
valued function near those points by Brakke \cite[6.1]{MR485012}. Therefore,
the condition in Question~3 also acts as replacement for more classical
notions of regularity which are known to possibly fail on a set of positive
$\| V \|$ measure.
Evidently, if ${m} \geq 2$ and $q = \infty$ the answer is in the negative
whenever $1 \leq p \leq \infty$ by Brakke \cite[6.1]{MR485012}. If $p < {m}$,
the answer is in the negative whenever $\alpha q > {m} p/({m}-p)$,
see~\cite[1.2\,(iv)]{snulmenn.isoperimetric}.
Turning to positive results, only the case $q=2$ has been systematically
studied. The sharpest known results were obtained in \cite{snulmenn.c2} building
partly on \cite{snulmenn.decay} and extending methods and results of Brakke
\cite[5.5, 5.7]{MR485012} and Sch\"atzle \cite[5.4]{MR1906780},
\cite[Theorem~5.1]{MR2064971}, and \cite[Theorem~3.1]{MR2472179}. Namely, if
${m} = 1$ or ${m} = 2$ and $p > 1$ or ${m} > 2$ and $p \geq 2
{m}/({m}+2)$, then the answer is in the affirmative for $\alpha = 1$ and $q
= 2$, see \cite[5.2\,(2)]{snulmenn.c2}, and if ${m} =2$ and $\alpha < 1$ or
$\sup \{ 2,p \} < {m}$ and $\alpha = {m} p/(2({m}-p)) < 1$ then the answer
is in the affirmative for $q = 2$, see \cite[5.2\,(1)]{snulmenn.c2}.
Therefore, the previously known results for Question 3 may be summarised as
follows. If $q = 2$, only the case $({m},p,\alpha) = (2,1,1)$ was left open.
If $2 < q < \infty$, the gap between positive results and known counterexamples
was quite large. And if $q = \infty$, only the case ${m} = 1$ remained
open. The initial motivation for the present work was to solve some of these
open cases of Question 3.
Finally, notice that Question~3 could be phrased for $1 \leq q < 2$ as well,
see \cite[p.~248, Problem (ii)]{snulmenn.isoperimetric}, and Question~2 could
include the case $p < \infty$. However, no effort has been made to resolve
these additional cases in the present study.
\subsection*{Results of the present article}
The results may be summarised as follows. All cases of Question~1 are answered;
in the negative if ${m} \geq 2$ and in the affirmative if ${m} = 1$, see
\ref{remark:no-gehring-improvement} and \ref{remark:sup-L1-est}. Question~2 is
answered in the negative if $q > 1$, see
\ref{remark:example-super-quadratic-tilt}. All cases of Question~3 except the
case $({m},p,\alpha) = (2,1,2/q)$ for $2 < q < \infty$ are treated, see
\ref{thm:positive_result},
\ref{remark:sharpness_superquadratric}--\ref{remark:m=2,p=1}, and
\ref{thm:decay-1d}. Finally, for the case $({m},p,q) = (2,1,2)$ of Question~3,
the precise decay rate is determined, see \ref{example:quadratic_tilt_excess}
and \ref{theorem:quadratic-tilt-decay}.
Beginning with the last item, the following theorem is established.
\begin{citing} [Theorem~A, see \ref{theorem:quadratic-tilt-decay}]
Suppose ${m} = 2$ and $p = 1$.
If ${n}$, $U$, $V$, and $\tau$ satisfy the conditions of the general
hypotheses, then
\begin{gather*}
\lim_{r \to 0+} r^{-4} ( \log (1/r) )^{-1} \tint{ \cball cr}{} | \tau
(z) - \tau (c) |^2 \ensuremath{\,\mathrm{d}} \| V \| z = 0
\end{gather*}
for $\| V \|$ almost all $c$.
\end{citing}
This result is sharply complemented by the following negative result.
\begin{citing} [Theorem~B, see \ref{example:quadratic_tilt_excess}]
Suppose ${m} = 2$, ${n} = 3$, $p = 1$, $U = \mathbf{R}^3$, and $\omega$ is a
modulus of continuity.
Then there exist $V$ and $\tau$ satisfying the conditions of the general
hypotheses and $C$ with $\| V \| (C) > 0$ satisfying
\begin{gather*}
\limsup_{r \to 0+} \omega (r)^{-1} r^{-4} ( \log (1/r) )^{-1} \tint{
\cball cr}{} | \tau (z) - \tau (c) |^2 \ensuremath{\,\mathrm{d}} \| V \| z > 0
\end{gather*}
whenever $c \in C$.
\end{citing}
In fact, the varifold constructed in Theorem~B may be chosen to be associated to
the graph of a Lipschitzian function with small Lipschitz constant. Theorem~B in
particular answers the case $({m},p,\alpha) = (2,1,1)$ of the part $q = 2$
of Question 3 in the negative.
\begin{citing} [Theorem~C, see \ref{example:quantitative_brakke_again} and
\ref{remark:second_fundamental_form_large}]
Suppose ${m} \geq 2$, ${n} = {m}+1$, $p = \infty$, $U =
\mathbf{R}^{n}$, and $\omega$ is a modulus of continuity.
Then there exist a curvature varifold $V \in \IVar_{m} ( \mathbf{R}^{n} )$ and
$\tau$ satisfying the conditions of the general hypotheses, an ${m}$
dimensional submanifold $M$ of $\mathbf{R}^{n}$ of class $\infty$ which is
relatively open in $\spt \| V \|$, $\varepsilon > 0$, $B \subset \{ r \with
r > 0 \}$ with $\inf B = 0$, and $C$ with $\| V \| (C)>0$ such that the
following properties hold whenever $c \in C$.
\begin{enumerate}
\item \label{item:brakke:tilt} If $r \in B$, then $$\| V \| ( \cball
cr \cap \{ z \with \| \tau(z)-\tau(c)\| \geq 1/3 \} ) \geq \omega (r)
r^{{m}+2} ( \log (1/r) )^{-2}.$$
\item \label{item:brakke:holes} If $r \in B$ and $T = \im \tau (c)$,
then $$\mathscr{H}^{m} ( H(T,c,r) ) \geq
\omega(r) r^{{m}+2} ( \log (1/r))^{-2},$$ where $H (T,c,r) = T \cap
\cball{\project{T}(c)}{r} \without \project{T} [ \cylinder Tcrr
\cap \spt \| V \| ]$.
\item \label{item:brakke:2nd_fundamental_form} If $r > 0$ and $1 < q < \infty$,
then
$$\tint{M \cap \cball cr}{} | \mathbf{b} ( M,z) |^q \ensuremath{\,\mathrm{d}}
\mathscr{H}^{m} z = \infty.$$
\item \label{item:brakke:density} If $0 < r \leq \varepsilon$, then
$$\| V \| ( \cball cr \cap \{ z \with \boldsymbol{\Theta}^{m} ( \| V \| , z )
\leq \boldsymbol{\Theta}^{m} ( \| V \|, c) - 1 \} ) \geq \omega (r) r^{m}.$$
\end{enumerate}
\end{citing}
If $\omega$ satisfies the Dini condition and ``$\omega(r)$'' is replaced by
``$\omega(r)^2$'' in parts \eqref{item:brakke:tilt} and
\eqref{item:brakke:holes}, then one may take $B = \{ r \with 0 < r \leq
\varepsilon \}$, see \ref{example:quantitative_brakke} and
\ref{remark:example-super-quadratic-tilt}.
Part \eqref{item:brakke:tilt} answers Question~3 in the negative whenever $q >
2$ and $\alpha > 2/q$. Notice that the $\| V \|$ measure of sets of the form
considered in part \eqref{item:brakke:tilt} occurs as upper bound on the size of
the exceptional sets of the usual approximation by $\mathbf{Q}_Q ( \mathbf{R}^{n-m})$
valued functions, \ref{remark:holes-tilt-large}. In part
\eqref{item:brakke:holes} we provide a lower bound on the size of ``holes'' of
the varifold. As these regions will always be part of one of the exceptional
sets of any approximation by $\mathbf{Q}_Q (\mathbf{R}^{n-m})$ valued functions, this
yields a corresponding lower bound on the size of this set. For any $\delta >
0$, this lower bound also rules out an upper bound on the size of the
exceptional sets by a suitable multiple of
\begin{gather*}
E^{1+\delta}, \quad \text{where $E = r^{-{m}} \tint{\cball cr}{} |
\tau (z)-\tau(c) |^2 \ensuremath{\,\mathrm{d}} \| V \| z$}.
\end{gather*}
Such upper bound would be the natural analogue for varifolds satisfying the
general hypotheses with ${m} \geq 2$ and $p =\infty$ of the aforementioned
second ``main analytic estimate'' for area minimising currents obtained by
Almgren in \cite[3.29\,(8)]{MR1777737}. Of course, our example does not
(obviously) preclude a possible bound involving $E ( \log (1/E))^{-2}$, for $E
> 0$, in place of $E^{1+\delta}$, or the original bound $E^{1+\delta}$ under
the additional hypothesis of stationarity.
Evidently, part \eqref{item:brakke:2nd_fundamental_form} provides a negative
answer to Question~2 whenever $q > 1$; in fact, even if the integral in
question is restricted to the ``regular part'' of~$V$. Part
\eqref{item:brakke:density} contains the observation that the regions of
significantly smaller density around a given point may be as large as the
approximate continuity of the density permits, see
\ref{remark:density-ap-lsc}.
If ${m} \geq 2$, part \eqref{item:brakke:tilt} of Theorem~C leaves little
room for positive answers to Question~3 for $q > 2$ except of those which
follow from the boundedness of $\tau$ and the known positive results for $q
=2$, see \ref{thm:positive_result} and
\ref{remark:sharpness_superquadratric}--\ref{remark:m=2,p=1}. In fact, only
the case $({m},p,\alpha) = (2,1,2/q)$ for $2 < q < \infty$ remains open.
This case is related to the question of availability of estimates of $\tau$ in
certain Lorentz spaces, see \ref{remark:m=2,p=1}.
The combination of positive and negative answers to Question~3 obtained so far
in particular implies that the answer to Question~1 is in the negative whenever
${m} \geq 2$, see~\ref{remark:no-gehring-improvement}.
Turning to the case ${m} = 1$, a positive answer to Question~1 follows from
\ref{remark:sup-L1-est}.
\begin{citing} [Theorem~D, see \ref{thm:decay-1d} and
\ref{corollary:1-dim-Q-graph}]
Suppose ${m} = 1$ and $p =1$.
If ${n}$, $U$, $V$, $Z$, and $\tau$ satisfy the general conditions, then
there exists a subset $A$ of $Z$ with $\| V || ( A \without Z ) = 0$ such
that the following two statements hold for $\| V \|$ almost all $z \in A$.
\begin{enumerate}
\item \label{item:1d:diff} The map $\tau | A$ is differentiable at~$z$
relative to $A$ and
\begin{gather*}
\Der ( \tau | A ) (z) = ( \| V \|, 1 ) \ap \Der \tau (z).
\end{gather*}
\item \label{item:1d:representation} If $\varepsilon > 0$ then there exist a
positive integer $Q$, $0 <r< \infty$, and $f_i : T \cap \cball{\project{T}
(z)}{r} \to T^\perp \cap \cball{ \perpproject{T} (z)}{r}$ with $\Lip f_i
\leq \varepsilon$ for $i = 1, \ldots, Q$ and
\begin{gather*}
\boldsymbol{\Theta}^1 ( \| V\|, \zeta ) = \card \big \{ i \with f_i (
\project{T} ( \zeta ) ) = \perpproject{T} (\zeta) \big \}
\end{gather*}
for $\mathscr{H}^1$ almost all $\zeta \in \cylinder Tzrr$.
\end{enumerate}
\end{citing}
Part \eqref{item:1d:diff} includes a positive answer to Question~3 for ${m}
= 1$, $\alpha = 1$, and $q = \infty$. Part \eqref{item:1d:representation}
expresses the varifold as finite sum of graphs of Lipschitzian functions.
\section*{Introduction}
\subsection*{Overview}
Integral varifolds constitute an analytically tractable model for singular
geometric objects which admit appropriate notions of tangent plane and mean
curvature vector, see Almgren \cite{Almgren:Vari}, Allard \cite{MR0307015},
and Simon \cite{MR756417}. Due to good compactness properties they also arise
naturally as weak limits of smooth submanifolds of some ambient space and may
be used to represent solutions to geometrical variational problems. Our
principal objective is the study of regularity properties entailed by
integrability conditions on the first variation of such varifolds by means of
decay rates of tilt-excess. For this purpose the classical quadratic
tilt-excess and the super-quadratic tilt-excess which arises in the study of
area minimising integral currents are employed. The three main results of
this study may be informally described as follows.
Firstly, for two-dimensional integral varifolds of locally bounded first
variation (i.e. ${m}=2$ and $p=1$ in the general hypotheses below), the
optimal decay rate almost everywhere of the quadratic tilt-excess is
established, see Theorems~A and~B for the decay rate and its sharpness
respectively. For all other values of $({m},p)$ the best decay rate amongst
\emph{powers} was determined in \cite{snulmenn.c2}. The present result not only
fills this gap but in fact sharply exhibits the best decay rate amongst
\emph{all} rates, not just powers. To obtain this precision, a coercive
estimate for the quadratic tilt-excess is derived which involves a height-excess
quantity measured in an Orlicz space occurring naturally in sharp embeddings of
Sobolev spaces. This seems to be the first time that a regularity estimate for
varifolds relies on Orlicz spaces; in fact, the only previous usage of Orlicz
spaces in the context of varifolds appears to be the Poincar{\'e} type embedding
results of Hutchinson, see \cite[Theorems 2 and 4]{MR1066398}.
Secondly, for at least two-dimensional integral varifolds with locally bounded
mean curvature and ``no boundary'' ($p = \infty$), negative results concerning
decay rates almost everywhere of super-quadratic tilt-excess are shown in
Theorem~C. The importance of these examples stems from the fact that they imply
that -- even just almost everywhere and in co-dimension one -- there is no
analogue in the present situation of the ``main analytic estimate'' in Almgren's
proof of interior almost everywhere regularity of area minimising integral
currents in arbitrary codimension, see \cite{MR574247} and
\cite[\S3]{MR1777737}. (Of course, Almgren in fact proved a stronger Hausdorff
dimension estimate on the interior singular set.) This provides a serious
obstacle to an, otherwise canonical, approach to the question of possible almost
everywhere regularity of stationary integral varifolds which is a key open
problem in varifold theory.
Thirdly, for one-dimensional integral varifolds of locally bounded first
variation ($p=1$), almost everywhere differentiability of the tangent plane
map (restricted to the set of points of approximate continuity) is proven, see
Theorem~D. This implies in particular that such varifolds near almost every
point are representable as graphs of a finite number of Lipschitzian functions
with small constant. These results as well as the estimates involved in
deriving them should prove useful in the study which parts of the structural
description of one-dimensional stationary varifolds with a uniform lower bound
on their density, see Allard and Almgren in \cite{MR0425741}, generalise to
locally bounded first variation.
In combination with previous results the preceding three main results
in particular yield a nearly complete picture concerning power decay rates
almost everywhere of quadratic and super-quadratic tilt-excess.
\section{Density properties of Lebesgue measure} \label{sec:lebesgue} The
purpose of this section is to provide two examples of subsets of positive
Lebesgue measure whose complement is as large as possible near each point of
the set, see \ref{example:cantor_set} and
\ref{example:yet_another_cantor_set}. In this respect the size of the
complement is measured either by the Lebesgue measure or by the behaviour of
the distance function on the complement (equivalently by the size of cubes
contained in the complement). In the light of known positive results, the
examples obtained are sharp, see \ref{remark:topsoe} and
\ref{remark:vitali_lebesgue}. Using these examples, certain varifolds will be
constructed in Sections~\ref{sec:qte-ex} and~\ref{sec:super_quadratic_tilt} to
demonstrate that the tilt-excess decays proven in
Sections~\ref{sec:quadratic_tilt_decay} and~\ref{sec:sqte-decay} are sharp in
many cases.
Both sets are constructed by removing a suitable disjointed subcollection of
the family of all dyadic subcubes from the unit cube.
\begin{miniremark}
\label{miniremark:family_of_cubes}
Here, we collect some useful terminology for later reference.
Suppose ${m}$ is a positive integer. Define \emph{open cubes} by
\begin{gather*}
C(a,r) = \classification{\mathbf{R}^{m}}{x}{ \text{$a_j < x_j <
a_j+r$ for $j=1, \ldots, {m}$} }
\end{gather*}
whenever $a \in \mathbf{R}^{m}$ and $0 < r < \infty$. Note that $a \notin
C(a,r)$. We shall work with the following definition of dyadic cubes.
Whenever $i$ is a~nonnegative integer define $W(i)$ to consist of all open
cubes
\begin{gather*}
C ( 2^{-i} a, 2^{-i} ) \subset C(0,1)
\end{gather*}
corresponding to $a \in \mathbf{Z}^{m}$ with $0 \leq a_j \leq 2^i-1$
for $j = 1, \ldots, {m}$. Let
\begin{gather*}
Z = {\textstyle\bigcup} \{ W(i) \with i = 0, 1, 2, \ldots \},
\quad
N = {\textstyle \bigcup_{i=1}^\infty \big ( ( \Clos C(0,1) ) \without \bigcup W(i) \big )}.
\end{gather*}
Note that $W(i)$ is disjointed and $W(i) \cap W(j) = \varnothing$ if
$i \neq j$ and $\mathscr{L}^{m} (N)=0$. Observe
\begin{gather*}
\text{either $Q \subset R$ or $R \subset Q$} \quad
\text{whenever $Q,R \in Z$ and $Q \cap R \neq \varnothing$}.
\end{gather*}
\end{miniremark}
\begin{example}
\label{example:cantor_set}
Suppose ${m}$, $W$, $Z$, and $N$ are as in
\ref{miniremark:family_of_cubes}, $\omega$ is a modulus of
continuity satisfying the Dini condition, and $0 \leq \lambda < 1$.
Then there exist $0 < \varepsilon \leq 1$ and a disjointed subfamily
$G$ of $Z$ such that
\begin{gather*}
A = {\textstyle C(0,1) \without ( N \cup \bigcup G)}
\end{gather*}
satisfies the following two conditions:
\begin{enumerate}
\item \label{item:cantor_set:measure} $\mathscr{L}^{m} (A)
\geq \lambda$.
\item \label{item:cantor_set:holes} If $a \in A$ and $0 < r
\leq \varepsilon$, then there exists $Q \in G$ with
\begin{gather*}
Q \subset \oball{a}{r} \quad \text{and} \quad
\mathscr{L}^{m} (Q) \geq \omega (r) r^{m}.
\end{gather*}
\end{enumerate}
\end{example}
\begin{proof} [Construction]
Choose $0 < s \leq 2^{-1} {m}^{-1/2}$ such that $\omega ( 2
{m}^{1/2} s) \leq 2^{-2{m}} {m}^{-{m}/2}$, define $\phi :
\classification{\mathbf{R}}{t}{ 0 \leq t \leq 1 } \to \mathbf{R}$ by
\begin{gather*}
\phi ( t ) = 2^{2{m}} {m}^{{m}/2} \omega ( 2 {m}^{1/2}
t ) \quad \text{for $0 \leq t \leq s$}, \qquad \phi ( t ) = 1
\quad \text{for $s < t \leq 1$},
\end{gather*}
and note that $\phi$ is a modulus of continuity satisfying the Dini
condition. Choose a positive integer $k$ such that
\begin{gather*}
2^{-k} \leq s, \quad ( \log 2 )^{-1} \tint{0}{2^{1-k}} \phi
(t) t^{-1} \ensuremath{\,\mathrm{d}} \mathscr{L}^1 t \leq 1-\lambda.
\end{gather*}
Define a sequence $\beta_i$ of integers by the requirement
\begin{gather*}
2^{{m} (i-\beta_i+1)} > \phi ( 2^{-i} ) \geq 2^{{m} (
i-\beta_i) }.
\end{gather*}
For $k \leq i \in \mathbf{Z}$ observe that $\beta_{i+1} \geq \beta_i \geq
i$ and inductively define families $F_i$ to consist of the open cubes
$C(b,2^{-\beta_i})$, see \ref{miniremark:family_of_cubes}, corresponding
to all $C(b,2^{-i}) \in W(i)$ satisfying the following condition:
\begin{gather*}
\text{If $j$ is an integer, $k \leq j \leq i-1$, and $R \in F_j$ then
$C(b,2^{-\beta_i}) \cap R = \varnothing$}.
\end{gather*}
Let $\varepsilon = {m}^{1/2} 2^{1-k}$ and note that $\varepsilon
\leq 1$. Define
\begin{gather*}
G = {\textstyle \bigcup} \{ F_i \with k \leq i \in \mathbf{Z} \},
\quad
A = {\textstyle C(0,1) \without ( N \cup \bigcup G)}.
\end{gather*}
Notice that $F_i$ is disjointed for $k \leq i \in \mathbf{Z}$ and $G$ is a
disjointed subfamily of $Z$. Estimating
\begin{gather*}
\tsum{i=k}{\infty} 2^{{m} (i-\beta_i)} \leq
\tsum{i=k}{\infty} \phi ( 2^{-i} ) \leq \tsum{i=k}{\infty}
(\log 2)^{-1} \tint{2^{-i}}{2^{1-i}} \phi (t) t^{-1} \ensuremath{\,\mathrm{d}}
\mathscr{L}^1 t \leq 1-\lambda,
\end{gather*}
and $\card F_i \leq \card W(i) = 2^{i{m}}$ whenever $k \leq i \in
\mathbf{Z}$, one obtains
\begin{gather*}
\mathscr{L}^{m} ( {\textstyle\bigcup G} ) \leq
\tsum{i=k}{\infty} 2^{{m}(i-\beta_i)} \leq 1-\lambda, \quad
\mathscr{L}^{m} ( A ) \geq \lambda.
\end{gather*}
Suppose $a \in A$ and $0 < r \leq \varepsilon$.
There exist $i$ and $b$ such that
\begin{gather*}
k \leq i \in \mathbf{Z}, \quad 2^{-i} \leq {m}^{-1/2} r \leq
2^{1-i}, \quad {\textstyle a \in S = C (b,2^{-i}) \in W(i)}.
\end{gather*}
The proof will be concluded by showing: \emph{There exists $Q = C
(c,t) \in G$ having the property
\begin{gather*}
t \geq 2^{-\beta_i} \quad \text{and} \quad Q \cap S \neq
\varnothing
\end{gather*}
and this property implies
\begin{gather*}
Q \subset \oball{a}{r}, \quad \mathscr{L}^{m} (Q) \geq
\omega (r) r^{m}.
\end{gather*}}
Concerning the existence of $Q$, if $C(b,2^{-\beta_i}) \cap R \neq
\varnothing$ for some integer $j$ with $k \leq j \leq i-1$ and some $R \in
F_j$, then one may take $Q = R$, and otherwise one may take $Q =
C(b,2^{-\beta_i}) \in F_i$. Concerning the implication of the property,
estimate
\begin{gather*}
2^{-i} \leq s, \quad \phi ( 2^{-i} ) \geq 2^{2{m}}
{m}^{{m}/2} \omega (r), \quad t^{m} \geq 2^{-{m}
\beta_i} > \phi (2^{-i}) 2^{{m}(-i-1)} \geq \omega (r)
r^{{m}}
\end{gather*}
and note that $Q \subset S$, since $a \in S \without Q$ and $S,Q \in
Z$, hence $Q \subset \oball{a}{r}$.
\end{proof}
\begin{remark} \label{remark:topsoe}
Whenever $\omega$ violates the Dini condition and $\lambda > 0$ there do
not exist $\varepsilon$, $G$ and $A$ as in \ref{example:cantor_set} as may
be verified by applying Tops{\o}e \cite[Theorem 3]{MR577961} with $\| x
\|$, $N$, $\mathscr{B}$, $\psi (r)$, and $\delta_0$ replaced by $\sup \{
|x_i| \with i = 1, \ldots, {m} \}$, ${m}$, $G$, $\omega (r)^{1/{m}}
r/2$, and $\varepsilon$.
\end{remark}
\begin{remark} \label{remark:mejlbro_topsoe}
The basic construction principle is that of Mejlbro and Tops{\o}e
\cite[Theorem 2]{MR0583557} who established the sharpness of the Dini
condition in a similar context.
\end{remark}
\begin{example}
\label{example:yet_another_cantor_set}
Suppose ${m}$, $W$, $Z$, and $N$ are as in
\ref{miniremark:family_of_cubes}, $\omega$ is a modulus of continuity, and
$0 \leq \lambda < 1$.
Then there exist $0 < \varepsilon \leq 1$, $B \subset \mathbf{R} \cap \{ r
\with r > 0 \}$ with $\inf B = 0$, and a disjointed subfamily $G$ of
$Z$ such that
\begin{gather*}
{\textstyle A = C(0,1) \without \left (
N \cup \bigcup G \right )} \quad \text{with} \quad
\mathscr{L}^{m} (A) \geq \lambda
\end{gather*}
satisfying the following condition: If $a \in A$ and $0 < r \leq
\varepsilon$, then there exists a subset $H$ of $G \cap \{ Q \with Q
\subset \oball{a}{r} \}$ such that
\begin{gather*}
\mathscr{L}^{m} ( {\textstyle \bigcup H} ) \geq \omega (r) r^{m}
\quad \text{and} \quad
\text{$\card H = 1$ if $r \in B$}.
\end{gather*}
\end{example}
\begin{proof} [Construction]
Choose $s$ with
\begin{gather*}
0 < s \leq 2^{-1} {m}^{-1/2},
\quad
\omega ( 2 {m}^{1/2} s) \leq 2^{-2{m}} {m}^{-{m}/2},
\end{gather*}
define $\phi : \mathbf{R} \cap \{ t
\with 0 \leq t \leq 1 \} \to \mathbf{R}$ by
\begin{gather*}
\phi (t) = 2^{2{m}} {m}^{{m}/2} \omega ( 2 {m}^{1/2} t)
\quad
\text{for $0 \leq t \leq s$},
\qquad
\phi (t) = 1
\quad
\text{for $s < t \leq 1$},
\end{gather*}
and note that $\phi$ is a modulus of continuity. Inductively choose
sequences $\alpha_i$ and $\beta_i$ of nonnegative integers subject to
the conditions
\begin{gather*}
\phi ( 2^{-\alpha_i} ) \leq ( 1- \lambda) 2^{-i},
\qquad
\alpha_{i+1} > \beta_i \geq \alpha_i,
\\
2^{{m} ( \alpha_{i+1} - \beta_{i+1} + 1)}
> \phi (2^{-\alpha_i} )
\geq 2^{{m} ( \alpha_{i+1} - \beta_{i+1} )}
\end{gather*}
whenever $i$ is a positive integer and, in case $i > 1$, define families
$F_i$ to consist of the open cubes $C(b,2^{-\beta_i})$, see
\ref{miniremark:family_of_cubes}, corresponding to all $C
(b,2^{-\alpha_i}) \in W ( \alpha_i )$ satisfying the following condition:
\begin{gather*}
\text{If $j$ is an integer, $1 < j \leq i-1$, and $R \in F_j$ then $C
(b,2^{-\beta_i}) \cap R = \varnothing$}.
\end{gather*}
Define $\varepsilon = \inf \{ 2^{-\alpha_1} {m}^{1/2}, 2 {m}^{1/2} s
\}$ and note that $\varepsilon \leq 1$. Let
\begin{gather*}
G = {\textstyle \bigcup} \{ F_i \with 1 < i \in \mathbf{Z} \},
\quad
{\textstyle A = C(0,1) \without \left ( N \cup \bigcup G \right )},
\\
B = \{ {m}^{1/2} 2^{1-\alpha_i} \with 1 < i \in \mathbf{Z} \}.
\end{gather*}
Observe that $G$ is disjointed. Since $\card F_i \le 2^{{m} \alpha_i}$
for any positive integer~$i$, we have
\begin{gather*}
\mathscr{L}^{m}({\textstyle \bigcup G})
\le \tsum{i = 2}{\infty} 2^{{m}(\alpha_i - \beta_i)}
\le \tsum{i = 2}{\infty} \phi(2^{-\alpha_{i-1}})
\le 1-\lambda .
\end{gather*}
Hence $\mathscr{L}^{m}(A) \ge \lambda$.
Suppose $a \in A$ and $0 < r \leq \varepsilon$.
There exist $k$ and $c$ satisfying
\begin{gather*}
\alpha_1 < k \in \mathbf{Z}, \quad 2^{-k} < {m}^{-1/2} r \leq
2^{1-k}, \quad a \in S = C (c,2^{-k}) \in W(k)
\end{gather*}
and $i$ such that
\begin{gather*}
1 < i \in \mathbf{Z} \quad \text{and} \quad \alpha_{i-1} < k
\leq \alpha_i.
\end{gather*}
Defining $I = W (\alpha_i) \cap \{ Q \with Q \subset S \}$, one notes
that
\begin{gather*}
\card I = 2^{{m}(\alpha_i-k)},
\end{gather*}
in particular $\card I = 1$ if $r \in B$. Moreover, one concludes
\begin{gather*}
C (b,2^{-\alpha_i}) \in I \quad \text{implies} \quad \text{$C
(b,2^{-\beta_i}) \subset R \subset S$ for some $R \in G$};
\end{gather*}
in fact, either $C (b,2^{-\beta_i}) \cap R \neq \varnothing$ for some
integer $j$ with $1 < j \leq i-1$ and some $R \in F_j$, hence
\begin{gather*}
R \cap S \neq \varnothing, \quad C (b,2^{-\beta_i}) \subset R
\subset S
\end{gather*}
as $\beta_i \geq \beta_j$ and $a \in S \without R$, because $a \in A \subset
C(0,1) \without \bigcup G$, or
\begin{gather*}
C (b,2^{-\beta_i}) \in F_i \subset G, \quad C (b,2^{-\beta_i})
\subset C (b,2^{-\alpha_i}) \subset S.
\end{gather*}
Consequently, there exists a subset $H$ of $G \cap \{ R \with R
\subset S \}$ such that
\begin{gather*}
{\textstyle \bigcup \{ C (b,2^{-\beta_i}) \with C
(b,2^{-\alpha_i}) \in I \} \subset \bigcup H}, \quad
\text{$\card H = 1$ if $r \in B$}.
\end{gather*}
Noting $2^{-1} {m}^{-1/2} r \leq \inf \{ 2^{-\alpha_{i-1}}, s \}$,
one infers
\begin{gather*}
\mathscr{L}^{m} ( {\textstyle \bigcup H} ) \geq 2^{{m} (
\alpha_i-\beta_i-k)} \geq 2^{-2{m}} {m}^{-{m}/2} \phi (
2^{-\alpha_{i-1}} ) r^{m} \geq \omega (r) r^{m}.
\end{gather*}
Since $S \subset \oball{a}{r}$, the conclusion follows.
\end{proof}
\begin{remark} \label{remark:vitali_lebesgue}
From the classical differentiation theory of Vitali and Lebesgue, see for
instance \cite[2.8.17, 2.9.11]{MR41:1976}, it is evident that the lower
bound on $\mathscr{L}^{m} ( \oball{a}{r} \without A )$ exhibited here is the
known optimal one, see Tolstoff \cite[Th\'eor\`eme 3, p.~263]{MR0013777}.
In order to demonstrate the sharpness of our results in
\ref{theorem:quadratic-tilt-decay}, the additional property ``$\card H =
1$ if $r \in B$'' will be employed in \ref{example:quadratic_tilt_excess}.
Clearly, $B$ may not be required to equal $\{ r \with 0 < r \leq
\varepsilon \}$ if $\omega$ violates the Dini condition and $\lambda > 0$
by \ref{remark:topsoe}.
\end{remark}
\section{Notation} \label{sec:notation}
The notation generally follows Federer \cite{MR41:1976} and Allard
\cite{MR0307015} with some modifications and additions described in \cite[\S
1]{snulmenn:tv.v2}. Here, concerning the paragraph ``Definitions in the text'' of
\cite[\S 1]{snulmenn:tv.v2}, only the notions of \emph{generalised weakly
differentiable function}, \emph{generalised weak derivative}, $\derivative
Vf$, and the associated space $\mathbf{T} (V,Y)$, see \cite[8.3]{snulmenn:tv.v2},
will be employed.
\paragraph{Less common symbols and terminology.} A family is disjointed if and
only if two distinct members have empty intersection, see
\cite[2.1.6]{MR41:1976}. Whenever $f$ is a function it is a subset of the
Cartesian product of its domain, $\dmn f$, and its image, $\im f$, see
\cite[p.~669]{MR41:1976}. The Lebesgue measure of the unit ball in
$\mathbf{R}^{m}$ will be denoted by $\unitmeasure {m}$, see
\cite[2.7.16]{MR41:1976}. Whenever $\mu$ measures $X$ and $Y$ is a Banach
space, $\mathbf{A} ( \mu, Y)$ denotes the vectorspace of those $Y$ valued
$\mu$ measurable functions $f$ such that there exists a separable subspace
$Z$ of $Y$ satisfying $\mu ( f^{-1} [ Y \without Z ] ) = 0$, see
\cite[2.3.8]{MR41:1976}. In this case,
\begin{gather*}
\Lpnorm{\mu}{p}{f} = \big ( \tint{}{} |f|^p \ensuremath{\,\mathrm{d}} \mu \big )^{1/p} \quad
\text{whenever $1 \leq p < \infty$}, \\
\Lpnorm{\mu}{\infty}{f} = \inf \{ t \with \mu ( \{ x \with |f(x)| > t
\} ) = 0 \}
\end{gather*}
whenever $f \in \mathbf{A} ( \mu, Y )$, see \cite[2.4.12]{MR41:1976}. Notice
also that no usage of equivalence classes of functions is made.
\paragraph{Modifications.} Extending Allard \cite[2.5\,(1)]{MR0307015},
whenever $M$ is a submanifold of $\mathbf{R}^{n}$ of class~$2$ and $a \in M$ the
\emph{second fundamental form of $M$ at $a$} is the unique symmetric bilinear
mapping
\begin{gather*}
\mathbf{b}(M,a) : \Tan (M,a) \times \Tan (M,a) \to \Nor (M,a)
\end{gather*}
such that whenever $g : M \to \mathbf{R}^{n}$ is of class $1$ and $g(z) \in \Nor
(M,z)$ for $z \in M$,
\begin{gather*}
u \bullet \langle v, \Der g (a) \rangle = - \mathbf{b} (M,a) ( u,v )
\bullet g (a) \quad \text{for $u,v \in \Tan (M,a)$}.
\end{gather*}
\paragraph{Additional notation.} By a~\emph{modulus of continuity} we mean
a~function $\omega : \{ t \with 0 \leq t \leq 1 \} \to \{ t \with 0 \leq t \leq
1 \}$ such that
\begin{gather*}
\lim_{t \to 0+} \omega (t) = 0, \qquad \text{$\omega (t) = 0$ if and only
if $t=0$ whenever $0 \leq t \leq 1$},
\\
\text{$\omega(s) \leq \omega(t)$ whenever $0 \leq s \leq t \leq 1$}.
\end{gather*}
Adapting Morrey \cite[p.~54]{MR0202511}, such $\omega$ is said to satisfy the
\emph{Dini condition} if and only if $\tint{0}{1} \omega(t)t^{-1} \ensuremath{\,\mathrm{d}} \mathscr{L}^1 t <
\infty$. Following \cite[p.~8]{snulmenn.decay}, whenever ${n},{m} \in
\mathscr{P}$, ${m} < {n}$, $T \in \grass{{n}}{{m}}$, $a \in \mathbf{R}^{n}$, $0 < r
< \infty$, and $0 < h < \infty$ we define the \emph{closed cuboid} by
\begin{gather*}
\cylinder Tarh = \mathbf{R}^{n} \cap \big \{ z \with |\project T (z-a)| \le r
\text{ and } |\perpproject T (z-a)| \le h \big \} .
\end{gather*}
\paragraph{Definitions in the text.} The notions of \emph{curvature varifold}
and its \emph{second fundamental form} $\mathbf{b}(V,z)$, are explained in
\ref{def:curvature_varifold}. The Orlicz space seminorm $\mu_{(\Phi)}$ is
defined in \ref{def:orlicz}. The space $\mathbf{Q}_Q ( Y )$ metrised by
$\mathscr{G}$ occurs in \ref{miniremark:almgren1}, the notion of \emph{affine}
$\mathbf{Q}_Q ( Y )$ valued function and the corresponding seminorm $\| f \|$ are
explained in \ref{miniremark:almgren2}. Finally, the notions of \emph{affinely
approximable} and \emph{approximately affinely approximable} for $\mathbf{Q}_Q (
Y)$ valued functions and the corresponding symbols $\Aff f$ and $\ap \Aff f$
are defined in \ref{miniremark:almgren2}.
\section{The one dimensional case}
\label{sec:one_dimensional_decay}
For completeness, we consider in this section one dimensional integral varifolds
of locally bounded first variation. In that case we prove that there is a set,
almost equal to the support of the weight measure of the varifold, such that the
tangent map of the varifold is differentiable relative to this set almost
everywhere, see \ref{thm:decay-1d}. This implies that near almost all points the
varifold may be expressed by a finite sum of graphs of Lipschitzian functions,
see \ref{corollary:1-dim-Q-graph}.
The differentiability result for the tangent map mainly relies on an adaptation
of a coercive estimate of Allard and Almgren in \cite[\S 5]{MR0425741} in
conjunction with differentiability results of approximate and integral nature
obtained for that map in~\cite[\S 5]{snulmenn.c2}. The corollary then follows
from a suitable approximation by Lipschitzian $\mathbf{Q}_Q ( \mathbf{R}^{{n}-1} )$
valued functions, see~\cite[3.15]{snulmenn.poincare}, in combination with
a~structural result for such function, see Almgren \cite[1.10]{MR1777737}.
\begin{theorem}
\label{lemma:sup-L1-est}
Suppose $U$ is an open subset of $\mathbf{R}^{n}$, $V \in \RVar_1 ( U )$, $M =
\{ z \with 0 < \boldsymbol{\Theta}^{{m}} ( \| V \|, z ) < \infty \}$, $\mu$ is a
Radon measure over $U$, $f : U \to \mathbf{R}$ is a Lipschitzian function,
\begin{gather*}
\tint{}{} \ap \Der f \bullet \ap \Der \theta \ensuremath{\,\mathrm{d}} \| V \| \leq \mu ( \theta )
\quad \text{whenever $\theta \in \mathscr{D} ( U, \mathbf{R} )$ and $\theta
\geq 0$},
\end{gather*}
where ``$\ap$'' denotes approximate differentiation with respect to $(\| V
\|, 1)$, and $\phi \in \mathscr{D} ( U, \mathbf{R} )$ with $\phi \geq 0$.
Then there holds
\begin{gather*}
\eqLpnorm{\mathscr{H}^1 \mathop{\llcorner} M}{\infty}{ \phi | \ap \Der f | \boldsymbol{\Theta}^1 ( \| V \|, \cdot )}
\leq \Lip ( \phi ) \tint{\spt \Der \phi}{} | \ap \Der f| \ensuremath{\,\mathrm{d}} \| V \| + \mu ( \phi ).
\end{gather*}
\end{theorem}
\begin{proof}
Abbreviate $\gamma = \Lip ( \phi) \tint{\spt \Der \phi}{} | \ap \Der f | \ensuremath{\,\mathrm{d}} \| V \|
+ \mu ( \phi )$ and define
\begin{gather*}
h(t) = \tint{M \cap \{ z \with f(z) = t \}}{}
\phi (z) | \ap \Der f(z) | \boldsymbol{\Theta}^1 ( \| V \|, z ) \ensuremath{\,\mathrm{d}} \mathscr{H}^0 z
\quad \text{whenever $t \in \mathbf{R}$}
\end{gather*}
and $T = \{ t \with h(t) \leq \gamma \}$. Since
\ref{miniremark:rect_varifold}\,\eqref{item:rect:varifold:coarea} and
Allard \cite[3.5\,(1b)]{MR0307015} imply
\begin{gather*}
\ap \Der f(z) = 0
\quad
\text{for $\mathscr{H}^1$ almost all $z \in M \cap f^{-1} [ N ]$}
\end{gather*}
whenever $\mathscr{L}^1 (N) = 0$, it is sufficient to prove $\mathscr{L}^1
( \mathbf{R} \without T ) = 0$.
Approximating $\theta$ by convolution and using
\cite[4.5\,(3)]{snulmenn.decay}, one obtains
\begin{gather*}
\tint{}{} \ap \Der f \bullet \ap \Der \theta \ensuremath{\,\mathrm{d}} \| V \| \leq \mu (\theta)
\end{gather*}
whenever $\theta : U \to \mathbf{R}$ is a nonnegative Lipschitzian function with
compact support. Employing
\ref{miniremark:rect_varifold}\,\eqref{item:rect:varifold:coarea} with
$g(z)$ replaced by $\phi (z) \psi'(f(z)) | \ap \Der f (z) |$ and taking
$\theta = \phi \cdot ( \psi \circ f )$ yields
\begin{align*}
\tint{}{} \psi' h \ensuremath{\,\mathrm{d}} \mathscr{L}^1
& = \tint{}{} \phi ( \psi' \circ f ) | \ap \Der f|^2 \ensuremath{\,\mathrm{d}} \| V \|
\\
& = \tint{}{} \ap \Der f \bullet \ap \Der \theta \ensuremath{\,\mathrm{d}} \| V \|
- \tint{}{} ( \psi \circ f ) \ap \Der f \bullet \ap \Der \phi \ensuremath{\,\mathrm{d}} \| V \|
\leq \gamma
\end{align*}
whenever $\psi \in \mathscr{E} ( \mathbf{R}, \mathbf{R} )$ and $0 \leq \psi \leq 1$.
Letting $\psi$ approach the characteristic function of $\{ u \with t < u
\}$ shows that $t \in T$ whenever $t$ is a Lebesgue point of $h$ and the
conclusion follows from \cite[2.8.18, 2.9.8]{MR41:1976}.
\end{proof}
\begin{remark}
\label{remark:sup-L1-est}
If $\theta \in \mathscr{D} (U,\mathbf{R})$ and $f \in \Hom ( \mathbf{R}^{n}, \mathbf{R})$,
then $( \delta V ) ( \theta \grad f ) = \tint{}{} \ap \Der f \bullet \ap \Der
\theta \ensuremath{\,\mathrm{d}} \| V \|$. Consequently, if $\| \delta V \|$ is a Radon measure,
then
\begin{gather*}
\Lpnorm{\| V \|}{\infty}{ \phi | \ap \Der L | }
\leq {n} \big ( \| L \| \| \delta V \| ( \phi )
+ \Lip ( \phi ) \tint{\spt \phi}{} | \ap \Der L | \ensuremath{\,\mathrm{d}} \| V \| \big )
\end{gather*}
whenever $L \in \Hom ( \mathbf{R}^{n}, \mathbf{R}^{n} )$.
\end{remark}
\begin{remark} \label{remark:laplace-beltrami}
The method of proof originates from Allard and Almgren
\cite[5\,(6)]{MR0425741}. Adapting the terminology of \cite[p.~41, p.~188,
p.~391]{MR1814364} to varifolds, our presentation views $f$ as a ``weak
subsolution to Poisson's equation for the Laplace-Beltrami operator on
$V$'', see also Allard \cite[7.5]{MR0307015}.
\end{remark}
\begin{theorem}
\label{thm:decay-1d}
Suppose $1 < {n} \in \mathscr{P}$, $U$ is an open subset of $\mathbf{R}^{n}$,
$V \in \IVar_1 ( U )$, $\| \delta V \|$ is a Radon measure,
\begin{gather*}
C = \{ (z,\cball{z}{r}) \with \text{$z \in U$, $0 < r <
\infty$} \}, \quad Z = U \cap \{ z \with \Tan^1 ( \| V \|, z )
\in \grass{{n}}{1} \},
\end{gather*}
$\tau : Z \to \Hom ( \mathbf{R}^{n}, \mathbf{R}^{n})$ satisfies $\tau(z) =
\project{\Tan^1 ( \| V \|, z )}$ for $z \in Z$, and $A$ is the set of
points in $\spt \| V \|$ at which $\tau$ is $( \| V \|, C )$
approximately continuous.
Then $\| V \| ( U \without A ) = 0$ and, for $\| V \|$ almost all $z
\in A$, $\tau|A$ is differentiable relative to $A$ at~$z$ with
\begin{gather*}
\Der (\tau|A) (z) = ( \| V \|, 1 ) \ap \Der \tau (z) .
\end{gather*}
\end{theorem}
\begin{proof}
First, notice that $\| V \| ( U \without A ) = 0$ by \cite[2.8.18,
2.9.13]{MR41:1976} and that, for $\| V \|$ almost all $z$, $\tau$ is
$( \| V \|, 1 )$ approximately differentiable at $z$ and
\begin{gather*}
\limsup_{r \to 0+} r^{-2} \tint{\cball{z}{r}}{} | \tau (
\zeta ) - \tau ( z )| \ensuremath{\,\mathrm{d}} \| V \| z < \infty
\end{gather*}
by \cite[5.2\,(2), 5.5]{snulmenn.c2}. If $z$ additionally satisfies
$\boldsymbol{\Theta}^{\ast 1} ( \| \delta V \|, z ) < \infty$, as $\| V \|$
almost all $z$ do by \cite[2.10.19\,(3)]{MR41:1976}, then
\begin{gather*}
\limsup_{A \owns \zeta \to z} | \tau (\zeta) - \tau (z)
|/|\zeta-z| < \infty;
\end{gather*}
in fact, it is sufficient to take $L = \id{\mathbf{R}^{n}} - \tau (z)$ and a
suitable $\phi$ in \ref{remark:sup-L1-est} since $| \tau (\zeta) - \tau
(z)|^2 = 2 | L \circ \tau (\zeta) |^2$ for $\zeta \in Z$ by Allard
\cite[8.9\,(1)\,(2)]{MR0307015},
Next, one obtains a sequence of functions $\tau_i : U \to \Hom (
\mathbf{R}^{n}, \mathbf{R}^{n} )$ of class $1$ such that the sets $Z_i = Z
\cap \{ z \with \tau (z) = \tau_i (z) \}$ cover $\| V \|$ almost all
of $U$ and
\begin{gather*}
( \| V \|, 1 ) \ap \Der \tau (z) = \Der \tau_i (z) | \Tan^1 ( \| V
\|, z) \quad \text{for $\| V \|$ almost all $z \in Z_i$}
\end{gather*}
by \cite[11.1\,(2)\,(4)]{snulmenn:tv.v2} and \cite[3.2.16]{MR41:1976}.
Defining $f_i = ( \tau-\tau_i ) |A$, the preceding paragraph yields
\begin{gather*}
\limsup_{\zeta \to z} |f_i (\zeta)| / |\zeta-z| <
\infty \quad \text{for $\| V \|$ almost all $z \in Z_i$},
\end{gather*}
hence, in view of \ref{thm:differentiation} and
\cite[3.1.22]{MR41:1976}, it follows
\begin{gather*}
\lim_{\zeta \to z} | f_i (\zeta)|/|\zeta-z| = 0, \quad \Der (
\tau |A ) (z) = \Der \tau_i (z) | \Tan (A,z)
\end{gather*}
for $\| V \|$ almost all $z \in Z_i$. Noting
\begin{gather*}
\Tan^1 ( \| V \|, z ) \subset \Tan (A,z) \subset \Tan (\spt \|
V \|, z) \quad \text{for $z \in U$}
\end{gather*}
by \cite[3.2.16]{MR41:1976}, the conclusion now follows from
\cite[11.3]{snulmenn:tv.v2}.
\end{proof}
\begin{corollary}
\label{corollary:1-dim-Q-graph}
Suppose $1 < {n} \in \mathscr{P}$, $U$ is an open subset of $\mathbf{R}^{n}$, $V
\in \IVar_1(U)$, $\| \delta V \|$ is a Radon measure, and $\varepsilon >
0$.
Then, for $V$ almost all $(z,T)$, there exist $Q \in \mathscr{P}$, $0 <r<
\infty$, and $f_i : T \cap \cball{\project{T} (z)}{r} \to T^\perp \cap
\cball{ \perpproject{T} (z)}{r}$ with $\Lip f_i \leq \varepsilon$ for $i =
1, \ldots, Q$ such that
\begin{gather*}
\boldsymbol{\Theta}^1 ( \| V\|, \zeta ) = \card \{ i \with f_i ( \project{T}
( \zeta ) ) = \perpproject{T} (\zeta) \}
\end{gather*}
for $\mathscr{H}^1$ almost all $\zeta \in \cylinder Tzrr$.
\end{corollary}
\begin{proof}
Let $Z$, $\tau$ and $A$ be defined as in~\ref{thm:decay-1d}, in particular
$\| V \| ( U \without A ) = 0$. In~view of~\ref{thm:decay-1d} and Allard
\cite[3.5\,(1)]{MR0307015} it is sufficient to prove the conclusion at
a~point $(z,T)$ such that $z \in A$, $T = \im \tau (z)$, $\| \delta V \| (
\{ z \} ) = 0$, $\tau|A$ is continuous at~$z$, and, for some $Q \in \mathscr{P}$,
also
\begin{gather*}
r^{-1} \tint{}{} k (r^{-1} (\zeta-z), S ) \ensuremath{\,\mathrm{d}} V(\zeta,S) \to Q \tint
T{} k ( \zeta, T ) \ensuremath{\,\mathrm{d}} \mathscr{H}^1 \zeta \quad \text{as $r \to 0+$}
\end{gather*}
for $k \in \mathscr{K} ( \mathbf{R}^{n} \times \grass {n} 1 )$. Define
$\delta_1 = \delta_2 = \delta_3 = 1/2$, $\delta_4 = 1/4$, and $\delta_5 = (2
\unitmeasure 1 \isoperimetric 1 )^{-1}$ and recall $\delta_5 \leq 1$, see
e.g.~\cite[2.4]{snulmenn.isoperimetric}. From \cite[3.15]{snulmenn.poincare}
one obtains
\begin{gather*}
\lambda = \varepsilon_{\text{\cite[3.15]{snulmenn.poincare}}} (
{n}-1, 1, Q, \varepsilon, 5Q, \delta_1, \delta_2, \delta_3,
\delta_4, \delta_5 ).
\end{gather*}
Choose $0 < r < \infty$ such that
\begin{gather*}
(Q - 1/2) \unitmeasure{1} r \le \|V\|(\cylinder Tzrr) \le (Q + 1/2)
\unitmeasure{1} r , \\
\|V\|(\cylinder Tzr{5r/4} \without \cylinder Tzr{r/2}) \le (1/2)
\unitmeasure{1} r , \\
\oball z{4r} \subset U, \quad \measureball{\|V\|}{\oball z{4r}} \le 5Q
\unitmeasure{1} r , \quad \measureball{\|\delta V\|}{\oball z{4r}} \le
\lambda, \\
| \tau(\zeta) - \tau(z) | \le \lambda \quad \text{whenever $\zeta \in
A \cap \oball z{4r}$}.
\end{gather*}
Applying \cite[3.15\,(1)--(3)]{snulmenn.poincare} with $m$, $n$, $L$, $M$,
$a$, $h$, and $\varepsilon_1$ by ${n}-1$, $1$, $\varepsilon$, $5Q$, $z$,
$r$, and $\lambda$ and noting that the set $B$ occurring there is empty,
one infers the existence of a function $f$ with values in $\mathbf{Q}_Q (
T^\perp )$ with $\dmn f \subset T \cap \cball{ \project{T} (z) }{r}$,
$\Lip f \leq \varepsilon$, and
\begin{gather*}
\spt f (x) \subset \cball { \perpproject{T} (z)} r \quad \text{for $x
\in \dmn f$}, \\
\boldsymbol{\Theta}^1 ( \| V \|, \zeta ) = \boldsymbol{\Theta}^0 ( \| f ( \project{T}
(\zeta) ) \|, \perpproject{T} (\zeta) ) \quad \text{for $\zeta \in
\cylinder Tzrr \cap \project{T}^{-1} [ \dmn f ]$}, \\
\mathscr{H}^1 ( T \cap
\cball{\project{T} (z)}{r} \without \dmn f) + \| V \| \big ( \cylinder
Tzrr \without \project{T}^{-1} [ \dmn f ] \big ) = 0.
\end{gather*}
Consequently, \cite[2.10.19\,(4)]{MR41:1976} implies
\begin{gather*}
\boldsymbol{\Theta}^1 ( \| V \|, \zeta ) = 0 \quad \text{for $\mathscr{H}^1$
almost all $\zeta \in \cylinder Tzrr \without \project{T}^{-1} [
\dmn f ]$}.
\end{gather*}
Defining $g = \Clos f$, one infers that $g$ is a function, $\dmn g = T \cap
\cball {\project{T}(z)}r$, $\Lip g \leq \varepsilon$, and $\spt g(x) \subset
\cball{\perpproject{T}(z)}{r}$ for $x \in \dmn g$. Now one readily verifies
\begin{gather*}
\boldsymbol{\Theta}^0 ( \| g ( \project{T} (\zeta) ) \|, \perpproject{T} (\zeta))
= 0 \quad \text{for $\mathscr{H}^1$ almost all $\zeta \in \cylinder
Tzrr \without \project{T}^{-1} [ \dmn f ]$}
\end{gather*}
and the conclusion, both by means of Almgren \cite[1.10\,(2),
1.10\,(1)\,(iii)]{MR1777737}.
\end{proof}
\begin{remark}
The fact $\mathscr{H}^{m} ( \{ (x,y) \with \text{$x \in X$ and $y \in \spt
g(x)$} \} ) = 0$ whenever $\mathscr{L}^{m} ( X ) = 0$ and $g : X \to
\mathbf{Q}_Q ( \mathbf{R}^{n-m} )$ is Lipschitzian, deduced from Almgren
\cite[1.10\,(2), 1.10\,(1)\,(iii)]{MR1777737} for ${m} = 1$ in the
preceding proof, clearly holds for arbitrary ${n} > {m} \in \mathscr{P}$ by
Almgren \cite[1.5\,(11)\,(iii)\,(c)]{MR1777737} or
\cite[2.5\,(1)]{snulmenn.poincare}.
\end{remark}
\section{Quadratic tilt-excess, decay rates}
\label{sec:quadratic_tilt_decay}
In this section we prove sharp decay rates of the quadratic tilt-excess for two
dimensional integral varifolds whose first variation is a Radon measure, see
\ref{theorem:quadratic-tilt-decay}. This result rests on two pillars. Firstly,
on the second order rectifiability of such varifolds obtained
in~\cite[4.8]{snulmenn.c2}. Secondly, on an approximate coercive estimate by
which we mean an estimate of the tilt-excess in terms of the first variation,
the height-excess measured on a set of suitably large weight measure and small
contributions from the tilt-excess, see \ref{lemma:app-coercive}.
Accordingly, in order to derive the approximate coercive estimate
\ref{lemma:app-coercive} from the coercive estimate
\ref{lemma:coercive-estimate}, one needs to estimate the height-excess
occurring in~\ref{lemma:coercive-estimate} by approximate height-quantities
together with the variation measure of the first variation and quadratic
tilt-excess. The approximation \ref{lemma:mini-lip-approx} reduces such an
estimate to the case of a real valued Lipschitzian functions which has been
treated in \ref{lemma:int-ineq} and \ref{corollary:orlicz-poincare}.
Since currently no analogous estimates to \ref{lemma:int-ineq} and
\ref{corollary:orlicz-poincare} are available for real valued Lipschitzian
function over varifolds, the authors have chosen the path using the
approximation by $\mathbf{Q}_Q ( \mathbf{R}^{n-m} )$ valued functions leading to
\ref{lemma:mini-lip-approx}. Yet, it would be of interest to investigate whether
the embedding theory for Lipschitzian functions on varifolds can be extended so
as to yield a proof without such approximation. In a somewhat different vein
much of that theory has been extended to generalised weakly differentiable
Functions in~\cite[\S 10]{snulmenn:tv.v2}.
\begin{lemma}
\label{lemma:app-coercive}
Suppose $2 < {n} \in \mathscr{P}$, $Q \in \mathscr{P}$, $c \in \mathbf{R}^{n}$, $0 < r <
\infty$, $V \in \IVar_2 ( \oball c{8r} )$,
\begin{gather*}
\measureball{\| V \|}{\cball c{2r}} \geq (Q-1/2) \unitmeasure 2 (2r)^2,
\quad
\measureball{\| V \|}{ \oball c{8r} } \leq (Q+1/4) \unitmeasure 2 (8r)^2,
\end{gather*}
$T \in \grass {n} 2$, $0 < \varepsilon \leq 1$, $Z$ is $\| V \|$
measurable, and
\begin{gather*}
\alpha = r^{-1} \measureball{\| \delta V\|}{ \oball c{8r} },
\quad
\beta = r^{-1} \big ( \tint{}{} \| \project{S} - \project{T} \|^2 \ensuremath{\,\mathrm{d}} V(z,S) \big )^{1/2},
\quad
\alpha + \beta \leq \varepsilon,
\\
\gamma = r^{-2} \big ( \tint{Z}{} \dist (z-c,T)^2 \ensuremath{\,\mathrm{d}} \| V \| z \big)^{1/2},
\quad
\| V \| ( \oball c{8r} \without Z ) \leq (\varepsilon r )^2,
\end{gather*}
and $\kappa : \{ t \with 0 \leq t < \infty \} \to \mathbf{R}$ satisfies (see~\ref{miniremark:kappa})
\begin{gather*}
\kappa (0) = 0,
\qquad
\kappa (t) = t \big ( 1 + (\log (1+1/t))^{1/2} \big )
\quad \text{for $0 < t < \infty$}.
\end{gather*}
Then there holds
\begin{gather*}
r^{-2} \tint{\oball cr \times \grass {n} 2}{}
\| \project{S} - \project{T} \|^2 \ensuremath{\,\mathrm{d}} V(z,S)
\leq \Gamma \big ( \kappa ( \alpha ( \alpha + \beta + \gamma ) )
+ \varepsilon \beta^2 + \varepsilon^{-1} \gamma^2 \big ),
\end{gather*}
where $\Gamma$ is a positive, finite number depending only on ${n}$
and~$Q$.
\end{lemma}
\begin{proof}
Considering $(\boldsymbol{\mu}_{1/r})_{\#}V$ in place of $V$ one may assume
$r = 1$ and, using isometries, one may assume $c = 0$ and $T = \im
\mathbf{p}^\ast$, see \ref{miniremark:pqT}. Moreover, one may assume $Z$ to be a
Borel set.
Define
\begin{gather*}
\Delta_1 = 1 + \Gamma_{\ref{lemma:mini-lip-approx}} ( 2 ,{n},Q),
\quad
\Delta_2 = \inf \big \{ 1/3, 2 \Delta_1^{-1/2} \big \},
\\
\Delta_3 = 2 (Q+1/2)^{1/2} (Q+3/8)^{-1/2},
\quad
\Delta_4 = (2Q)^{1/2}+\Delta_1^{1/2},
\\
\Delta_5 = \Delta_1^{5/4} \Gamma_{\ref{lemma:int-ineq}}(2) 2^{1/2},
\quad
\Delta_6 = \inf \big \{ 1, ( \Delta_3^2 - 4 )^{1/2}/4 \big \},
\quad
\Delta_7 = \Delta_5 ( 1 + \Delta_4 ),
\\
\Delta_8 = 4 \Delta_1 ( \Gamma_{\ref{corollary:orlicz-poincare}} ( 2) +1 ) ( \Delta_4 + 1 ),
\\
\Delta_9 = \sup \{ \Delta_2^{-1}, \Delta_6^{-2} \},
\quad
\Gamma = \sup \{ \Delta_9, \Gamma_{\ref{lemma:coercive-estimate}} ( 2
) ( 1 + 3 \Delta_7^2 + \Delta_8 ) \}.
\end{gather*}
If $\varepsilon > \Delta_2$ then $\beta^2 \le \Delta_2^{-1} \varepsilon
\beta^2 \leq \Delta_9 \varepsilon \beta^2$ and if $\gamma > \Delta_6$,
then $\beta^2 \le \varepsilon^2 \le 1 \le \Delta_6^{-2} \varepsilon^{-1}
\gamma^2 \leq \Delta_9 \varepsilon^{-1} \gamma^2$. Therefore one may
assume $\varepsilon \leq \Delta_2$ and $\gamma \leq \Delta_6$.
Abbreviate $C = \cylinder T022$, $K = \mathbf{R}^{n} \cap \{ z \with \dist
(z,C) \leq 4 \}$, and
\begin{gather*}
H = C \cap \{ z \with \measureball{\| V \|}{\cball zs} \geq (80
\isoperimetric{ 2 } )^{-2} s^2 \text{ for $0 < s < 4$} \}.
\end{gather*}
Notice that
\begin{gather*}
\oball 01 \subset \{ z \with \oball z1 \subset C \},
\quad
K \subset \oball 08,
\\
C \cap H_{\ref{lemma:coercive-estimate}} \subset H,
\quad
\text{where $H_{\ref{lemma:coercive-estimate}}$ denotes the set
named ``$H$'' in \ref{lemma:coercive-estimate}}.
\end{gather*}
In order to apply \ref{lemma:mini-lip-approx} with ${m}$ and $r$
replaced by $2$ and $2$, one estimates
\begin{gather*}
\measureball{\| V \|}{ \oball 08} \leq 100 Q \unitmeasure 2,
\\
\begin{aligned}
\| V \| ( \cylinder T024 \without \cylinder T021 )
& \leq \| V \| ( \oball 08 \cap \{ z \with \dist (z,T) \geq 1 \} )
\\
& \leq \| V \| ( \oball 08 \without Z ) + \gamma^2
\leq \varepsilon^2 + 1
\leq 2 \unitmeasure{2},
\end{aligned}
\end{gather*}
and, noting $2 < \Delta_3 < 8$ and
\begin{gather*}
C \subset \oball 0{\Delta_3} \cup
\big (
\oball 08 \cap \{ z \with \dist (z,T)^2 \geq \Delta_3^2 - 4 \} \big
),
\end{gather*}
one employs the monotonicity identity (see for instance \cite[4.5,
4.6]{snulmenn:tv.v2}) and the bounds on $\varepsilon$ and $\gamma$ to infer
\begin{align*}
\| V \| ( C )
&\leq \measureball{\| V \|}{ \oball 0{\Delta_3} } + \| V \| ( \oball
08 \cap \{ z \with \dist (z,T)^2 \geq \Delta_3^2-4 \} ) \\
&\leq \Delta_3^2 \big (
8^{-2} \measureball{\| V \|}{ \oball 08 }
+ \tint 28 t^{-2} \measureball{\| \delta V \|}{ \oball 0t } \ensuremath{\,\mathrm{d}} \mathscr{L}^1t
\big )
\\
& \qquad + \| V \| ( \oball 08 \without Z ) + ( \Delta_3^2-4)^{-1}
\gamma^2 \\
&\leq \Delta_3^2 \big (
(Q+1/4) \unitmeasure 2 + \tfrac 38 \alpha
+ \varepsilon^2 + ( \Delta_3^2-4)^{-1} \gamma^2
\big )
\\
&\leq \Delta_3^2 ( Q + 3/8 ) \unitmeasure 2
= 4 ( Q + 1/2 ) \unitmeasure 2.
\end{align*}
Using the hypotheses one also gets $\|V\|(C) \ge 4 ( Q - 1/2 )
\unitmeasure{2}$. Therefore, applying \ref{lemma:mini-lip-approx} with
${m}$ and $r$ replaced by~$2$ and $2$ in conjunction with Kirszbraun's
theorem, see~\cite[2.10.43]{MR41:1976}, one obtains a Borel set $X$ and a
function $f : \mathbf{R}^2 \to \mathbf{R}$ such that $f|X$ satisfies the conditions of
\ref{lemma:mini-lip-approx} and $\Lip f \leq 1$, in particular $f$ is
weakly differentiable with $\operatorname{\mathbf{D}} f(x) = \Der f(x)$ for $\mathscr{L}^2$ almost all
$x$ by \cite[2.13, 2.14]{MR2003a:49002}. Define
\begin{gather*}
A = X \without \mathbf{p} [ C \cap
\{ z \with \boldsymbol{\Theta}^2 ( \| V \|, z ) \in \mathscr{P} \} \without Z ]
\end{gather*}
and notice that $A$ is $\mathscr{L}^2$ measurable by \cite[2.55]{MR2003a:49002}
and~\cite[2.2.13]{MR41:1976}. Since
\begin{multline*}
\mathscr{L}^2 ( \mathbf{p} [ C \cap
\{ z \with \boldsymbol{\Theta}^2 ( \| V \|, z ) \in \mathscr{P} \} \without Z ] )
\\
\leq \mathscr{H}^2 ( C \cap \{ z \with \boldsymbol{\Theta}^2 ( \| V \|, z ) \in \mathscr{P} \} \without Z )
\leq \| V \| ( C \without Z )
\leq \varepsilon^2
\end{multline*}
by \cite[2.10.35]{MR41:1976} and Allard \cite[3.5\,(1b)]{MR0307015}, one
infers from
\ref{lemma:mini-lip-approx}\,\eqref{item:mini-lip-approx:bad-set} that
\begin{gather*}
\mathscr{L}^2 ( \oball 02 \without A )
\le \mathscr{L}^2 ( \cball 02 \without X ) + \mathscr{L}^2 ( X \without A )
\le \Delta_1 \varepsilon^2
\le 2 \unitmeasure 2.
\end{gather*}
Noting \cite[3.5\,(1c)]{MR0307015} and~\cite[2.8.17, 2.9.11,
3.1.2]{MR41:1976} and observing that from the definition of $A$ it follows
that $H \cap \mathbf{p}^{-1} [ A ] \cap \{ z \with \boldsymbol{\Theta}^2 ( \| V \|, z
) \in \mathscr{P} \} \subset Z$, one applies
\ref{lemma:mini-lip-approx}\,\eqref{item:mini-lip-approx:inverse-2-height-control}
and~\ref{lemma:mini-lip-approx}\,\eqref{item:mini-lip-approx:bad-set}\,\eqref{item:mini-lip-approx:tilt-control}
to obtain the following auxiliary estimates
\begin{gather*}
\eqLpnorm{ \mathscr{L}^2 \mathop{\llcorner} A }{2}{f} \leq \gamma, \quad
\eqLpnorm{ \mathscr{L}^2 \mathop{\llcorner} \oball 02}{2}{\Der f} \leq \Delta_4 (
\alpha + \beta ).
\end{gather*}
Next, defining $\Phi$ as in \ref{miniremark:Phi}, it will be shown that
\begin{align*}
\eqLpnorm{\| V \| \mathop{\llcorner} H} 2 {\perpproject T} & \leq \Delta_7 (
\alpha + \varepsilon^{1/2} \beta + \varepsilon^{-1/2} \gamma ), \\
\eqLpnorm{ \| V \| \mathop{\llcorner} H}{\Phi}{\perpproject T} & \leq \Delta_8 (
\alpha + \beta + \gamma ).
\end{align*}
To prove the first estimate, one notes $\alpha^2 \le \alpha$, $\beta^2 \le
\varepsilon \beta \le \varepsilon^{1/2} \beta $ and applies
\ref{lemma:mini-lip-approx}\,\eqref{item:mini-lip-approx:2-height-control}
and \ref{lemma:int-ineq} with $r$ and $\varepsilon$ replaced by $2$ and
$2^{-1} \Delta_1^{1/2} \varepsilon$ to deduce
\begin{align*}
& \eqLpnorm{\| V \| \mathop{\llcorner} H}{2} { \perpproject T}
\leq \Delta_1 \big (
\eqLpnorm{\mathscr{L}^2 \mathop{\llcorner} \oball 02} 2f + \alpha^2 + \beta^2
\big ) \\
& \qquad \leq \Delta_5 \big (
\varepsilon^{1/2} \eqLpnorm{\mathscr{L}^2 \mathop{\llcorner} \oball 02}{2}{\Der f}
+ \varepsilon^{-1/2} \eqLpnorm{\mathscr{L}^2 \mathop{\llcorner} A} 2f
+ \alpha + \varepsilon^{1/2} \beta
\big ),
\end{align*}
hence the estimate follows using the auxiliary estimates. To prove the
second estimate, one employs
\ref{remark:basic_orlicz}\,\eqref{item:basic_orlicz:estimate},
\ref{lemma:mini-lip-approx}\,\eqref{item:mini-lip-approx:Phi-height-control},
and \ref{corollary:orlicz-poincare} to infer
\begin{align*}
& \eqLpnorm{\| V \| \mathop{\llcorner} H}{\Phi} { \perpproject{T} }
\leq 4 \Delta_1 \big (
\eqLpnorm{\mathscr{L}^2 \mathop{\llcorner} \oball 02}{\Phi/4}{f} + \alpha + \beta
\big )
\\
& \quad \leq 4 \Delta_1 (\Gamma_{\ref{corollary:orlicz-poincare}} ( 2 ) + 1 )
\big (
\eqLpnorm{\mathscr{L}^2 \mathop{\llcorner} \oball 02} 2 {\Der f} +
\eqLpnorm{\mathscr{L}^2 \mathop{\llcorner} A}{2}{f} + \alpha + \beta
\big ),
\end{align*}
hence the estimate follows from the auxiliary estimates.
To conclude the proof, one employs \ref{lemma:coercive-estimate} to obtain
\begin{align*}
& \tint{ \oball 01 \times \grass {n} 2}{} \| \project{S} -
\project{T} \|^2 \ensuremath{\,\mathrm{d}} V(z,S) \\
& \qquad \leq \Gamma_{\ref{lemma:coercive-estimate}} (2) \big (
\alpha^2 + \kappa ( \Delta_8 \alpha ( \alpha + \beta + \gamma ) ) + 3
\Delta_7^2 ( \alpha^2 + \varepsilon \beta^2 + \varepsilon^{-1}
\gamma^2 ) \big ),
\end{align*}
hence, noting $\alpha^2 \leq \kappa ( \alpha^2 )$ and $\kappa ( \Delta_8
\alpha ( \alpha + \beta + \gamma ) ) \leq \Delta_8 \kappa ( \alpha ( \alpha
+ \beta + \gamma ) )$ by \ref{miniremark:kappa}, the conclusion is now
readily derived.
\end{proof}
\begin{theorem}
\label{theorem:quadratic-tilt-decay}
Suppose $2 < {n} \in \mathscr{P}$, $U$ is an open subset of~$\mathbf{R}^{n}$, $V
\in \IVar_2(U)$, and $\| \delta V \|$ is a Radon measure.
Then, for $V$ almost all $(z,T)$, there holds
\begin{gather*}
\lim_{r \to 0+} r^{-4} ( \log (1/r) )^{-1}
\tint{\cball zr \times \grass {n} 2}{}
\| \project{S} - \project{T} \|^2 \ensuremath{\,\mathrm{d}} V (\zeta,S) = 0.
\end{gather*}
\end{theorem}
\begin{proof}
Define $Z = U \cap \{ z \with \Tan^2 ( \| V \|, z ) \in \grass {n} 2 \}$
and $\tau : Z \to \Hom ( \mathbf{R}^{n}, \mathbf{R}^{n} )$ by $\tau (z) =
\project{\Tan^2 ( \| V \|, z )}$ for $z \in Z$. Recall that
\begin{gather*}
V(k) = \tint{Z}{} k(z,\tau(z)) \boldsymbol{\Theta}^2 ( \| V \|, z ) \ensuremath{\,\mathrm{d}} \mathscr{H}^2 z
\quad \text{for $k \in \mathscr{K} (U \times \grass {n} 2 )$}
\end{gather*}
from Allard \cite[3.5\,(1b)]{MR0307015} and that there exists a countable
collection $C$ of $2$ dimensional submanifolds of $\mathbf{R}^{n}$ of class $2$
such that $\| V\| (U \without \bigcup C ) = 0$ from
\cite[4.8]{snulmenn.c2}. Notice that
\begin{gather*}
\Tan (M,z) = \Tan^2 ( \| V \|, z )
\quad \text{for $\| V \|$ almost all $z \in U \cap M$}
\end{gather*}
for $M \in C$ by \cite[2.8.18, 2.9.11, 3.2.17]{MR41:1976} and Allard
\cite[3.5\,(2)]{MR0307015}. In~particular, one may construct a sequence of
functions $\tau_i : U \to \Hom ( \mathbf{R}^{n}, \mathbf{R}^{n} )$ of class $1$ such
that the sets $Z_i = U \cap \{ z \with \tau(z) = \tau_i (z) \}$ cover $\| V
\|$ almost all of~$U$. For $i \in \mathscr{P}$, applying \ref{thm:O_o} with
${m}$, $p$, $\omega(r)$, $Z$, $f$, and $q$ replaced by $2$, $1$, $r^2
(1 + \log(1/r))$, $Z_i$, $\| \tau - \tau_i \|^2$, and $\infty$ one infers, for $\|
V \|$ almost all $z \in Z_i$, that
\begin{align*}
& \limsup_{r \to 0+} r^{-4} ( \log (1/r) )^{-1} \tint{ \cball zr }{}
\| \tau(\zeta)-\tau(z)\|^2 \ensuremath{\,\mathrm{d}} \| V \| \zeta
\\
& \qquad = \limsup_{r \to 0+} r^{-4} ( \log (1/r))^{-1} \tint{\cball zr}{}
\| \tau - \tau_i \|^2 \ensuremath{\,\mathrm{d}} \| V \| \in \{ 0, \infty \}.
\end{align*}
Therefore it is sufficient to prove for $\|V\|$ almost all $c$ that
\begin{gather*}
\limsup_{r \to 0+} r^{-4} ( \log (1/r) )^{-1} \tint{\cball cr}{}
\| \tau (z)- \tau(c) \|^2 \ensuremath{\,\mathrm{d}} \| V \| z < \infty.
\end{gather*}
For $\| V \|$ almost all $c \in U$ there exist $Q \in \mathscr{P}$ and $M \in C$
such that
\begin{gather*}
\boldsymbol{\Theta}^2 ( \| V \|, c ) = Q,
\quad
\boldsymbol{\Theta}^{\ast 2} ( \| \delta V \|, c ) < \infty,
\quad
\boldsymbol{\Theta}^2 ( \| V \| \mathop{\llcorner} U \without M, c) = 0,
\\
\lim_{r \to 0+} r^{-2}
\tint{\cball cr}{} \| \tau (z) - \tau (c) \| \ensuremath{\,\mathrm{d}} \| V \| z = 0
\end{gather*}
by Allard \cite[3.5\,(1c)]{MR0307015} and \cite[2.8.18, 2.9.5, 2.9.9,
2.9.11]{MR41:1976}. Considering such $c$, $Q$ and~$M$ and abbreviating $T =
\project{\Tan(M,c)}$, it follows
\begin{gather*}
\tau (c) = \project T,
\quad
\limsup_{s \to 0+} s^{-6}
\tint{ \cball cs \cap M}{} \dist (z-c,T)^2 \ensuremath{\,\mathrm{d}} \| V \| z < \infty
\end{gather*}
since $M$ is a submanifold of class $2$. Defining
\begin{gather*}
\eta = \sup \{ 1, \Gamma_{\ref{lemma:app-coercive}} ( {n}, Q ) \},
\quad
\varepsilon = 2^{-14} \eta^{-1},
\end{gather*}
one obtains the existence of $0 < r \leq 1/4$ and $1 \leq \xi < \infty$
such that $\oball c {8r} \subset U$ and for $0 < s \le r$
\begin{gather*}
s^{-1} \measureball{\| \delta V \|}{ \oball c{8s} } + s^{-2}
\big (
\tint{ \oball c{8s} \cap M}{} \dist (z-c,T)^2 \ensuremath{\,\mathrm{d}} \| V \| z
\big )^{1/2} \leq \xi s,
\\
\text{and $V$ satisfies the hypotheses of \ref{lemma:app-coercive}}
\\
\text{ with $r$ and $Z$ replaced by $s$ and $\oball c{8s} \cap M$.}
\end{gather*}
Abbreviating
\begin{gather*}
f(s) = s^{-2} \tint{ \oball cs \times \grass {n} 2}{}
\| \project S - \project T \|^2 \ensuremath{\,\mathrm{d}} V(z,S)
\quad \text{for $0 < s \leq 8r$},
\\
\Delta = \sup \big \{
2^{20} \eta^2 \varepsilon^{-1} \xi^2,
2^6 f (8r) r^{-2} ( \log (1/(8r)))^{-1}
\big \},
\end{gather*}
one inductively proves that
\begin{gather*}
f (s) \leq \Delta s^2 \log (1/s)
\quad \text{whenever $0 < s \leq 8r$};
\end{gather*}
in fact, the inequality is evident if $r \leq s \leq 8r$ and if it holds
with $s$ replaced by~$8s$ for some $0 < s \leq r$, then, recalling $r \le
1/4$, one notes that
\begin{gather*}
s^2 \leq \xi s \big ( \xi s + 2^6 \Delta^{1/2} s ( \log (1/s ) )^{1/2}
\big ) \leq 8^{-1} \eta^{-1} \Delta s^2(\log (1/s))^{1/2}, \\
1 + ( \log ( 1 + 1/s^2 ) )^{1/2} \leq 4 ( \log (1/s) )^{1/2},
\quad
\eta \varepsilon^{-1} \xi^2 s^2 \le 4^{-1} \Delta s^2 \log(1/s),
\end{gather*}
to infer from \ref{lemma:app-coercive} that
\begin{align*}
f(s) & \leq \eta \Big ( \kappa \big ( \xi s ( \xi s + 2^6
\Delta^{1/2} s ( \log (1/s) )^{1/2} ) \big ) + 2^{12} \varepsilon
\Delta s^2 \log (1/s) + \varepsilon^{-1} \xi^2 s^2 \Big ) \\
& \leq 8^{-1} \Delta s^2 ( \log ( 1/s ) )^{1/2}
\big ( 1 + ( \log (1+1/s^2))^{1/2} \big ) + 2^{-1} \Delta s^2 \log (1/s)
\\
& \leq \Delta s^2 \log (1/s),
\end{align*}
where $\kappa$ is as in \ref{lemma:app-coercive}.
\end{proof}
\begin{remark}
In view of \ref{example:quadratic_tilt_excess} the decay rate is sharp for
integral varifolds. For curvature varifolds a stronger conclusion is
attainable, see \cite[15.9]{snulmenn:tv.v2}.
\end{remark}
\begin{remark}
It is an open problem whether the integrality hypothesis on $V$ could be
replaced by the requirement ``$\boldsymbol{\Theta}^{m} ( \| V \|, z ) \geq 1$ for
$\| V \|$ almost all $z$''.
\end{remark}
\section{Quadratic tilt-excess, an example}
\label{sec:qte-ex}
The purpose of this section is to exhibit a two dimensional integral
varifold whose first variation is representable by integration in
order to render the decay estimates of
Section~\ref{sec:quadratic_tilt_decay} sharp. The varifold constructed
for that purpose in~\ref{example:quadratic_tilt_excess} will in fact
be associated to the graph of a Lipschitzian function with small
Lipschitz constant.
The basic building block of this example is a varifold consisting of a~piece of
a~sphere, a~piece of a~catenoid and a~smooth join to a~plane,
see~\ref{lemma:ball_catenoid}. Suitably rescaled copies of~this varifold will
then be used to fill the holes of a~set previously constructed
in~\ref{example:yet_another_cantor_set}.
\begin{figure}[!htb]
\centering
\includegraphics[width=\textwidth, keepaspectratio=true]{ex-qte.mps}
\caption{Rotating the solid line around the vertical axis illustrates the
support of the varifold constructed in \ref{lemma:ball_catenoid}.}
\label{F:catenoid}
\end{figure}
\begin{miniremark} [see \protect{\cite[5.1.9]{MR41:1976}}]
\label{miniremark:pqT}
Suppose $1 < {n} \in \mathscr{P}$ and $\mathbf{p} : \mathbf{R}^{n} \to
\mathbf{R}^{{n}-1}$ and $\mathbf{q} : \mathbf{R}^{n} \to \mathbf{R}$ satisfy
\begin{gather*}
\mathbf{p} (z) = (z_1,\ldots,z_{{n}-1}) \quad \text{and} \quad \mathbf{q} (z) =
z_{n} \quad \text{whenever $z = (z_1,\ldots,z_{n}) \in
\mathbf{R}^{n}$}.
\end{gather*}
Then the statements of
Allard \cite[8.9]{MR0307015} may be supplemented as follows.
\begin{enumerate}
\item \label{item:pqT:norm_equiv} If $S, T \in \grass {n}{{n}-1}$,
then $| \project S - \project T | = 2^{1/2} \| \project S - \project T
\|$.
\item \label{item:pqT:graphs} If $L \in \Hom ( \mathbf{R}^{{n}-1}, \mathbf{R}
)$, $S = \mathbf{R}^{n} \cap \{ z \with L ( \mathbf{p} ( z )) = \mathbf{q} (z) \}$, and
$T = \im \mathbf{p}^\ast$, then $| L | = \| L \|$ and $\| \project S -
\project T \| = ( 1 + \| L \|^2 )^{-1/2} \| L \|$.
\end{enumerate}
In fact, if $v \in \mathbf{S}^{{n}-1}$ and with $S = \{ z \with z
\bullet v = 0 \}$ then \eqref{item:pqT:norm_equiv} is implied by
\begin{gather*}
2^{-1} | \project S - \project T |^2 = \project T \bullet \perpproject
S = | \project T (v) |^2 = \| \project{T} \circ \perpproject{S} \|^2 =
\| \project{S} - \project{T} \|^2
\end{gather*}
and in case of \eqref{item:pqT:graphs} one may take $v = ( 1 + \| L \|^2
)^{-1/2} ( \mathbf{p}^\ast ( L^\ast (1)) - \mathbf{q}^\ast (1) )$.
\end{miniremark}
\begin{miniremark}
\label{miniremark:radial_function}
If $1 < {n} \in \mathscr{P}$, $I$ is an open subset of $\{ t \with 0 < t <
\infty \}$, and $g : I \to \mathbf{R}$ is of class $2$, then $N = \mathbf{R}^{n}
\cap \{ z \with \mathbf{q} (z) = g ( |\mathbf{p} (z)| ) \}$, see \ref{miniremark:pqT},
is an ${n}-1$ dimensional submanifold of $\mathbf{R}^{n}$ of class $2$ and
if $z \in N$ and $t = | \mathbf{p} (z) |$ then
\begin{align*}
| \mathbf{h} (N,z) | & = ( 1 + g'(t)^2 )^{-1/2} \big | ({n}-2)
t^{-1} g' ( t ) + (1+g'(t)^2)^{-1} g'' ( t ) \big | \\
& \leq ({n}-2) t^{-1} |g'(t)| + |g''(t)|, \\
\| \mathbf{b} (N,z) \| & = ( 1 + g'(t)^2 )^{-1/2} \sup \{ t^{-1}
|g'(t)|, (1+g'(t)^2 )^{-1} |g''(t)| \} \\
& \leq \sup \{ t^{-1} |g'(t)|, | g''(t)| \}
\end{align*}
as may be verified using the formulae occurring in
\cite[p.~356--357,~388--391]{MR1814364}.
\end{miniremark}
\begin{miniremark} \label{miniremark:arcosh}
We will employ the area cosinus hyperbolicus, $\ach : \{ t \with 1 \leq
t < \infty \} \to \mathbf{R}$, given by $\ach (t) = \log \big ( t +
(t^2-1)^{1/2} \big )$ for $1 \leq t < \infty$. Notice that
\begin{gather*}
\ach' (t) = (t^2-1)^{-1/2}, \quad \ach''(t) = -t (t^2-1)^{-3/2}
\end{gather*}
for $1 < t < \infty$, hence for $3/2 \leq t < \infty$ also that
\begin{gather*}
\log t \leq \ach(t) \leq 3 \log t, \quad 1/t \leq \ach'(t) \leq 3/t,
\quad - 3/t^2 \leq \ach''(t) < 0.
\end{gather*}
Moreover, $\ach ( t ) - \ach ( s ) \geq \log ( t/s)$ for $1 \leq s \leq t
< \infty$.
\end{miniremark}
\begin{lemma} \label{miniremark:catenoid}
Suppose $N = \mathbf{R}^3 \cap \{ z \with | \mathbf{q} (z) | = \ach ( | \mathbf{p} (z) | )
\}$, see \ref{miniremark:pqT} and \ref{miniremark:arcosh}.
Then $N$ is a $2$ dimensional submanifold of $\mathbf{R}^3$ of class $\infty$,
$\mathbf{h} (N, z ) = 0$ for $z \in N$, and whenever $1 \leq r < \infty$
there holds
\begin{gather*}
\mathscr{H}^2 \big (N \cap \mathbf{p}^{-1}[ \cball 0r ] \big ) = 2
\unitmeasure{2} \big (\ach(r) + r (r^2 - 1)^{1/2} \big ), \\
\tint{N \cap \mathbf{p}^{-1}[ \cball 0r ]}{} | \project{\Tan(N,z)} -
\mathbf{p}^\ast \circ \mathbf{p} |^2 \ensuremath{\,\mathrm{d}} \mathscr{H}^2 z = 8 \unitmeasure{2} \ach(r) .
\end{gather*}
\end{lemma}
\begin{proof}
The asserted equations are readily verified by means of
\ref{miniremark:pqT}--\ref{miniremark:arcosh}.
\end{proof}
\begin{remark}
The surface $N$ is known as the catenoid, see e.g.~\cite[p.~18]{MR852409}.
\end{remark}
\begin{lemma}
\label{lemma:ball_catenoid}
Suppose ${n} = 3$, $\mathbf{p}$ and $\mathbf{q}$ are related to ${n}$ as in
\ref{miniremark:pqT}, and $2 \leq s \leq r/2 < \infty$.
Then there exist $h : \mathbf{R}^2 \to \mathbf{R}$ of class $1$ and $V \in \IVar_2 (
\mathbf{R}^3 )$ satisfying
\begin{gather*}
\Lip h \leq \Gamma/s, \quad \Lip \Der h< \infty, \quad \| V \| = \mathscr{H}^2
\mathop{\llcorner} \im ( \mathbf{p}^\ast + \mathbf{q}^\ast \circ h ), \\
\| V \| \mathop{\llcorner} \mathbf{p}^{-1} [ \mathbf{R}^2 \without \oball 0r ] = \mathscr{H}^2
\mathop{\llcorner} \mathbf{p}^\ast [ \mathbf{R}^2 \without \oball 0r ], \\
1 \leq \unitmeasure{2}^{-1} r^{-2} \| V \| \big ( \mathbf{p}^{-1} [
\oball{0}{r} ] \big ) \leq 1 + \Gamma r^{-2} \log r, \quad \| \delta
V \| ( \mathbf{R}^3 ) \leq \Gamma, \\
\text{$\| \delta V \|$ is absolutely continuous with respect to $\| V
\|$}, \\
\text{$\boldsymbol{\Theta}^2 ( \| V \|,z ) = 1$ and $| \mathbf{q} (z)| \leq 3 \log r$
whenever $z \in \spt \| V \|$}, \\
\tint{\mathbf{p}^{-1} [ \oball{0}{r} ] \times \grass{3}{2}}{} |
\project{S} - \mathbf{p}^\ast \circ \mathbf{p} |^2 \ensuremath{\,\mathrm{d}} V (z,S) \geq \log (r/(2s)),
\end{gather*}
where $\Gamma$ is a universal, positive, finite number.
\end{lemma}
\begin{proof}
Abbreviate $f_1 = \ach$, see \ref{miniremark:arcosh}. Define $d : \{
\sigma \with 1 < \sigma < \infty \} \to \mathbf{R}$ by
\begin{gather*}
d(\sigma) = f_1(\sigma) + \sigma/f_1'(\sigma)
\end{gather*}
for $1 < \sigma < \infty$ and note that $d(\sigma)-\sigma^2$ is a
nondecreasing as a function of $\sigma$ with $d(\sigma)-\sigma^2 \to -1$
as $\sigma \to 1+$, hence
\begin{gather*}
d(\sigma)-\sigma^2 \geq -1 \quad \text{for $1 < \sigma < \infty$}.
\end{gather*}
Define $f_2 : \{ t \with -s^2 < t < s^2 \} \to \mathbf{R}$ by
\begin{gather*}
f_2 (t) = a(s)- \big ( s^4 - t^2 \big )^{1/2} \quad
\text{for $-s^2 < t < s^2$},
\end{gather*}
hence $f_2 (t) \geq -1$ for $-s^2 < t < s^2$. Choose $\gamma \in
\mathscr{E} ( \mathbf{R}, \mathbf{R} )$ with $0 \leq \gamma \leq 1$ and
\begin{gather*}
\gamma (t) = 1 \quad \text{if $t \leq 1/2$}, \qquad \gamma(t) = 0
\quad \text{if $t \geq 1$}, \qquad -3 \leq \gamma'(t) \leq 0
\end{gather*}
whenever $t \in \mathbf{R}$. Noting
\begin{gather*}
s^4-s^2 = s^2/f_1'(s)^2 > 0, \quad f_2(s) = f_1(s), \quad
f_2'(s) = f_1'(s),
\end{gather*}
one defines a function $g : \mathbf{R} \to \mathbf{R}$ of class $1$ with $\Lip g'
< \infty$ by
\begin{gather*}
g (t) = f_2 (|t|) - f_1 (r) \quad \text{if $|t| \leq s$},
\qquad g (t) = ( f_1(|t|) - f_1(r) ) \gamma ( |t|/r ) \quad
\text{else}
\end{gather*}
whenever $t \in \mathbf{R}$.
One computes
\begin{gather*}
\begin{aligned}
g'(t) & = f_1'(t) \gamma(t/r) + (f_1(t)-f_1(r)) r^{-1}
\gamma'(t/r), \\
g''(t) & = f_1''(t) \gamma(t/r) + 2 f_1'(t) r^{-1}
\gamma'(t/r) + (f_1(t)-f_1(r)) r^{-2} \gamma''(t/r)
\end{aligned}
\end{gather*}
for $s < t < \infty$, hence, taking into account \ref{miniremark:arcosh},
there exists a positive, finite number $\Delta_1$ determined by $\gamma$
such that
\begin{gather*}
- \Delta_1 \log r \leq g(t) \leq 0
\quad \text{for $0 \leq t < \infty$},
\\
0 \leq g'(t) \leq \Delta_1 \inf \{ 1/s, 1/t \}
\quad \text{for $0 \leq t < r$},
\\
g'(t) \geq 1 / t
\quad \text{for $s \leq t \leq r/2$},
\qquad |g''(t)| \leq \Delta_1 r^{-2}
\quad \text{for $r/2 < t < r$}.
\end{gather*}
Define $h : \mathbf{R}^2 \to \mathbf{R}$ by
\begin{gather*}
h(x) = g(|x|) \quad \text{for $x \in \mathbf{R}^2$}
\end{gather*}
and note that $h$ is of class $1$ with
\begin{gather*}
\Lip h \leq \Delta_1/s, \quad \Lip \Der h < \infty.
\end{gather*}
Let $M = \mathbf{R}^3 \cap \{ z \with \mathbf{q}(z) = h(\mathbf{p}(z)) \}$, define $V \in
\IVar_2 ( \mathbf{R}^3 )$ by
\begin{gather*}
V(k) = \tint{M}{} k(z,\Tan(M,z)) \ensuremath{\,\mathrm{d}} \mathscr{H}^2 z \quad \text{for $k \in
\mathscr{K} ( \mathbf{R}^3 \times \grass{3}{2} )$}
\end{gather*}
and observe
\begin{gather*}
\delta V(\theta) = - \tint{M}{} \mathbf{h}(M,z) \bullet \theta(z) \ensuremath{\,\mathrm{d}}
\mathscr{H}^2 z \quad \text{for $\theta \in \mathscr{D} (\mathbf{R}^3, \mathbf{R}^3 )$}.
\end{gather*}
Since $M \cap \mathbf{p}^{-1} [ \oball{0}{s} ]$ is a~piece of a~sphere of
radius~$s^2$ one computes
\begin{gather*}
| \mathbf{h}(M,z) | = 2 s^{-2}
\quad \text{whenever $z \in M \cap \mathbf{p}^{-1} [ \oball{0}{s} ]$}, \\
\tint{M \cap \mathbf{p}^{-1} [ \oball{0}{s} ]}{} |
\mathbf{h}(M,z) | \ensuremath{\,\mathrm{d}} \mathscr{H}^2 z
= 4 \unitmeasure{2} s / (s + (s^2 - 1)^{1/2}) \leq 4 \unitmeasure 2
\end{gather*}
and since $M \cap \mathbf{p}^{-1} [ \cball{0}{r/2} \without \cball{0}{s} ]$
is a~piece of a~catenoid one obtains
\begin{gather*}
\mathbf{h}(M,z) = 0
\quad
\text{whenever $z \in M \cap \mathbf{p}^{-1} [ \cball{0}{r/2} \without
\cball{0}{s} ]$}
\end{gather*}
from \ref{miniremark:catenoid}. Moreover, recalling $2 \le s \le r/2$, one
estimates
\begin{gather*}
\begin{aligned}
0 \leq \mathscr{H}^2 \big ( M \cap \mathbf{p}^{-1} [ \oball{0}{r} ] \big )
- \unitmeasure{2} r^2 & = 2 \unitmeasure 2 \tint{0}{r} \big ( ( 1
+ g'(t)^2)^{1/2} -1 \big ) t \ensuremath{\,\mathrm{d}} \mathscr{L}^1 t \\
& \leq \unitmeasure 2 \tint{0}{r} g'(t)^2 t \ensuremath{\,\mathrm{d}} \mathscr{L}^1 t
\leq 2 \unitmeasure{2} \Delta_1^2 \log r.
\end{aligned}
\end{gather*}
From \ref{miniremark:radial_function} and the estimates for $g'$ and
$g''$, one obtains a positive, finite number $\Delta_2$ determined by
$\gamma$ such that
\begin{gather*}
| \mathbf{h}(M,z) | \leq \Delta_2 r^{-2}
\quad
\text{whenever $z \in M \cap \mathbf{p}^{-1} [ \oball{0}{r} \without
\cball{0}{r/2} ]$}.
\end{gather*}
Combining the preceding estimates, we obtain, recalling $r \geq 2$, that
\begin{gather*}
\| \delta V \| ( \mathbf{R}^3 ) \leq 4 \unitmeasure 2 + \unitmeasure{2}
\Delta_2(1 + \Delta_1^2) .
\end{gather*}
Since $M \cap \mathbf{p}^{-1}[ \oball{0}{r/2} \without \cball{0}{s} ]$ is
a~piece of a~catenoid, \ref{miniremark:catenoid} implies
\begin{gather*}
\tint{\mathbf{p}^{-1} [ \oball{0}{r} ] \times \grass{3}{2}}{} |
\project{S} - \mathbf{p}^\ast \circ \mathbf{p} |^2 \ensuremath{\,\mathrm{d}} V (z,S) \geq 8
\unitmeasure 2 ( \ach ( r/2) - \ach (s) ),
\end{gather*}
hence \ref{miniremark:arcosh} implies the conclusion.
\end{proof}
\begin{miniremark}
\label{miniremark:open_cube}
Occasionally, we denote the open cube with centre~$a$ and side length~$2r$
by
\begin{gather*}
\ocube ar = \mathbf{R}^{m} \cap \{ (x_1,\ldots,x_{m}) \with |x_i - a_i|
< r \text{ for $i = 1,\ldots,{m}$}\}
\end{gather*}
for ${m} \in \mathscr{P}$, $a = (a_1,\ldots,a_{m}) \in \mathbf{R}^{m}$, and $0 < r
< \infty$.
\end{miniremark}
\begin{example}
\label{example:quadratic_tilt_excess}
Suppose ${n} = 3$, $\mathbf{p}$ and $\mathbf{q}$ are related to ${n}$ as in
\ref{miniremark:pqT}, $T = \im \mathbf{p}^\ast$, $\varepsilon > 0$, and $\omega$
is a modulus of~continuity.
Then there exist $f : \mathbf{R}^2 \to \mathbf{R}$, $C \subset T$, and $V \in \IVar_2
( \mathbf{R}^3 )$ satisfying
\begin{gather*}
\Lip f \leq \varepsilon, \quad \| V \| = \mathscr{H}^2
\mathop{\llcorner} \im ( \mathbf{p}^\ast + \mathbf{q}^\ast \circ f ), \quad \| V \|
( C ) > 0, \\
\| \delta V \| ( \mathbf{R}^3 ) < \infty, \quad \text{$\| \delta V
\|$ is absolutely continuous with respect to $\| V \|$}, \\
\limsup_{r \to 0+} r^{-1} ( \log (1/r) )^{-1/2}
\omega(r)^{-1} \big ( \tfint{\cball{c}{r} \times
\grass{3}{2}}{} | \project{S} - \project{T} |^2 \ensuremath{\,\mathrm{d}} V (z,S)
\big )^{1/2} > 0
\end{gather*}
whenever $c \in C$, here $0^{-1} = \infty$.
\end{example}
\begin{proof} [Construction]
Take $G$ and $A$ as furnished by \ref{example:yet_another_cantor_set} with
${m}$ and $\lambda$ replaced by $2$ and $1/2$, abbreviate $\Delta =
\Gamma_{\ref{lemma:ball_catenoid}}$, $\lambda = \inf \{ 1/4, \varepsilon
/(2\Delta) \}$, and let $C = \mathbf{p}^\ast [ A ]$. Define $W \in \IVar_2
( \mathbf{R}^3 )$ by
\begin{gather*}
W(k) = \tint{T \without \mathbf{p}^\ast [ \bigcup G ]}{} k (z,T) \ensuremath{\,\mathrm{d}}
\mathscr{H}^2 z \quad \text{for $k \in \mathscr{K} ( \mathbf{R}^3 \times
\grass{3}{2} )$}.
\end{gather*}
Whenever $Q = \ocube at \in G$ and $t > \lambda$ let $f_Q : \mathbf{R}^2 \to \mathbf{R}$
and $X_Q \in \IVar_2 ( \mathbf{R}^3 )$ be defined by
\begin{gather*}
f_Q (x) = 0 \quad \text{for $x \in \mathbf{R}^2$}, \qquad \| X_Q \|
= \mathscr{H}^2 \mathop{\llcorner} \im ( \mathbf{p}^\ast + \mathbf{q}^\ast \circ f_Q
).
\end{gather*}
Whenever $Q = \ocube{a}{t} \in G$ and $t \leq \lambda$ apply
\ref{lemma:ball_catenoid} with $s$ and $r$ replaced by $(2\lambda)^{-1}$
and $1/t$ to construct $f_Q : \mathbf{R}^2 \to \mathbf{R}$ and $X_Q \in \IVar_2 (
\mathbf{R}^3 )$ such that $t^{-2} f_Q \circ \boldsymbol{\tau}_a \circ
\boldsymbol{\mu}_{t^2}$ and $( \boldsymbol{\mu}_{t^{-2}} \circ
\boldsymbol{\tau}_{-a} )_\# X_Q$ satisfy the conditions of
\ref{lemma:ball_catenoid} in place of $h$ and $V$ implying
\begin{gather*}
\spt f_Q \subset \Clos Q ,
\quad
\Lip f_Q \leq \varepsilon,
\quad
\| X_Q \| = \mathscr{H}^2 \mathop{\llcorner} \im ( \mathbf{p}^\ast + \mathbf{q}^\ast \circ f_Q ),
\\
\| X_Q \| \mathop{\llcorner} \mathbf{p}^{-1} [ \mathbf{R}^2 \without Q ] = \mathscr{H}^2
\mathop{\llcorner} \mathbf{p}^\ast [ \mathbf{R}^2 \without Q ], \\
\| \delta X_Q \| ( \mathbf{R}^3 ) \leq \Delta t^2,
\quad
\text{$\| \delta X_Q \|$ is absolutely continuous with respect to $\|
X_Q \|$}, \\
| \mathbf{q} (z) | \leq 3 t^2 \log ( 1/t ) \quad \text{whenever $z \in \spt
\| X_Q \|$}, \\
\tint{\mathbf{p}^{-1} [ Q ] \times \grass{3}{2}}{}
| \project{S} - \project{T} |^2 \ensuremath{\,\mathrm{d}} V (z,S) \geq t^4 \big ( \log (1/t)
- \log (1/\lambda) \big ).
\end{gather*}
Recall that $G$ is disjointed and define $f : \mathbf{R}^2 \to \mathbf{R}$ and $V \in
\IVar_2 ( \mathbf{R}^3 )$ by
\begin{gather*}
f(x) = \sum_{Q \in G} f_Q (x) \quad \text{for $x \in \mathbf{R}^2$},
\qquad \| V \| = \mathscr{H}^2 \mathop{\llcorner} \im ( \mathbf{p}^\ast + \mathbf{q}^\ast \circ f ).
\end{gather*}
Note that $V = W + \sum_{Q \in G} X_Q \mathop{\llcorner} ( \mathbf{p}^{-1} [ Q ]
\times \grass{3}{2} )$ and $\| V \| ( C ) \geq 1/2$. Observe
\begin{gather*}
\| \delta V \| ( \mathbf{R}^3 ) < \infty, \quad \text{$\| \delta V
\|$ is absolutely continuous with respect to $ \| V \|$}.
\end{gather*}
Suppose $c \in C$ and $\delta > 0$.
Then there exist $r$ and $\ocube{a}{t} = Q$ such that
\begin{gather*}
0 < r \leq \delta, \quad Q \in G, \quad Q \subset
\oball{a}{r}, \quad \mathscr{L}^2 ( Q ) \geq \omega (r) r^2 ,
\end{gather*}
hence $t \leq r$. Since $(2t)^2 \geq \omega (r) r^2$ and
\begin{gather*}
t^4 \big ( \log (1/t) - \log (1/\lambda) \big ) \geq 2^{-4} \omega
(r)^2 r^4 \big ( \log (1/r) - \log (1/\lambda) \big ),
\end{gather*}
the estimates for $X_Q$ imply the assertion.
\end{proof}
\section{Super-quadratic tilt-excess, decay rates}
\label{sec:sqte-decay}
The present section concerns integral varifolds of at least two dimensions,
deferring the one dimensional case to Section~\ref{sec:one_dimensional_decay}.
Its purpose is to complement the examples concerning the decay rates of the
super-quadratic tilt-excess constructed in \ref{example:quantitative_brakke},
\ref{example:quantitative_brakke_again} and \cite[\S 1]{snulmenn.isoperimetric}
by positive results, see \ref{thm:positive_result}. This yields a~sharp
dividing line in most cases, see
\ref{remark:sharpness_superquadratric}--\ref{remark:m=2,p=1}. Additionally, we
prove that the examples constructed in \ref{example:quantitative_brakke} and
\ref{example:quantitative_brakke_again} are essentially sharp also with respect
to the size of holes the varifolds contain, see \ref{corollary:holes}.
The positive results follow readily from the existing theory. For the
super-quadratic tilt-excess, these are the second order rectifiability and its
consequences for the decay of the quadratic tilt-excess in conjunction with the
differentiation theory both obtained in~\cite[4.8, 5.2]{snulmenn.c2} and
\cite[\S 3]{snulmenn.isoperimetric} respectively. Concerning the estimate for
the size of the holes, we additionally employ an approximation by $\mathbf{Q}_Q (
\mathbf{R}^{n-m} )$ valued functions, see~\cite[3.15]{snulmenn.poincare}, and more
basic results on the size of the set where the first variation is large
from~\cite[\S 2]{snulmenn.isoperimetric}.
\begin{theorem}
\label{thm:positive_result}
Suppose ${m}$, ${n}$, $p$, $U$, and $V$ satisfy the hypotheses of
\ref{miniremark:situation_general}, $V \in \IVar_{{m}}(U)$, $2 < q <
\infty$, and either
\begin{enumerate}
\item ${m} = 2$ and $p>1$, or
\item ${m}>2$ and $p \geq 2{m}/({m}+2)$.
\end{enumerate}
Then for $V$ almost all $(z,T)$ there holds
\begin{gather*}
\lim_{r \to 0+} r^{-{m}-2} \tint{\cball{z}{r} \times
\grass{{n}}{{m}}}{} | \project{S} - \project{T} |^q \ensuremath{\,\mathrm{d}} V
(\zeta,S) = 0.
\end{gather*}
\end{theorem}
\begin{proof}
Assume ${m} < {n}$. First, note that since the function mapping $S \in
\grass{{n}}{{m}}$ to $|\project S - \project T|$ is bounded for any $T
\in \grass{{n}}{{m}}$, we have
\begin{gather*}
\limsup_{r \to 0+} r^{-{m}-2} \tint{\cball{z}{r} \times
\grass{{n}}{{m}}}{} | \project{S} - \project{T} |^q \ensuremath{\,\mathrm{d}} V
(\zeta,S) < \infty
\end{gather*}
for $V$ almost all $(z,T)$ by \cite[5.2\,(2)]{snulmenn.c2} and H\"older's
inequality. Second, note that \cite[4.8]{snulmenn.c2} implies the
existence of a sequence of functions $\tau_i : U \to \Hom ( \mathbf{R}^{n},
\mathbf{R}^{n} )$ of class $\class{1}$ such that
\begin{gather*}
\| V \| \big ( U \without {\textstyle\bigcup_{i=1}^{\infty}}
Z_i \big ) = 0,
\end{gather*}
where $Z_i = \classification{U}{z}{ \tau_i(z) = \project{\Tan^{m} ( \| V
\|, z )} }$, hence
\begin{gather*}
\lim_{r \to 0+} r^{-{m}-2} \tint{\cball{z}{r}}{} | \tau_i (\zeta) -
\project{\Tan^{m} ( \| V \|, \zeta )} |^q \ensuremath{\,\mathrm{d}} \| V \|\zeta = 0
\end{gather*}
for $\| V \|$ almost all $z \in Z_i$
by \cite[3.7\,(i)]{snulmenn.isoperimetric} with $Z$,
$f$, $\alpha$, $r$, and $g$ replaced by $\Hom ( \mathbf{R}^{n},
\mathbf{R}^{n})$, $\project{\Tan^{m} ( \| V \|, \cdot )}$, $2/q$, $\infty$,
and $\tau_i$. The conclusion then follows, since the functions $\tau_i$
are of class~$1$.
\end{proof}
\begin{remark}
The concept of proof is the same as in \cite[5.2\,(1)]{snulmenn.c2}.
\end{remark}
\begin{remark}
\label{remark:sharpness_superquadratric}
Note that the number $2$ in $r^{-{m}-2}$ cannot be replaced by any larger
number by \ref{example:quantitative_brakke} even if ${n} = {m}+1$ and
``$\lim$'' is replaced by ``$\liminf$''.
\end{remark}
\begin{remark}
Note the following proposition: \emph{If ${m}$, ${n}$, $p$, $U$, and
$V$ are as in \ref{miniremark:situation_general}, $V \in
\IVar_{{m}}(U)$, ${m} > 2$, $p < 2{m}/({m}+2)$, and $2 \leq q <
\infty$, then
\begin{gather*}
\lim_{r \to 0+} r^{-{m}-{m} p/({m}-p)} \tint{\cball{z}{r}
\times \grass{{n}}{{m}}}{} | \project{S} - \project{T} |^q \ensuremath{\,\mathrm{d}} V
(z,S) = 0
\end{gather*}
for $V$ almost all $(z,T)$;} in fact, it suffices to combine
\cite[5.2\,(1)]{snulmenn.c2} with H\"older's inequality. Taking $\alpha_1 =
\alpha_2$ slightly larger than $q^{-1} {m} p({m}-p)^{-1}$ in
\cite[1.2]{snulmenn.isoperimetric}, one infers that ${m} p/({m}-p)$ cannot
be replaced by any larger number in the preceding statement even if ${n} =
{m}+1$ and ``$\lim$'' is replaced by ``$\liminf$''.
\end{remark}
\begin{remark}
\label{remark:m=2,p=1}
Note the following proposition: \emph{If ${m}$, ${n}$, $p$, $U$, and
$V$ are as in \ref{miniremark:situation_general}, $V \in
\IVar_{{m}}(U)$, ${m}=2$, $p=1$, $0 < s < 2$, and $2 \leq q < \infty$,
then
\begin{gather*}
\limsup_{r \to 0+} r^{-2-s} \tint{\cball{z}{r} \times
\grass{{n}}{{m}}}{} | \project{S} - \project{T} |^q \ensuremath{\,\mathrm{d}} V
(\zeta,S) < \infty
\end{gather*}
for $V$ almost all $(z,T)$;} in fact, again, it suffices to combine
\cite[5.2\,(1)]{snulmenn.c2} with H\"older's inequality. Taking $\alpha_1 =
\alpha_2$ slightly larger than $2q^{-1}$ in \cite[1.2,
1.3]{snulmenn.isoperimetric}, one infers that $s$ cannot be replaced by any
number larger than $2$ in the preceding statement and $s$ cannot be replaced
by $2$ in case $q=2$ by \ref{example:quadratic_tilt_excess} both even if
${n} = {m}+1$ and ``$\limsup$'' is replaced by ``$\liminf$''. This leaves
open the case $s=2$ and $q>2$. An affirmative answer to the latter case would
be implied by interpolation if one would know
\begin{gather*}
\limsup_{r \to 0+} r^{-1} \phi (z,r,T) < \infty
\quad
\text{for $V$ almost all $(z,T)$},
\end{gather*}
where $\phi(z,r,T)$ abbreviates
\begin{gather*}
r^{-1} \sup \big\{ t V ( \eqclassification{\cball{z}{r} \times
\grass{{n}}{{m}} }{(\zeta,S)}{| \project{S} - \project{T} | >
t})^{1/2} \with 0 < t < \infty \big \}.
\end{gather*}
\end{remark}
\begin{remark}
\label{remark:no-gehring-improvement}
\emph{If $2 \leq {m} \in \mathscr{P}$, $1 \leq p < \infty$, and $2 < q <
\infty$, then there exist $V \in \IVar_{m} (U)$ related to ${m}$,
${n} = {m}+1$, $p$, and $U = \mathbf{R}^{n}$ as in
\ref{miniremark:situation_general} and $A$ with $V(A)>0$ satisfying
\begin{gather*}
\lim_{r \to 0+} \frac{\big ( r^{-{m}} \tint{\cball zr \times \grass
{n} {m} }{} | \project S - \project T |^q \ensuremath{\,\mathrm{d}} V(\zeta,S) \big
)^{1/q}}{\big ( r^{-{m}} \tint{\cball zr \times \grass {n}
{m}}{} | \project S - \project T|^2 \ensuremath{\,\mathrm{d}} V(\zeta,S) \big )^{1/2} +
r^{1-{m}/p} \psi ( \cball zr)^{1/p}} = \infty
\end{gather*}
whenever $(z,T) \in A$, where $\psi$ is as in
\ref{miniremark:situation_general};} in fact, choosing $\alpha$ such that
$q^{-1} {m} p({m}-p)^{-1} < \alpha < 2^{-1} {m} p ( {m}-p )^{-1}$
if ${m} > 2$ and $p < 2 {m}/({m}+2)$ and $2/q < \alpha < 1$
otherwise, \ref{remark:sharpness_superquadratric}--\ref{remark:m=2,p=1}
yield $V \in \IVar_{m} (U)$ related ${m}$, ${n} = {m}+1$, $p$, and
$U = \mathbf{R}^{n}$ as in \ref{miniremark:situation_general} and $A$ with
$V(A)>0$ such that
\begin{gather*}
\liminf_{r \to 0+} r^{-\alpha} \big ( r^{-{m}} \tint{\cball zr
\times \grass {n} {m}}{} | \project S - \project T|^q \ensuremath{\,\mathrm{d}} V (
\zeta, S ) \big )^{1/q} > 0
\end{gather*}
whenever $(z,T) \in A$, hence \cite[5.2]{snulmenn.c2} and \cite[2.8.18,
2.9.5, 2.9.8]{MR41:1976} imply the assertion. \emph{The same statement
holds for $p = \infty$ if $\psi ( \cball zr )^{1/p}$ is replaced by
$\eqLpnorm{\| V \| \mathop{\llcorner} \cball zr}{\infty}{ \mathbf{h}(V,\cdot)}$.}
\end{remark}
\begin{theorem}
\label{corollary:holes} Suppose ${m}$, ${n}$, $p$, $U$, and $V$
satisfy the hypotheses of \ref{miniremark:situation_general}, $V \in
\IVar_{{m}}(U)$, and either
\begin{enumerate}
\item ${m} = 2$ and $p>1$, or
\item ${m}>2$ and $p \geq 2{m}/({m}+2)$.
\end{enumerate}
Then for $V$ almost all $(c,T)$ there holds
\begin{gather*}
\lim_{r \to 0+} r^{-{m}-2} \mathscr{H}^{m} ( H (T,c,r) ) = 0,
\end{gather*}
where $H (T,c,r) = T \cap \cball{\project{T}(c)}{r} \without \project{T}
[ \cylinder Tcrr \cap \{ z \with \boldsymbol{\Theta}^{\ast {m}} ( \| V \|, z ) >
0 \} ]$.
\end{theorem}
\begin{proof}
Assume $1 < p < {m}$. If ${m} = {n}$, then $\delta V = 0$ by
\cite[4.8]{snulmenn.c2}, hence the conclusion follows from Allard
\cite[4.6\,(3)]{MR0307015}. Therefore assume ${m} < {n}$.
Suppose $Q$ is a positive integer. Recalling
\cite[2.4]{snulmenn.isoperimetric}, define
\begin{gather*}
\lambda = \varepsilon_{\text{\cite[3.15]{snulmenn.poincare}}}
( {n-m}, {m}, Q, 1, 5^{m} Q, 1/4, 1/4, 1/4, 1/4, ( 2
\isoperimetric{{m}} {m})^{-{m}} / \unitmeasure{{m}} ),
\\
Z = U \cap \{ z \with \Tan^{m} ( \| V \|, z ) \in
\grass{{n}}{{m}} \}
\end{gather*}
and $\tau \with Z \to \Hom ( \mathbf{R}^{n}, \mathbf{R}^{n})$ by
\begin{gather*}
\tau (z) = \project{\Tan^{m} ( \| V \|, z )} \quad
\text{whenever $z \in Z$}.
\end{gather*}
Let $B_i$ consist of all $z \in \spt \| V \|$ such that either
$\cball{z}{1/i} \not \subset U$ or
\begin{gather*}
\measureball{\| \delta V \|}{ \cball{z}{s} } > \lambda \,
\| V \| ( \cball{z}{s} )^{1-1/{m}} \quad \text{for some $0 <
s < 1/i$}
\end{gather*}
whenever $i$ is a positive integer. Note that $B_{i+1} \subset B_i$.
Moreover, let $D_i(c)$ denote the set of all $z \in U$ such that
either $\cball{z}{1/i} \not \subset U$ or
\begin{gather*}
\tint{\cball{z}{s}}{} | \tau(\zeta) - \tau(c)| \ensuremath{\,\mathrm{d}} \| V \| \zeta >
\lambda \, \measureball{\| V \|}{\cball{z}{s}} \quad
\text{for some $0 < s < 1/i$}
\end{gather*}
whenever $c \in Z$ and $i$ is a positive integer. Note that $D_{i+1}
(c) \subset D_i(c)$.
Next, the following assertion will be proven. \emph{For $\| V \|$ almost
all $c$ there exists $i$ such that
\begin{gather*}
\lim_{r \to 0+} r^{-{m}-2} \| V \| ( B_i \cap \cball{c}{r} ) = 0,
\quad \lim_{r \to 0+} r^{-{m}-2} \| V \| ( D_i(c) \cap
\cball{c}{r} ) = 0.
\end{gather*}}
Noting ${m} p / ({m}-p) \geq 2$ and applying \cite[2.9,
2.10]{snulmenn.isoperimetric} with $m$, $n$, $\mu$, $s$, $\varepsilon$,
and $\Gamma$ replaced by ${n-m}$, ${m}$, $\| V \|$, ${m}$, $\inf \big
\{ ( 2 \isoperimetric{{m}} )^{-p/({m}-p)}, \lambda^{p/({m}-p)}
\big \}$, and $8 \isoperimetric{{m}} {m}$ yields the first equality.
In view of \cite[5.2\,(2)]{snulmenn.c2}, applying
\cite[3.7\,(ii)]{snulmenn.isoperimetric} with $n$, $m$, $\mu$, $Z$, $f$,
$\alpha$, $q$, and $r$ replaced by ${m}$, ${n-m}$, $\| V \|$, $\Hom (
\mathbf{R}^{n}, \mathbf{R}^{n} )$, $\tau$, $1$, $2$, and $\infty$ one obtains the
second equality.
Note that for $V$ almost all $(c,T)$ with density $\boldsymbol{\Theta}^{m} ( \| V
\|, c ) = Q$ the hypotheses of~\cite[3.15]{snulmenn.poincare} (Lipschitz
approximation theorem) with $m$, $n$, $L$, $M$, $\delta_1$, $\delta_2$,
$\delta_3$, $\delta_4$, $\delta_5$, $a$, $h$, and $\mu$ replaced by
${n-m}$, ${n}$, $1$, $5^{m} Q$, $1/4$, $1/4$, $1/4$, $1/4$, $(2
\isoperimetric{{m}} {m} )^{-{m}} / \unitmeasure{{m}}$, $c$, $r$,
and $\| V \|$ are satisfied for all sufficiently small $r>0$. Therefore
the conclusions (1)--(3) of \cite[3.15]{snulmenn.poincare} with
$\varepsilon_1$ replaced by $\lambda$ in conjunction with the assertion of
the preceding paragraph yield the conclusion.
\end{proof}
\begin{remark}
Possibly up to logarithmic factors, the estimate obtained is sharp even in
case ${n-m} = 1$ and $p = \infty$ by \ref{example:quantitative_brakke} and
\ref{example:quantitative_brakke_again}.
\end{remark}
\section{Super-quadratic tilt-excess, an example}
\label{sec:super_quadratic_tilt}
In this section we provide examples of curvature varifolds satisfying the
conditions of \ref{miniremark:situation_general} with $p = \infty$, hence in
particular having bounded generalised mean curvature vector, for which there is
a~set of positive weight measure such that in arbitrarily small balls around the
points of that set there is a portion of relatively large measure where the tilt
is greater than~$1/3$. In~fact, the Hausdorff measure of the regions in the
affine tangent planes which are not covered by the varifold, i.e., the size of
``holes'', is large at these scales which is essentially a stronger statement,
see~\ref{remark:holes-tilt-large}. The power of the decay of the super-quadratic
tilt-excess exhibited by these varifolds is the smallest possible, see
\ref{remark:example-super-quadratic-tilt} and \ref{thm:positive_result}. In
\ref{example:quantitative_brakke_again}, the example is modified so as to yield
the largest possible size of points of small density permitted by the
approximate lower semicontinuity of the density,
see~\ref{remark:density-ap-lsc}.
The qualitative construction principle was described by Brakke in
\cite[6.1]{MR485012} for two dimensional integral varifolds. Our implementation
employs additionally the estimates obtained in \ref{lemma:bent_catenoid} for
certain varifolds, see Figure \ref{F:bent-catenoid}, and the sets constructed
in~\ref{example:cantor_set} and~\ref{example:yet_another_cantor_set}.
\begin{figure}[!htb]
\centering
\includegraphics[width=\textwidth, keepaspectratio=true]{ex-sqte.mps}
\caption{Rotating the solid line around the vertical axis illustrates the
support of the varifold constructed in \ref{lemma:bent_catenoid}.}
\label{F:bent-catenoid}
\end{figure}
\begin{miniremark}
\label{miniremark:absorb}
If $\phi$ is a measure, $A$ is $\phi$ measurable, $\phi (A) < \infty$, $f
\in \Lp{\infty} ( \phi )$, $\varepsilon > 0$, and $\varepsilon \phi (A) \leq
\tint A{} f \ensuremath{\,\mathrm{d}} \phi$, then
\begin{gather*}
( \varepsilon / 2 ) \phi (A) \leq \phi ( A \cap \{ x \with f(x) \geq
\varepsilon / 2 \} ) \Lpnorm{\phi}{\infty}{f}.
\end{gather*}
\end{miniremark}
\begin{lemma}
\label{lemma:bent_catenoid}
Suppose ${n} = 3$, $\mathbf{p}$ and $\mathbf{q}$ are related to ${n}$ as in
\ref{miniremark:pqT}, and $4 \leq r < \infty$.
Then there exists a~curvature varifold $V \in \IVar_2 ( \mathbf{R}^3 )$
satisfying
\begin{gather*}
\spt \| V \| \subset \im \mathbf{p}^\ast \cup \mathbf{p}^{-1} [ \oball 0r ],
\\
\| V \| \mathop{\llcorner} \mathbf{p}^{-1} [ \mathbf{R}^2 \without \oball{0}{r} ]
= 2 \mathscr{H}^2 \mathop{\llcorner} \im \mathbf{p}^\ast \without \oball{0}{r},
\\
\text{$\mathbf{p}^{-1} [ \oball 0r ] \cap \spt \| V \|$
is a two dimensional submanifold of $\mathbf{R}^3$ of class $\infty$},
\\
\boldsymbol{\Theta}^2 ( \| V \|, z ) = 1
\quad
\text{for $z \in \mathbf{p}^{-1} [ \oball 0r ] \cap \spt \| V \|$},
\\
0 \leq \| V \| \big ( \mathbf{p}^{-1} [ \oball 0r ] \big )
- 2 \measureball{\mathscr{L}^2}{ \oball 0r }
\leq \Gamma ( \log r )^2,
\\
\| \delta V \| \leq \Gamma ( r^{-2} \log r ) \| V \|,
\quad \tint{}{} \| \mathbf{b} (V,z) \| \ensuremath{\,\mathrm{d}} \| V \| z \leq \Gamma \log r,
\\
\text{$| \mathbf{p} (z) | \geq 1$ and $| \mathbf{q} (z) | \leq 3 \log r$}
\quad
\text{whenever $z \in \spt \| V \|$},
\\
V \big( \eqclassification{\mathbf{R}^3 \times \grass{3}{2}}{(z,S)}
{\| \project{S} - \mathbf{p}^\ast \circ \mathbf{p} \| \geq 1/3 } \big) \geq 1,
\end{gather*}
where $\Gamma$ is a universal positive, finite number.
\end{lemma}
\begin{proof}
Choose $\gamma \in \mathscr{D} ( \mathbf{R}, \mathbf{R} )$ with $\{ t \with \gamma (t) >
0 \} = \oball 01$ and
\begin{gather*}
0 \leq \gamma (t) \leq 1 \quad \text{for $t \in \mathbf{R}$},
\qquad
\gamma (t) = 1
\quad
\text{for $-1/2 \leq t \leq 1/2$}.
\end{gather*}
Recalling \ref{miniremark:arcosh}, define $g : \classification{\mathbf{R}}{t}{ 1
< t < \infty} \to \mathbf{R}$ by
\begin{gather*}
g(t) = \ach(t) \gamma (t/r) \quad \text{for $1 < t < \infty$},
\end{gather*}
hence there exists a positive, finite number $\Delta_1$ determined by
$\gamma$ such that
\begin{gather*}
|g'(t)| \leq \Delta_1 r^{-1} \log r \quad \text{and} \quad |g''(t)| \leq
\Delta_1 r^{-2} \log r \qquad \text{for $r/2 \leq t \leq r$}.
\end{gather*}
Defining $h : \mathbf{R}^2 \without \cball{0}{1} \to \mathbf{R}$ by $h(x) = g (|x|)$
for $x \in \mathbf{R}^2 \without \cball{0}{1}$, let
\begin{gather*}
M = \classification{\mathbf{R}^3}{z}{ \mathbf{q} (z) = h ( \mathbf{p} (z) ) }.
\end{gather*}
Notice that $\mathbf{h} ( M,z ) = 0$ for $z \in M \cap \mathbf{p}^{-1} [
\oball 0{r/2} ]$ by \ref{miniremark:catenoid} and
\begin{gather*}
M \cap \mathbf{p}^{-1} [ \mathbf{R}^2 \without \oball 0r ] = \im \mathbf{p}^\ast
\without \oball 0r, \\
\mathbf{p} [ M ] \subset \mathbf{R}^2 \without \oball 01, \quad \mathbf{q} [ M
] \subset \cball 0{3 \log r}, \\
M \cap \mathbf{p}^{-1} [ \cball 02 ] \subset \{ z \with \| \project{\Tan
( M,z)} - \mathbf{p}^\ast \circ \mathbf{p} \| \geq 1/3 \}.
\end{gather*}
by \ref{miniremark:pqT}\,\eqref{item:pqT:graphs} and
\ref{miniremark:arcosh}. Recalling $r \geq 4$, one may deduce from
\ref{miniremark:arcosh} and \ref{miniremark:catenoid} that
\begin{gather*}
0 \leq \mathscr{H}^2 \big ( M \cap \mathbf{p}^{-1} [ \oball{0}{r/2} ] \big ) -
\mathscr{L}^2 ( \oball 0{r/2} ) \leq 12 \log r.
\end{gather*}
Noting that $(1+s)^{1/2} \leq 1 + s/2$ for $-1 \leq s < \infty$, one
estimates
\begin{align*}
0 & \leq \mathscr{H}^2 \big ( M \cap \mathbf{p}^{-1} [ \oball 0r \without \oball
0{r/2} ] \big ) - \mathscr{L}^2 ( \oball 0r \without \oball 0{r/2}
) \\
& \leq 8 \tint{r/2}{r} \big ( ( 1 + g'(t)^2 )^{1/2} - 1 \big ) t \ensuremath{\,\mathrm{d}}
\mathscr{L}^1 t \leq 4 \tint{r/2}{r} g'(t)^2 t \ensuremath{\,\mathrm{d}} \mathscr{L}^1 t \leq
2 \Delta_1^2 ( \log r )^2.
\end{align*}
In view of the estimates for $g'$ and $g''$ and
\ref{miniremark:radial_function}, one notes
\begin{gather*}
\| \mathbf{b} (M,z) \| \leq 2 \Delta_1 r^{-2} \log r \quad
\text{whenever $z \in M$ and $r/2 \leq | \mathbf{p} (z) | \leq r$},
\end{gather*}
hence, using \ref{miniremark:radial_function} and \ref{miniremark:arcosh},
one obtains
\begin{gather*}
\tint{M}{} \| \mathbf{b}(M,z) \| \ensuremath{\,\mathrm{d}} \mathscr{H}^2 z \leq \Delta_2 \log r,
\end{gather*}
where $\Delta_2$ is a positive, finite number determined by $\gamma$.
Employing the reflection $L : \mathbf{R}^3 \to \mathbf{R}^3$ given by $L(z)=
\mathbf{p}^\ast(\mathbf{p}(z)) - \mathbf{q}^\ast(\mathbf{q}(z))$ for $z \in \mathbf{R}^3$, one defines a
curvature varifold
$V \in \IVar_2 ( \mathbf{R}^3 )$ by
\begin{gather*}
V(k) = \tint{M}{} k(z,\Tan(M,z)) + k (L(z), L [ \Tan (M,z) ] ) \ensuremath{\,\mathrm{d}} \mathscr{H}^2 z
\end{gather*}
for $k \in \mathscr{K} ( \mathbf{R}^3 \times \grass{3}{2} )$. Hence one may take
$\Gamma = 4 \sup \{ (3 + \Delta_1)^2, \Delta_2 \}$.
\end{proof}
\begin{example}
\label{example:quantitative_brakke}
Suppose ${m}$ is an integer with ${m} \geq 2$ and $\omega$ is a
modulus of continuity satisfying the Dini condition.
Then there exist $\varepsilon$, $C$, $M$, $R$, $T$, and $V$ satisfying
\begin{gather*}
\varepsilon > 0,
\quad
R \in \grass{{m}+1}{{m}-2},
\quad
T \in \grass{{m}+1}{{m}},
\quad
\text{$C$ is a Borel subset of $T$},
\\
\text{$M$ is an ${m}$ dimensional submanifold
of $\mathbf{R}^{{m}+1}$ of class $\infty$},
\\
\text{$V \in \IVar_{m} ( \mathbf{R}^{{m}+1} )$ is a~curvature varifold
with $\boldsymbol{\Theta}^{m} ( \| V \|, z ) = 1$ for $z \in M$},
\\
\text{$C$, $M$, and $T$, are invariant under translations in
directions belonging to $R$}, \\
\spt \| V \| \subset M\cup T,
\quad
\| \delta V \| \leq \| V \|,
\quad
\| V \| (C) > 0, \quad \boldsymbol{\Theta}^{m} ( \| V \|, c ) = 2,
\\
\| V \| ( \cball{c}{r} \cap \{ z \with \boldsymbol{\Theta}^{m} ( \| V \|, z) = 1 \} )
\geq \omega (r) r^{m},
\\
\begin{aligned}
& \inf \big \{
V( \eqclassification{\cball cr \times \grass{{m}+1}{{m}}}{(z,S)}
{\| \project{S} - \project{T} \| \geq 1/3 } ),
\\
& \phantom{\inf \big \{\ }
\mathscr{H}^{m} ( T \cap \cball cr \without \project T [ \spt \| V \| ] )
\big \}
\geq \omega (r)^2 r^{{m}+2} ( \log (1/r))^{-2}
\end{aligned}
\end{gather*}
whenever $c \in C$ and $0 < r \leq \varepsilon$ and, if ${m} > 2$, then
there also exists a~curvature varifold $V' \in \IVar_2 ( \ker \project R )$
such that
\begin{gather*}
V(k) = \tint{\mathbf{R}^{{m}+1} \times R}{} k(x+y, \im (\project{P} +
\project R)) \ensuremath{\,\mathrm{d}} V' \times \mathscr{H}^{{m}-2} ((x,P),y)
\end{gather*}
whenever $k \in \mathscr{K} \big ( \mathbf{R}^{{m}+1}, \grass {{m}+1} {m}
\big)$.
\end{example}
\begin{proof} [Construction]
Suppose $\mathbf{p}$ and $\mathbf{q}$ as related to ${n}={m}+1$ as in
\ref{miniremark:pqT} define $T = \im \mathbf{p}^\ast$. Notice that it is sufficient
to prove the assertion obtained by replacing closed balls ``$\cball cr$'' by
open cubes ``$\ocube cr$'', see \ref{miniremark:open_cube}. Hence, in view
of \ref{thm:product_varifold}, the construction may be reduced to the case
${m} = 2$ by considering suitable products with ${m}-2$ dimensional
planes if ${m} > 2$. Let $\Delta = 3 \sup \{
\Gamma_{\ref{lemma:bent_catenoid}}, 3 \}$ and choose $0 < \eta \leq 1$ such
that $\omega ( \eta ) \leq 2^{-6} \Delta^{-2}$.
Define a modulus of continuity $\psi$ satisfying the Dini condition such
that $\psi (r) = \sup \{ 8 \Delta \omega (r), 4r^2 \}$ for $0 \leq r \leq
\eta$. Apply \ref{example:cantor_set} with ${m}$, $\omega$, and
$\lambda$ replaced by $2$, $\psi$, and $1/2$ to obtain a number $\delta$,
named ``$\varepsilon$'' there, as well as $G$ and $A$. Let
\begin{gather*}
\varepsilon = \inf \big \{ \delta, \Delta^{-1}, \eta \big \}, \quad B
= \{ r \with 0 < r \leq \varepsilon \}.
\end{gather*}
Define $W \in \IVar_2 ( \mathbf{R}^3 )$ by
\begin{gather*}
W ( k ) = 2 \tint{T \without \mathbf{p}^\ast [ \bigcup G ]}{} k (z,T) \ensuremath{\,\mathrm{d}} \mathscr{H}^2 z
\quad
\text{for $k \in \mathscr{K} (\mathbf{R}^3 \times \grass{3}{2} )$}.
\end{gather*}
Whenever $Q = \ocube{a}{s} \in G$ and $s > \Delta^{-1}$ let $X_Q \in \IVar_2
( \mathbf{R}^3 )$ be defined by
\begin{gather*}
X_Q ( k ) = 2 \tint{T}{} k(z,T) \ensuremath{\,\mathrm{d}} \mathscr{H}^2 z
\quad
\text{for $k \in \mathscr{K} ( \mathbf{R}^3 \times \grass{3}{2} )$}
\end{gather*}
and set $M_Q = \varnothing$. Whenever $Q = \ocube{a}{s} \in G$ and $s \leq
\Delta^{-1}$ apply \ref{lemma:bent_catenoid} with $r$ replaced by $\Delta
s^{-1} \log (1/s)$ to construct a~curvature varifold $X_Q \in \IVar_2 (
\mathbf{R}^3 )$ such that $( \boldsymbol{\mu}_{\Delta s^{-2} \log (1/s)} \circ
\boldsymbol{\tau}_{-a})_\# X_Q$ satisfies the conditions of
\ref{lemma:bent_catenoid} in place of $V$ implying
\begin{gather*}
\| X_Q \| \mathop{\llcorner} \mathbf{p}^{-1} [ \mathbf{R}^2 \without \oball as ]
= 2 \mathscr{H}^2 \mathop{\llcorner} T \without \oball {\mathbf{p}^\ast(a)}s,
\\
\text{$M_Q$ is a two dimensional submanifold of $\mathbf{R}^3$ of class $\infty$},
\\
\boldsymbol{\Theta}^2 ( \| X_Q \|, z ) = 1 \quad \text{for $z \in M_Q$},
\\
\| \delta X_Q \| \leq \| X_Q \|,
\qquad
| \mathbf{q} (z) | \leq s^2 \quad \text{for $z \in \spt \| X_Q \|$},
\\
\| X_Q \| \big ( \ocube{\mathbf{p}^\ast(a)} s \cap
\{ z \with \boldsymbol{\Theta}^2 ( \| X_Q \|, z ) = 1 \} \big )
\geq 2^{-1} \mathscr{L}^2 ( Q ),
\\
\| X_Q \| \big ( \mathbf{p}^{-1} [ Q ] \big )
\leq 2 \mathscr{L}^2 ( Q) + s^4,
\quad
\tint{}{} \| \mathbf{b} (X_Q,z) \| \ensuremath{\,\mathrm{d}} \| X_Q \| z \leq s^2,
\\
\begin{aligned}
& \inf \big \{ X_Q ( \eqclassification{\mathbf{p}^{-1} [ Q ] \times
\grass{3}{2}}{(z,S)}{\| \project{S} - \project{T} \| \geq 1/3 } ),
\\
& \phantom{ \inf\big \{ } \
\mathscr{H}^2 ( \mathbf{p}^\ast [ Q ] \without
\project{T} [ \spt \| X_Q \| ] ) \big \}
\geq \Delta^{-2} s^4 ( \log (1/s))^{-2},
\end{aligned}
\end{gather*}
where $M_Q = \mathbf{p}^{-1} [ \oball as ] \cap \spt \| X_Q \|$. Now, let
$M = \bigcup \{ M_Q \with Q \in G \}$ and define $V \in \IVar_2 ( \mathbf{R}^3
)$ by
\begin{gather*}
V = W + \tsum{Q \in G}{} X_Q \mathop{\llcorner}
( \mathbf{p}^{-1} [ Q ] \times \grass 32 ).
\end{gather*}
Note that $\| V \| (\mathbf{p}^\ast [ A ]) \geq 1$ and $\boldsymbol{\Theta}^2 ( \| V \|,
c ) = 2$ for $\| V \|$ almost all $c \in \mathbf{p}^\ast [ A ]$ by Allard
\cite[2.8\,(4a), 3.5\,(2)]{MR0307015}. Let $C = \classification{\mathbf{p}^\ast
[ A ]}{c}{\boldsymbol{\Theta}^2(\|V\|, c) = 2 }$. Moreover, observe that $V$
is a~curvature varifold with $\| \delta V \| \leq \| V \|$. Finally, if $c
\in \mathbf{p}^\ast [ A ]$ and $r \in B$, then there exists $\ocube as = Q
\in G$ with $Q \subset \oball{\mathbf{p}(c)}{r}$ and $\mathscr{L}^2( Q) = 4s^2 \geq \psi(r)
r^2$, in particular $\ocube{\mathbf{p}^\ast(a)} s \subset \ocube cr$, $s \leq
\Delta^{-1}$, and
\begin{gather*}
\Delta^{-2} s^4 ( \log (1/s) )^{-2}
\geq 2^{-6} \Delta^{-2} \psi (r)^2 r^4 ( \log (1/r) )^{-2}
\geq \omega(r)^2 r^4 ( \log(1/r))^{-2}
\end{gather*}
since $r \geq s \geq \psi (r)^{1/2} r/2 \geq r^2$.
\end{proof}
\begin{remark}
\label{remark:holes-tilt-large}
Concerning the relation of the two terms occurring in the infimum, the
following observation is particularly appropriate. \emph{If $n$, $Q$, $L$,
$M$, $\delta_1$, $\delta_2$, $\delta_3$, $\delta_4$, $\varepsilon$, $m$,
$s$, $S$, $U$, $V$, $\delta$, and $B$ are as in \cite[4.1]{snulmenn.c2},
$p = {m}$, $\psi$ is related to ${m}$, ${n}$, $p$, $U$ and $V$ as in
\ref{miniremark:situation_general}, and $\psi ( U )^{1/{m}} \leq
\delta$, then}
\begin{gather*}
\| V \| (B) \leq 2 \delta^{-1} {n}^{1/2} \besicovitch {n} V ( ( U
\times \grass {n} {m} ) \cap \{(z,R) \with | \project R - \project S
| \geq \delta/2 \} );
\end{gather*}
in fact, this follows from \ref{miniremark:absorb}, Allard
\cite[8.9\,(3)]{MR0307015}, and the Besicovitch-Federer covering theorem.
\end{remark}
\begin{remark}
\label{remark:example-super-quadratic-tilt}
Taking $\omega$ in \ref{example:quantitative_brakke} such that $\omega (t) =
( \log (1/t))^{-1} (\log(\log(1/t)))^{-2}$ for $0 < t \leq e^{-e}$, where
$e$ denotes the Euler's number, one obtains
\begin{gather*}
\lim_{r \to 0+} r^{-{m}-2-\delta}
\tint{\cball{c}{r} \times \grass{{n}}{{m}}}{}
\| \project{S} - \project{T} \|^\iota \ensuremath{\,\mathrm{d}} V (z,S) = \infty
\end{gather*}
whenever $c \in C$, $\delta > 0$, and $1 \leq \iota < \infty$. Taking $\iota
> 2$ and $\delta = \iota -2$, one infers
\begin{gather*}
\tint{M \cap \oball cr}{} \| \mathbf{b} (M,z) \|^q \ensuremath{\,\mathrm{d}} \mathscr{H}^{m} z
= \infty
\end{gather*}
whenever $c \in \spt ( \| V \| \mathop{\llcorner} C )$, $0 < r < \infty$, and $1 < q
< \infty$; in fact, the Cartesian product structure of $M$ and $V$ reduces
the problem to the case ${m} = 2$ in which, in view of
\ref{lemma:second_fundamental_form}\,\eqref{item:second_fundamental_form:norm_equiv}
and \ref{remark:curvature_varifold}, one may apply
\cite[11.4\,(3)]{snulmenn:tv.v2} with $f(z)$ replaced by $\project{\Tan^2 (
\| V \|,z)}$.
\end{remark}
\begin{remark}
\label{remark:brakke_example}
For comparison note the following well known proposition: \emph{If
${m}$ and ${n}$ are positive integers, ${m} \leq {n}$, $0 \leq
K < \infty$, $V \in \IVar_{m} ( \mathbf{R}^{n} )$ with $\| \delta V \|
\leq K \| V \|$ then there exists a relatively open, dense subset $A$
of $\spt \| V \|$ such that for any $1 \leq q < \infty$ there holds
\begin{gather*}
\limsup_{r \to 0+} r^{-{m}-q} \tint{\cball{a}{r}
\times \grass{{n}}{{m}}}{} | \project{S} - \project{T} |^q
\ensuremath{\,\mathrm{d}} V (z,S) < \infty
\end{gather*}
for $V$ almost all $(a,T) \in A \times \grass{{n}}{{m}}$}; in fact,
one may combine Allard \cite[8.1\,(1)]{MR0307015} with elliptic regularity
theory as provided, e.g.,~in \cite[3.6, 3.21]{snulmenn.c2} and properties of
Sobolev functions, see Calder\'on and Zygmund \cite[Theorem~12,
p.~204]{MR0136849} or \cite[Theorem 3.4.2]{MR1014685}. In particular, the
tangent plane behaviour exhibited in the preceding example may not occur at
$V$ almost all points.
\end{remark}
\begin{remark}
Example \ref{example:quantitative_brakke} is a refinement of the example
described by Brakke in \cite[6.1]{MR485012}.
\end{remark}
\begin{example}
\label{example:quantitative_brakke_again}
Suppose ${m}$ is an integer with ${m} \geq 2$ and $\omega$ is a
modulus of continuity.
Then there exist $\varepsilon$, $B$, $C$, $M$, $T$, and $V$ satisfying
\begin{gather*}
\varepsilon > 0, \quad B \subset \mathbf{R} \cap \{ t \with t > 0 \}, \quad
R \in \grass{{m}+1}{{m}-2}, \quad T \in \grass{{m}+1}{{m}}, \\
\inf B = 0, \quad \text{$C$ is a Borel subset of $T$}, \\
\text{$M$ is an ${m}$ dimensional submanifold of $\mathbf{R}^{{m}+1}$ of
class $\infty$}, \\
\text{$V \in \IVar_{m} ( \mathbf{R}^{{m}+1} )$ is a~curvature varifold
with $\boldsymbol{\Theta}^{m} ( \| V \|, z ) = 1$ for $z \in M$}, \\
\text{$C$, $M$, and $T$ are invariant under translations in directions
belonging to $R$}, \\
\spt \| V \| \subset M \cup T, \quad \| \delta V \| \leq \| V \|,
\quad \| V \| (C) > 0, \quad \boldsymbol{\Theta}^{m} ( \| V \|, c ) = 2, \\
\| V \| ( \cball{c}{r} \cap \{ z \with \boldsymbol{\Theta}^{m} ( \| V \|, z) = 1
\} ) \geq \omega (r) r^{m} \quad \text{for $0 < r \leq
\varepsilon$}, \\
\begin{aligned}
& \inf \big \{ V( \eqclassification{\cball cr \times
\grass{{m}+1}{{m}}}{(z,S)} {\| \project{S} - \project{T} \|
\geq 1/3 } ), \\
& \phantom{\inf \big \{\ } \mathscr{H}^{m} ( T \cap \cball cr
\without \project T [ \spt \| V \| ] ) \big \} \geq \omega
(r) r^{{m}+2} ( \log (1/r))^{-2} \quad \text{for $r \in B$}
\end{aligned}
\end{gather*}
whenever $c \in C$ and, if ${m} > 2$, then there also exists a~curvature
varifold $V' \in \IVar_2 ( \ker \project R)$ such that
\begin{gather*}
V(k) = \tint{\mathbf{R}^{{m}+1} \times R}{}
k(x+y, \im (\project{P} + \project R))
\ensuremath{\,\mathrm{d}} V' \times \mathscr{H}^{{m}-2} ((x,P),y)
\end{gather*}
whenever $k \in \mathscr{K} \big ( \mathbf{R}^{{m}+1}, \grass {{m}+1} {m}
\big)$.
\end{example}
\begin{proof} [Construction]
Modify the construction of \ref{example:quantitative_brakke} by replacing
its second paragraph by ``Define a modulus of continuity $\psi$ such that
$\psi (r) = \sup \{ 8 \Delta \omega (r)^{1/2}, 4r^2 \}$ for $0 \leq r \leq
\eta$. Apply \ref{example:yet_another_cantor_set} with ${m}$, $\omega$,
and $\lambda$ replaced by $2$, $\psi$, and $1/2$ to obtain a number
$\delta$, named `$\varepsilon$' there, as well as $B$, $G$ and $A$. Let
$\varepsilon = \inf \big \{ \delta, \Delta^{-1}, \eta \big \}$.'' and
``$\omega(r)^2$'' in the last displayed inequality by ``$\omega(r)$'', and
adding ``If $c \in \mathbf{p}^\ast [ A ]$ and $0 < r \leq \varepsilon$ there
exists $H$ such that $H \subset G \cap \{ Q \with Q \subset \oball{\mathbf{p} (c)}r
\}$ and $\mathscr{L}^2 ( \bigcup H ) \geq \psi (r)r^2$, in particular
$\ocube as \in H$ implies $\ocube{\mathbf{p}^\ast(a)}s \subset \ocube cr$ and $s
\leq \Delta^{-1}$.'' at the end, to obtain a construction for the present
assertion.
\end{proof}
\begin{remark}
The main modification of the construction of
\ref{example:quantitative_brakke_again} in comparison to
\ref{example:quantitative_brakke} is the usage of
\ref{example:yet_another_cantor_set} in place of \ref{example:cantor_set}
and that $B$ is a (countable) set constructed in
\ref{example:yet_another_cantor_set} rather than an interval.
\end{remark}
\begin{remark}
\label{remark:second_fundamental_form_large}
As in \ref{remark:example-super-quadratic-tilt}, one obtains
\begin{gather*}
\tint{M \cap \oball cr}{} \| \mathbf{b} (M,z) \|^q \ensuremath{\,\mathrm{d}}
\mathscr{H}^{m} z = \infty
\end{gather*}
whenever $c \in \spt ( \| V \| \mathop{\llcorner} C )$, $0 < r < \infty$, and $1 <
q < \infty$.
\end{remark}
\begin{remark} \label{remark:density-ap-lsc}
Since $\boldsymbol{\Theta}^{m} ( \| V \|, c ) = 2$ for $c \in C$, the lower bound
on $$\| V \| ( \cball cr \cap \{ z \with \boldsymbol{\Theta}^{m} ( \| V \|, z ) = 1
\} )$$ is the largest one permitted by the approximate continuity of
$\boldsymbol{\Theta}^{m} ( \| V \|, \cdot )$ with respect to $\| V\|$ and the
standard Vitali relation, see \cite[2.8.18, 2.9.13]{MR41:1976}.
\end{remark}
\section{Cartesian product of varifolds}
\label{sec:varifold-theory}
The purpose of this section is to establish basic properties of the
Cartesian product of varifolds. The construction preserves rectifiability,
integrality and maps curvature varifolds to curvature varifolds, see
\ref{thm:product_varifold}. We also note a version of the coarea formula for
rectifiable varifolds, see
\ref{miniremark:rect_varifold}\,\eqref{item:rect:varifold:coarea}.
The proof of the rectifiability of the Cartesian product of rectifiable
varifolds needs to take into account that \emph{$(\mathscr{H}^{m},{m})$
rectifiable sets do not possess a similar stability property}, see
\ref{remark:product_rectifiable_sets}. Our treatment of curvature varifolds is
based on the characterisation of such varifolds in terms of generalised weakly
differentiable functions obtained in~\cite[15.6]{snulmenn:tv.v2}.
In the present paper only products of varifolds with planes are employed.
More general products will be required in the study of the geodesic distance
on the support of the weight measure of certain varifolds, see \cite[\S
6]{snulmenn:sobolev.v2}.
\begin{lemma} \label{lemma:second_fundamental_form}
Suppose ${m}, {n} \in \mathscr{P}$, ${m} \leq {n}$, $M$ is an ${m}$
dimensional submanifold of $\mathbf{R}^{n}$ of class $2$, $\tau : M \to
\Hom ( \mathbf{R}^{n}, \mathbf{R}^{n} )$ is defined by $\tau (z) = \project{\Tan
(M,z)}$ for $z \in M$, and $\Hom ( \mathbf{R}^{n}, \mathbf{R}^{n} )$ is normed by
$\| \cdot \|$.
Then the following three statements hold.
\begin{enumerate}
\item \label{item:second_fundamental_form:b_from_tau} If $z \in M$ and
$u,v \in \Tan (M,z)$, then $\mathbf{b} (M,z) (u,v) = \langle u,
\langle v, \Der \tau (z) \rangle \rangle$.
\item \label{item:second_fundamental_form:tau_from_b} If $z \in M$ and
$u,v,w \in \mathbf{R}^{n}$ then
\begin{align*}
& \langle v, \langle u, \Der \tau (z) \circ \tau (z) \rangle
\rangle \bullet w \\
& \quad = \mathbf{b} (M,z) ( \langle u, \tau(z) \rangle, \langle
v, \tau (z) \rangle ) \bullet w + \mathbf{b} (M,z) ( \langle u,
\tau (z) \rangle, \langle w, \tau (z) \rangle ) \bullet v.
\end{align*}
\item \label{item:second_fundamental_form:norm_equiv} If $z \in M$,
then $\| \mathbf{b} (M,z) \| = \| \Der \tau (z) \circ \tau (z) \|$.
\end{enumerate}
\end{lemma}
\begin{proof}
Define $\nu : M \to \Hom ( \mathbf{R}^{n}, \mathbf{R}^{n} )$ by $\nu (z) =
\id{\mathbf{R}^{n}} - \tau (z)$ for $z \in M$. Differentiating the equations
$\tau (z) \circ \tau (z) = \tau (z)$ and $\nu (z) \circ \tau (z) = 0$
for $z \in M$, one obtains for $z \in M$ and $u \in \Tan (M,z)$ that
\begin{gather*}
\tau (z) \circ \langle u, \Der \tau (z) \rangle \circ \tau (z) = 0,
\quad
\nu (z) \circ \langle u, \Der \tau (z) \rangle \circ \nu (z) = 0.
\end{gather*}
In order to prove \eqref{item:second_fundamental_form:b_from_tau}, suppose
$z \in M$ and $u,v \in \Tan (M,z)$, notice that $\langle u, \langle v,
\Der \tau (z) \rangle \rangle \in \Nor (M,z)$, and differentiate the
equation $\langle u, \tau (\zeta) \rangle \bullet g ( \zeta ) = 0$ for
$\zeta \in M$, to obtain $\langle u, \langle v,\Der \tau (z) \rangle
\rangle \bullet g(z) = - u \bullet \langle v, \Der g(z) \rangle $.
Expressing
\begin{align*}
\langle v, \langle u, \Der \tau (z) \circ \tau (z) \rangle \rangle
\bullet w & = \langle \langle v, \tau (z) \rangle, \langle u, \Der
\tau (z) \circ \tau (z) \rangle \rangle \bullet \langle w, \nu (z)
\rangle \\
& \phantom{=} \ + \langle \langle v, \nu (z) \rangle, \langle u, \Der
\tau (z) \circ \tau (z) \rangle \rangle \bullet \langle w, \tau (z)
\rangle
\end{align*}
for $z \in M$ and $u,v,w \in \mathbf{R}^{n}$,
\eqref{item:second_fundamental_form:tau_from_b} follows from the symmetry
of $\langle u, \Der \tau (z) \circ \tau (z) \rangle$ and
\eqref{item:second_fundamental_form:b_from_tau}. Finally, noting
\begin{gather*}
| \langle v, \tau (z) \rangle || \langle w, \nu (z) \rangle | + |
\langle v, \nu (z) \rangle | | \langle w, \tau (z) \rangle | \leq
|v||w|
\end{gather*}
for $z \in M$ and $v,w \in \mathbf{R}^{n}$ by H{\"o}lder's inequality,
\eqref{item:second_fundamental_form:norm_equiv} follows from
\eqref{item:second_fundamental_form:b_from_tau} and
\eqref{item:second_fundamental_form:tau_from_b}.
\end{proof}
\begin{remark}
Items
\ref{lemma:second_fundamental_form}\,\eqref{item:second_fundamental_form:b_from_tau}\,\eqref{item:second_fundamental_form:tau_from_b}
are in analogy with Hutchinson \cite[5.1.1\,(i)\,(ii)]{MR825628}.
\end{remark}
\begin{definition} \label{def:curvature_varifold}
Suppose ${m}, {n} \in \mathscr{P}$, ${m} \leq {n}$, and $U$ is an open
subset of $\mathbf{R}^{n}$.
Then $V$ is called \emph{${m}$ dimensional curvature varifold in $U$} if
and only if the following three conditions are satisfied:
\begin{enumerate}
\item $V$ is an ${m}$ dimensional integral varifold in $U$.
\item $\| \delta V \|$ is a Radon measure absolutely continuous with
respect to $\| V\|$.
\item If $Y = \Hom ( \mathbf{R}^{n}, \mathbf{R}^{n} ) \cap \{ \sigma \with
\sigma = \sigma^\ast \}$, $Z = U \cap \{ z \with \Tan^{m} ( \|V \|,
z ) \in \grass {n} {m} \}$, and $\tau : Z \to Y$ is defined by
$\tau (z) = \project{\Tan^{m} ( \| V \|, z )}$ for $z \in Z$, then
$\tau$ is a generalised $V$ weakly differentiable function.
\end{enumerate}
In this case one defines for $z \in Z \cap \dmn \derivative V \tau$ the
\emph{second fundamental form of~$V$ at~$z$} by
\begin{gather*}
\mathbf{b} (V,z) : \Tan^{m} ( \| V \|, z ) \times \Tan^{m} ( \| V \|, z ) \to \mathbf{R}^{n},
\\
\mathbf{b} (V,z) (u,v) = \langle u, \langle v, \derivative{V}{\tau}
(z) \rangle \rangle \quad \text{whenever $u,v \in \Tan^{m} ( \| V
\|, z )$}.
\end{gather*}
\end{definition}
\begin{remark} \label{remark:curvature_varifold} In view of
\cite[15.4--15.6]{snulmenn:tv.v2} the preceding definition of curvature
varifold is equivalent to Hutchinson's original definition in
\cite[5.2.1]{MR825628}. Recalling \cite[4.8]{snulmenn.c2} and
\cite[11.2]{snulmenn:tv.v2}, one infers that, for $\| V \|$ almost all $z$, the
second fundamental form of $V$ at $z$ is a symmetric bilinear map with
values in $\Nor^{m} ( \| V \|, z)$ which is related to
$\derivative{V}{\tau} (z)$ as $\mathbf{b}(M,z)$ is related to $\Der \tau (z)
\circ \tau (z)$ in
\ref{lemma:second_fundamental_form}\,\eqref{item:second_fundamental_form:tau_from_b}\,\eqref{item:second_fundamental_form:norm_equiv};
in fact, if $M$ is an ${m}$ dimensional submanifold of $\mathbf{R}^{n}$ of
class~$2$ and $\sigma : M \to \Hom ( \mathbf{R}^{n} , \mathbf{R}^{n} )$ satisfies
$\sigma (z) = \project{\Tan (M,z)}$ for $z \in M$, then
\begin{gather*}
\Tan(M,z) = \Tan^{m}(\|V\|,z) \quad \text{and} \quad \Der \sigma (z) = (
\| V \|, {m} ) \ap \Der \tau (z)
\end{gather*}
for $\| V\|$ almost all $z \in U \cap M$ by \cite[2.8.18, 2.9.11,
3.2.17]{MR41:1976} and Allard \cite[3.5\,(2)]{MR0307015}. Notice also
that, for $\| V\|$ almost all $z$,
\begin{gather*}
\mathbf{b} (V,z) \circ
( \project{\Tan^{m} ( \| V \|, z )} \times \project{\Tan^{m} ( \| V \|, z )} )
\end{gather*}
corresponds to the generalised second fundamental form at $(z,\Tan^{m} (
\| V \|, z) )$ in the sense of Hutchinson \cite[5.2.5]{MR825628}.
\end{remark}
\begin{miniremark}
\label{miniremark:traces}
Suppose $V$ and $W$ are finite dimensional vectorspaces, $f \in \Hom (
V,W)$, and $g \in \Hom (W,V)$. Then $\trace ( g \circ f ) = \trace ( f \circ
g)$; in fact, the argument of \cite[1.4.5]{MR41:1976} for the case $V=W$
applies to the present case as well.
\end{miniremark}
\begin{miniremark}
\label{miniremark:rect_varifold}
Suppose ${m}, {n} \in \mathscr{P}$, ${m} \leq {n}$, $U$ is an open subset
of $\mathbf{R}^{n}$, and $V \in \RVar_{m} (U)$. Then the following two
statements hold.
\begin{enumerate}
\item
\label{item:rect_varifold:c1}
There exist sequences of compact subset $C_j$ of ${m}$ dimensional
submanifolds of $U$ of class $1$ and $0 < d_j < \infty$ such that
\begin{gather*}
V(k) = \tsum{j=1}{\infty} d_j \tint{C_j}{} k (z,\Tan^{m}
(\mathscr{H}^{m} \mathop{\llcorner} C_j,z)) \ensuremath{\,\mathrm{d}} \mathscr{H}^{m} z
\end{gather*}
for $k \in \mathscr{K} ( U \times \grass {n} {m} )$.
\item \label{item:rect:varifold:coarea} If ${m} \geq \mu \in \mathscr{P}$,
$M = \{ z \with 0 < \boldsymbol{\Theta}^{m} ( \| V \|, z ) < \infty \}$, and $f
: U \to \mathbf{R}^\mu$ is Lipschitzian, then $f$ is $(\| V \|, {m})$
differentiable at $\| V \|$ almost all $z$ and there holds
\begin{align*}
& \tint{}{} g(z) {\textstyle \big \| \bigwedge_\mu (
\| V \|, {m} ) \ap \Der f(z) \big \|} \ensuremath{\,\mathrm{d}} \| V \| z \\
& \qquad = \tint{}{} \tint{M \cap f^{-1} [ \{ y \} ]}{}
g(z) \boldsymbol{\Theta}^{m} ( \| V \|, z ) \ensuremath{\,\mathrm{d}}
\mathscr{H}^{{m}-\mu} z \ensuremath{\,\mathrm{d}} \mathscr{L}^\mu y
\end{align*}
whenever $g$ is a $\| V \|$ integrable $\overline \mathbf{R}$ valued
function, where $\infty \cdot 0 = 0$.
\end{enumerate}
\eqref{item:rect_varifold:c1} may be verified by means of \cite[2.8.18,
2.9.11, 3.2.17, 3.2.29]{MR41:1976}. The first half of
\eqref{item:rect:varifold:coarea} is implied by
\cite[4.5\,(2)]{snulmenn.decay}. Concerning the second half of
\eqref{item:rect:varifold:coarea}, one notices that $g|C_j$ is
$\mathscr{H}^{m} \mathop{\llcorner} C_j$ integrable and
\begin{gather*}
\boldsymbol{\Theta}^{m} ( \| V \|, z ) = \tsum{j \in J(z)}{} d_j \quad
\text{for $\mathscr{H}^{m}$ almost all $z \in U$}, \\
( \| V \|, {m} ) \ap \Der f(z) = ( \mathscr{H}^{m} \mathop{\llcorner} C_j, {m}
) \ap \Der f(z) \quad \text{for $\mathscr{H}^{m}$ almost all $z \in C_j$},
\end{gather*}
where $J(z) = \mathscr{P} \cap \{j \with z \in C_j \}$, by Allard
\cite[2.8\,(4a), 3.5\,(2)]{MR0307015} and obtains
\begin{align*}
& {\textstyle \tint{C_j}{} g (z) \big \| \bigwedge_\mu (
\mathscr{H}^{m} \mathop{\llcorner} C_j, {m} ) \ap \Der f(z) \big \| \ensuremath{\,\mathrm{d}}
\mathscr{H}^{m} z} \\
& \qquad = \tint{}{} \tint{C_j \cap f^{-1} [ \{ y \} ]}{} g (z)
\ensuremath{\,\mathrm{d}} \mathscr{H}^{{m}-\mu} z \ensuremath{\,\mathrm{d}} \mathscr{L}^\mu y
\end{align*}
from \cite[2.10.35, 3.2.22\,(3)]{MR41:1976}, hence multiplying by $d_j$
and summing over $j$ by means of \cite[2.10.25]{MR41:1976} yields the
conclusion since $M$ and $\bigcup_{j=1}^\infty C_j$ are
$\mathscr{H}^{m}$ almost equal by Allard \cite[3.5\,(1b)]{MR0307015} and
the equation for $\boldsymbol{\Theta}^{m} ( \| V \|, z )$.
\end{miniremark}
\begin{theorem} \label{thm:product_varifold}
Suppose for $i \in \{ 1,2 \}$, ${m}_i, {n}_i \in \mathscr{P}$, ${m}_i
\leq {n}_i$, $p_i : \mathbf{R}^{{n}_1} \times \mathbf{R}^{{n}_2} \to
\mathbf{R}^{{n}_i}$ satisfy $p_i (z_1,z_2) = z_i$ for $(z_1,z_2) \in
\mathbf{R}^{{n}_1} \times\mathbf{R}^{{n}_2}$, $U_i$ are open subsets of
$\mathbf{R}^{{n}_i}$, $V_i \in \Var_{{m}_i} ( U_i )$, and $W \in
\Var_{{m}_1+{m}_2} ( U_1 \times U_2 )$ satisfies
\begin{gather*}
W(k) = \tint{}{} k ((z_1,z_2), S_1 \times S_2) \ensuremath{\,\mathrm{d}} (V_1 \times V_2)
((z_1,S_1),(z_2,S_2))
\end{gather*}
for $k \in \mathscr{K} \big ( ( U_1 \times U_2 ) \times
\grass{\mathbf{R}^{{n}_1} \times \mathbf{R}^{{n}_2}}{ {m}_1+{m}_2 } \big )$.
Then the following eight statements hold.
\begin{enumerate}
\item \label{item:product_varifold:weight} There holds $\| W \| = \|
V_1 \| \times \| V_2 \|$.
\item \label{item:product_varifold:rectifiable} If $V_1$ and $V_2$ are
rectifiable, so is $W$ and, for $\| W \|$ almost all $(z_1,z_2)$,
\begin{gather*}
\Tan^{{m}_1+{m}_2} ( \| W \|, (z_1,z_2)) = \Tan^{{m}_1} ( \|
V_1 \| , z_1 ) \times \Tan^{{m}_2} ( \| V_2 \|, z_2 ), \\
\boldsymbol{\Theta}^{{m}_1+{m}_2} ( \| W \|, (z_1,z_2) ) =
\boldsymbol{\Theta}^{{m}_1} ( \| V_1 \|, z_1 ) \boldsymbol{\Theta}^{{m}_2} ( \|V_2\|
, z_2 ).
\end{gather*}
\item \label{item:product_varifold:integral} If $V_1$ and $V_2$ are
integral, so is $W$.
\item \label{item:product_varifold:planes} If $S_i \in \grass
{{n}_i}{{m}_i}$ for $i \in \{1,2\}$ and $h \in \Hom (
\mathbf{R}^{{n}_1} \times \mathbf{R}^{{n}_2}, \mathbf{R}^{{n}_1} \times
\mathbf{R}^{{n}_2} )$, then
\begin{gather*}
\eqproject{S_1 \times S_2} = p_1^\ast \circ \eqproject{S_1} \circ
p_1 + p_2^\ast \circ \eqproject{S_2} \circ p_2, \\
\eqproject{S_1 \times S_2} \bullet h = \eqproject{S_1} \bullet (
p_1 \circ h \circ p_1^\ast ) + \eqproject{S_2} \bullet ( p_2
\circ h \circ p_2^\ast ).
\end{gather*}
\item \label{item:product_varifold:variation} If $\theta \in
\mathscr{D} (U_1 \times U_2, \mathbf{R}^{{n}_1} \times \mathbf{R}^{{n}_2} )$,
then
\begin{align*}
( \delta W ) ( \theta ) & = \tint{}{} ( \delta V_1)_{z_1} ( p_1 (
\theta (z_1,z_2))) \ensuremath{\,\mathrm{d}} \| V_2 \| z_2 \\
& \phantom{=}\ + \tint{}{} ( \delta V_2)_{z_2} ( p_2 ( \theta
(z_1,z_2))) \ensuremath{\,\mathrm{d}} \| V_1 \| z_1.
\end{align*}
\item \label{item:product_varifold:variation_measure} If $\| \delta
V_i \|$ are Radon measures for $i \in \{ 1,2 \}$, then
\begin{gather*}
\| \delta W \| \leq \| \delta V_1 \| \times \| V_2 \| + \| V_1 \|
\times \| \delta V_2 \|
\end{gather*}
and, for $\theta \in \Lp{1} \big ( \| \delta V_1 \| \times \| V_2 \| +
\| V_1 \| \times \| \delta V_2 \|, \mathbf{R}^{{n}_1} \times
\mathbf{R}^{{n}_2} \big )$, the equation in
\eqref{item:product_varifold:variation} holds.
\item \label{item:product_varifold:tv_again} If, for $i \in \{ 1,2
\}$, $V_i$ are rectifiable, $\| \delta V_i \|$ are Radon measures,
$Y_i$ are finite dimensional normed vectorspaces and $f_i \in \mathbf{T}
(V_i, Y_i)$, and $f : \dmn f_1 \times \dmn f_2 \to Y_1 \times Y_2$
satisfies
\begin{gather*}
f(z_1,z_2) = (f_1(z_1),f_2(z_2)) \quad \text{for $z_1 \in \dmn
f_1$ and $z_2 \in \dmn f_2$},
\end{gather*}
then $f \in \mathbf{T} (W, Y_1 \times Y_2)$ and, for $\| W \|$ almost all
$(z_1,z_2)$,
\begin{gather*}
\derivative{W}{f} (z_1,z_2) (u_1,u_2) = ( \derivative{V_1}{f_1}
(z_1)(u_1), \derivative{V_2}{f_2} (z_2)(u_2) )
\end{gather*}
whenever $u_1 \in \mathbf{R}^{{n}_1}$ and $u_2 \in \mathbf{R}^{{n}_2}$.
\item \label{item:product_varifold:curvature} If $V_1$ and $V_2$ are
curvature varifolds, then so is $W$ and, for $\| W \|$ almost all
$(z_1,z_2)$,
\begin{gather*}
\mathbf{b} (W,(z_1,z_2)) ( (u_1,u_2), (v_1,v_2)) = ( \mathbf{b}
(V_1,z_1) (u_1,u_2), \mathbf{b} (V_2,z_2) (u_2,v_2) )
\end{gather*}
whenever $u_1,v_1 \in \mathbf{R}^{{n}_1}$ and $u_2,v_2 \in
\mathbf{R}^{{n}_2}$.
\end{enumerate}
\end{theorem}
\begin{proof} [Proof of
\eqref{item:product_varifold:weight}]
\eqref{item:product_varifold:weight} follows from Fubini's theorem.
\end{proof}
\begin{proof} [Proof of \eqref{item:product_varifold:rectifiable}]
If $C_i$ are compact subsets of ${m}_i$ dimensional submanifolds of $U_i$
of class $1$ for $i \in \{ 1,2 \}$, then $C_1 \times C_2$ is
$(\mathscr{H}^{{m}_1+{m}_2},{m}_1+{m}_2)$ rectifiable with
\begin{gather*}
( \mathscr{H}^{{m}_1} \mathop{\llcorner} C_1 ) \times ( \mathscr{H}^{{m}_2}
\mathop{\llcorner} C_2 ) = \mathscr{H}^{{m}_1+{m}_2} \mathop{\llcorner} ( C_1 \times
C_2 ),
\end{gather*}
by \cite[3.2.23]{MR41:1976} and, for $\mathscr{H}^{{m}_1+{m}_2}$
almost all $(z_1,z_2) \in C_1 \times C_2$,
\begin{align*}
& \Tan^{{m}_1} ( \mathscr{H}^{{m}_1} \mathop{\llcorner} C_1, z_1 ) \times
\Tan^{{m}_2} ( \mathscr{H}^{{m}_2} \mathop{\llcorner} C_2, z_2 ) \\
& \qquad = \Tan^{{m}_1+{m}_2} ( \mathscr{H}^{{m}_1+{m}_2}
\mathop{\llcorner} ( C_1 \times C_2 ), (z_1,z_2) )
\end{align*}
by \cite[2.8.18, 2.9.11]{MR41:1976}. Therefore the assertion may be verified
by means of \ref{miniremark:rect_varifold}\,\eqref{item:rect_varifold:c1},
Allard \cite[2.8\,(4a), 3.5\,(2)]{MR0307015}, \cite[2.1.1\,(11)]{MR41:1976},
and \eqref{item:product_varifold:weight}.
\end{proof}
\begin{proof} [Proof of \eqref{item:product_varifold:integral}]
This is a consequence of \eqref{item:product_varifold:rectifiable} and
Allard \cite[3.5\,(1c)]{MR0307015}.
\end{proof}
\begin{proof} [Proof of \eqref{item:product_varifold:planes}\,\eqref{item:product_varifold:variation}\,\eqref{item:product_varifold:variation_measure}]
The first equation of \eqref{item:product_varifold:planes} is obvious and
implies the second equation by \ref{miniremark:traces}.
\eqref{item:product_varifold:variation} is a consequence of
\eqref{item:product_varifold:planes} and Fubini's theorem.
\eqref{item:product_varifold:variation} implies
\eqref{item:product_varifold:variation_measure}.
\end{proof}
\begin{proof} [Proof of \eqref{item:product_varifold:tv_again}]
Assume $\dim ( Y_1 \times Y_2 ) \geq 1$. First, consider the case $\dim
(Y_1 \times Y_2) \geq 2$. Define $F : p_1^{-1} [ \dmn
\derivative{V_1}{f_1} ] \cap p_2^{-1} [ \dmn \derivative{V_2}{f_2}
] \to \Hom ( \mathbf{R}^{{n}_1} \times \mathbf{R}^{{n}_2}, Y_1 \times Y_2 )$ by
\begin{gather*}
F (z_1,z_2)(u_1,u_2) = ( \derivative{V_1}{f_1} (z_1)(u_1),
\derivative{V_2}{f_2}(z_2)(u_2) )
\end{gather*}
whenever $z_1 \in \dmn \derivative{V_i}{f_i}$ and $u_i \in \mathbf{R}^{{n}_i}$
for $i \in \{ 1,2 \}$. By \cite[8.4]{snulmenn:tv.v2} it sufficient to prove
\begin{align*}
& ( \delta W ) ( ( \gamma \circ f ) \theta ) \\
& \qquad = \tint{}{} \gamma (f(z)) \project{T} \bullet \Der \theta
(z) \ensuremath{\,\mathrm{d}} W (z,T) + \tint{}{} \langle \theta (z), \Der \gamma ( f(z))
\circ F (z) \rangle \ensuremath{\,\mathrm{d}} \| W \| z
\end{align*}
whenever $\theta \in \mathscr{D} (U_1 \times U_2, \mathbf{R}^{{n}_1} \times
\mathbf{R}^{{n}_2} )$ and $\gamma \in \mathscr{D} ( Y_1 \times Y_2, \mathbf{R} )$.
Recalling \cite[2.15, 3.1]{snulmenn:tv.v2},\footnote{The topologies on
$\mathscr{D} ( Y_1 \times Y_2, \mathbf{R} )$ considered in
\cite[2.13]{snulmenn:tv.v2} and \cite[4.1.1]{MR41:1976} differ but possess
the same convergent sequences, see \cite[2.15, 2.17\,(3)]{snulmenn:tv.v2}.}
one may assume additionally that for some $\gamma_i \in \mathscr{D}
(Y_i,\mathbf{R})$ for $i \in \{1,2 \}$, there holds
\begin{gather*}
\gamma (y_1,y_2) = \gamma_1 (y_1) \gamma_2 (y_2) \quad \text{for
$(y_1,y_2) \in Y_1 \times Y_2$}.
\end{gather*}
In this case one computes, noting
\begin{align*}
& ( \delta V_i )_{z_i} ( \gamma_i (f_i(z_i)) p_i ( \theta (z_1,z_2)) )
\\
& \qquad = \tint{}{} \gamma_i (f_i(z_i)) \eqproject{S_i} \bullet ( p_i
\circ \Der \theta (z_1,z_2) \circ p_i^\ast ) \ensuremath{\,\mathrm{d}} V_i (z_i,S_i) \\
& \qquad \phantom = \ + \tint{}{} \langle p_i (\theta (z_1,z_2)), \Der
\gamma_i ( f_i (z_i) ) \circ \derivative{V_i}{f_i} (z_i) \rangle \ensuremath{\,\mathrm{d}}
\| V_i \| z_i
\end{align*}
whenever $i,j \in\{ 1,2 \}$, $i \neq j$ and $z_j \in \dmn f_j$, and using
\eqref{item:product_varifold:planes} and
\eqref{item:product_varifold:variation_measure},
\begin{align*}
& ( \delta W ) ( ( \gamma \circ f ) \theta ) \\
& \qquad = \tint{}{} \gamma_2 ( f_2 (z_2) ) ( \delta V_1 )_{z_1} (
\gamma_1 ( f_1 (z_1)) p_1 ( \theta (z_1,z_2)) ) \ensuremath{\,\mathrm{d}} \| V_2 \| z_2 \\
& \qquad \phantom = \ + \tint{}{} \gamma_1 ( f_1 (z_1) ) ( \delta V_2
)_{z_2} ( \gamma_2 ( f_2 (z_2)) p_2 ( \theta (z_1,z_2)) ) \ensuremath{\,\mathrm{d}} \| V_1
\| z_1 \\
& \qquad = \tint{}{} \gamma (f(z)) \project{T} \bullet \Der \theta
(z) \ensuremath{\,\mathrm{d}} W (z,T) + \tint{}{} \langle \theta (z), \Der \gamma ( f(z))
\circ F (z) \rangle \ensuremath{\,\mathrm{d}} \| W \| z.
\end{align*}
If $\dim ( Y_1 \times Y_2 ) = 1$, one may assume $\dim Y_1 = 1$, hence
$\dim Y_2 = 0$, and similarly consider $\gamma_1 \in \mathscr{E} (Y_1,
\mathbf{R} )$ such that $\spt \Der \gamma_1$ is compact and $\gamma_2 = 1$.
\end{proof}
\begin{proof} [Proof of \eqref{item:product_varifold:curvature}]
Define $Y_i = \Hom ( \mathbf{R}^{{n}_i}, \mathbf{R}^{{n}_i} ) \cap \{ \sigma \with
\sigma = \sigma^\ast \}$ and functions $f_i \in \mathbf{T} ( V_i, Y_i)$ by $f_i
(z_i) = \project{\Tan^{{m}_i} ( \| V_i \|, z_i )}$ whenever $z_i \in U_i$
and $\Tan^{{m}_i} ( \| V_i \|, z_i ) \in \grass {{n}_i}{{m}_i}$ for $i
\in \{ 1, 2 \}$. Associate $f \in \mathbf{T} (W,Y_1 \times Y_2 )$ to $f_1$ and
$f_2$ as in \eqref{item:product_varifold:tv_again}. Define $Y$ to be the
vectorspace of symmetric endomorphisms of $\mathbf{R}^{{n}_1} \times
\mathbf{R}^{{n}_2}$ and let the linear map $L : Y_1 \times Y_2 \to Y$ be defined
by
\begin{gather*}
L ( \sigma_1, \sigma_2 ) = p_1^\ast \circ \sigma_1 \circ p_1 + p_2^\ast
\circ \sigma_2 \circ p_2 \quad \text{for $\sigma_1 \in Y_1$ and
$\sigma_2 \in Y_2$}.
\end{gather*}
In view of \eqref{item:product_varifold:rectifiable} and
\eqref{item:product_varifold:planes}, one infers
\begin{gather*}
\project{\Tan^{{m}_1+{m}_2} ( \| W \|, (z_1,z_2))} = L ( f
(z_1,z_2)) \quad \text{for $\| W \|$ almost all $(z_1,z_2)$}.
\end{gather*}
Therefore $W$ is a curvature varifold by
\eqref{item:product_varifold:weight},
\eqref{item:product_varifold:integral},
\eqref{item:product_varifold:variation_measure}, and
\cite[8.6]{snulmenn:tv.v2} and, by \eqref{item:product_varifold:tv_again},
there holds, for $\| W \|$ almost all $(z_1,z_2)$,
\begin{gather*}
\derivative{W}{f} (z_1,z_2) (u_1,u_2) = L ( \derivative{V_1}{f_1}
(z_1)(u_1), \derivative{V_2}{f_2} (z_2)(u_2) )
\end{gather*}
for $u_1 \in \mathbf{R}^{{n}_1}$ and $u_2 \in \mathbf{R}^{{n}_2}$. Recalling
\eqref{item:product_varifold:rectifiable}, the equation
for the second fundamental form of $W$ now follows.
\end{proof}
\begin{remark}
\label{remark:product_rectifiable_sets}
The behaviour of rectifiable varifolds described in
\eqref{item:product_varifold:rectifiable} is in contrast with the more
subtle behaviour of $(\mathscr{H}^{m},{m})$ rectifiable sets; in fact,
if ${m}, {n} \in \mathscr{P}$ and ${m} < {n}$ there exist compact subsets
$C_1$ and $C_2$ of $\mathbf{R}^{n}$ such that
\begin{gather*}
\mathscr{H}^{m} ( C_1 ) = \mathscr{H}^{m} (C_2) =0, \quad
\mathscr{H}^l ( C_1 \times C_2 ) = \infty
\end{gather*}
whenever $0 \leq l < {m} + {n}$, see \cite[2.10.29,
3.2.24]{MR41:1976}.
\end{remark}
\begin{remark}
Concerning \eqref{item:product_varifold:tv_again} and
\eqref{item:product_varifold:curvature}, notice that, for general $W$,
neither is membership in $\mathbf{T} (W, Y_1 \times Y_2 )$ implied by membership
of the component functions in $\mathbf{T} (W,Y_i)$ nor is $\mathbf{T} (W, Y_1 \times
Y_2 )$ closed under addition, see \cite[8.25]{snulmenn:tv.v2}.
\end{remark}
|
\section{Introduction}\label{sec:intro}
We consider the problem in Figure~\ref{fig:mac}, in which $K$ users design matrices $B_1,\ldots,B_K$, where $B_i$ is an $n \times m_i$ matrix, designed by the $i^{\textrm{th}}$ user, with full column rank $m_i \leq n$. A destination obtains the matrix $B_D=\left[\Lambda_1 B_1 \quad \Lambda_2 B_2 \quad \cdots \quad \Lambda_K B_K \right]$, where $\Lambda_i$ is an $n \times n$ diagonal matrix with random diagonal entries, and the set of all random coefficients is drawn from a continuous joint distribution.
It is easy to see that $R=\min\left(\sum_{i=1}^K m_i, n\right)$ is the maximum rank that the matrix $B_D$ can have. However, by a careful design, one can reduce the rank of this matrix. In particular, the question that we address in this work is: Under what conditions on the design of the matrices $B_1,\ldots,B_K$ will the matrix $B_D$ lose rank by $\tau$ almost surely, i.e., $\text{rank}(B_D) \overset{a.s.}{\leq} R - \tau$, while each of the matrices $B_1,\ldots,B_K$ has full column rank?
\begin{figure}[h]
\centering
\includegraphics[trim = 2.71in 2.95in 1.82in 2.81in, clip,width=0.4\textwidth]{MAC}
\caption{Multiple users, each one having a matrix $B_i$ and another user receiving a concatenation of these matrices with randomly-scaled rows.}
\label{fig:mac}
\end{figure}
The aforementioned problem arises naturally when understanding the fundamentals of interference management in wireless networks, in particular for the case where no information about the channel state is assumed to be available at the transmitters except for the knowledge about the network topology. In this setting, the $K$ users in Figure \ref{fig:mac} represent transmitters that are interfering at a single receiver. The matrix $B_i$ represents the beamforming matrix at the $i^{\textrm{th}}$ transmitter and the $j^{\textrm{th}}$ diagonal element of the matrix $\Lambda_i$ represents the channel coefficient in the $j^{\textrm{th}}$ time slot. The column span of the matrix $B_D$ represents the space of the received interference. The aforementioned problem is that of determining the conditions on the beamforming matrices that result in an alignment of the interference caused by users $1$ to $K$ in a subspace whose dimension is at most $R-\tau$.
Another motivating example arises in recommendation systems, where for a fixed set of items, each matrix $B_i$ represents the ratings given by the $i^{\textrm{th}}$ group of users.
Each row represents one user's ratings for the set of items, and each diagonal element of $\Lambda_i$ represents a random scaling factor that reflects the user's own bias. For example, one user can give a rating of $10$ to his most favorite movie and $5$ for his least favorite while another user can have a highest rating of $8$ and a lowest rating of $4$. The problem in this setting would be to understand conditions on the ratings of each user group that result in a rank loss of the ensemble. Understanding conditions of rank loss in this application is useful for completion algorithms that recover missing ratings. In other words, the answer of our question in this setting specifies the structure of matrix entries which suffice to complete the entire rating matrix.
As a special case of this problem, one can consider each $B_i$ being only a column vector. This is equivalent to the case where instead of each row, each individual element is scaled by a random coefficient. In this case, the problem can be shown to have a combinatorial structure, depending on the position of zero/non-zero elements of $B_i$'s. In particular, it can be verified that in this case, the ensemble matrix $B_D$ will lose rank by $\tau$ (i.e., $\text{rank}(B_D)\leq R-\tau$) if and only if there does not exist a matching of size greater than $R-\tau$ between rows and columns of $B_D$, where a row is connected to a column if and only if the element with the corresponding row and column indices is non-zero (see e.g.,\cite{path-matching,lovasz,factor} and also the Zippel-Schwartz Lemma \cite{sch,zip}). However, once $B_i$'s are not column vectors, the problem becomes much more complicated since there is a structure in the random scaling; elements of the same row in each matrix are scaled by the same coefficient.
In this paper, we characterize a necessary and sufficient condition for rank-loss of the matrix $B_D$. In particular, we determine under what condition on $B_i$'s, the matrix $B_D$ loses rank by $\tau$ almost surely. The proof consists of a succession of equivalence steps. In the first step, we connect the rank of $B_D$ to the rank of the union of certain matroids. This is first done by expanding the determinants of specific submatrices of $B_D$ and then using the Zippel-Schwartz Lemma to express the almost-sure rank loss condition of $B_D$ in terms of the products of the determinants of certain submatrices of $B_i$'s being equal to zero. Then we use the notion of \emph{sparse subspaces} in order to connect the rank loss of $B_D$ to the column span of $B_i$'s and then to the rank loss of the union of carefully-defined matroids. After representing the rank condition of $B_D$ in terms of the rank loss of the union of certain matroids, in the second step, we make use of the matroid union theorem. This theorem enables us to derive an equivalent condition on the structure of $B_1,...,B_K$ for the rank loss of the union of matroids.
In the final step of the proof, we simplify the above condition by constructing appropriate bipartite graphs in which one partite set represents the column vectors of a carefully chosen set of bases for the column span of each of the matrices $B_1,...,B_K$ and the other partite set represents the row indices. A column vector vertex is connected to a row vertex if and only if the vector has a non-zero entry in the corresponding row. We use a variation of the Hall's marriage theorem to reach a final condition expressed through matching sizes on the constructed graphs.
As an application of the derived rank-loss condition, we utilize the result in the context of the topological interference management problem. This problem focuses on the scenario in which interference management primarily relies on a coarse knowledge about channel states in the network, namely the ``topology'' of the network. Network topology simply refers to the 1-bit feedback information for each link between each transmitter and each receiver, indicating whether or not the signal of the transmitter is received above the noise floor at the corresponding receiver.
There have been several prior works in the literature that have considered similar scenarios. In~\cite{localview}, it is assumed that transmitters are aware of the network topology as well as the actual channel gains within a local neighborhood; the concept of normalized capacity was introduced by characterizing the maximum achievable rate using this channel knowledge as a fraction of the achievable rate using global channel state information. In \cite{jafar}, the authors considered a more restrictive scenario, called topological interference management (TIM), in which the transmitters are only aware of the topology with no information about the channel coefficients. It has been shown that if the channel gains in the network remain constant for a sufficiently large time, then topological interference management is closely connected to the classical index coding problem (see, e.g., \cite{birk,baryossef,rouayheb,local_coloring_index_coding}), and via this connection, a class of linear interference management schemes has been introduced which rely on the network topology knowledge to align the interference over time at unintended receivers. In~\cite{topology,topology_isit}, the authors considered a class of retransmission-based schemes that only allows transmitters to resend their symbols in order to assist with the neutralization of interference at the receivers. Besides their simplicity, these schemes are robust to channel variations over time and were shown to be optimal in terms of the symmetric degrees-of-freedom (DoF) in many classes of topologies via the outer bounds developed in \cite{topology,topology_isit}.
In this paper, we use our derived rank loss condition in order to solve two problems in the topological interference management framework. In this setting, Figure~\ref{fig:mac} can be seen as a set of $K$ transmitters causing interference at a destination, and the considered rank-loss problem can be seen as reducing the dimension of the subspace occupied by interference at the receiver through careful design of the beamforming matrices at the transmitters.
We first characterize the ``best'' topologies for topological interference management; i.e., the topologies for which half symmetric DoF is achievable for the case of time-varying channels. It is easy to see that in topologies with at least one interference link, the symmetric DoF is upper bounded by $\frac{1}{2}$. Thus, the topologies in which half symmetric DoF is achievable represent the best topologies that one can hope for (from the degrees-of-freedom perspective). For the case where the channel gains in the network are assumed to remain constant for a long-enough period of time, the necessary and sufficient condition on the network topology for achieving half symmetric DoF was characterized in \cite{jafar}. For the case of case of time-varying channels, a sufficient condition for the achievability of half symmetric DoF was derived in \cite{multialign}. In this work, we close the gap in the results on half symmetric DoF by introducing a necessary and sufficient condition under which the symmetric DoF of $\frac{1}{2}$ is achievable for the case of time-varying channels (i.e., without requiring the channels to remain fixed for a long enough time).
Second, we use our condition to characterize the linear symmetric degrees of freedom (DoF) for a class of network topologies with exclusive interference sets. This helps to resolve the characterization of the symmetric DoF for a set of previously open problems considered in~\cite{topology}.
\section{Problem Formulation and Main Result}\label{sec:model}
Consider $K$ matrices $B_1,\ldots,B_K$, where each matrix $B_i, i\in[K]$ has size $n\times m_i$, $m_i\leq n$ (we use $[K]$ to denote the set $\{1,\ldots,K\}$ for any positive integer $K$). Without loss of generality, we assume that each matrix is full-column rank, since the linearly-dependent columns can be removed from each matrix. Furthermore, consider $K$ diagonal matrices $\Lambda_1,\ldots,\Lambda_K$, each of size $n\times n$, where their diagonal elements are drawn from a joint continuous distribution.
In this paper, the problem under consideration is the rank loss of the matrix $B_D=\left[\Lambda_1 B_1 \quad \Lambda_2 B_2 \quad \cdots \quad \Lambda_K B_K \right]$. To be precise, we aim to find an equivalent condition for when
\begin{align}\label{eq:rankloss}
\text{rank} ([\Lambda_1 B_1 ~~\Lambda_2 B_2~~ \ldots ~~ \Lambda_K B_K])\overset{a.s.}{\leq} R -\tau,
\end{align}
where $R=\min\left(\sum_{i=1}^K m_i,n\right)$ denotes the maximum possible rank of $[\Lambda_1 B_1 ~~\Lambda_2 B_2~~ \ldots ~~ \Lambda_K B_K]$ and $\tau\in \mathbb{Z}^+$.
\textbf{Notation:} For a matrix $B\in\mathbb{R}^{n\times m}$, we use calligraphic $\mathcal{B}$ to denote the subspace in $\mathbb{R}^n$ spanned by the columns of $B$. Also, for any $X\subseteq[n]$ and $Y\subseteq[m]$, $B_{X,Y}$ denotes the submatrix of $B$ created by removing the rows with indices outside $X$ and removing the columns with indices outside $Y$, and $B_{*,Y}$ denotes the submatrix of $B$ created by removing the columns with indices outside $Y$. Besides, for any $X\subseteq[n]$, $P_X^K$ denotes the set of all partitions of $X$ to $K$ disjoint subsets (each one possibly empty). Finally, for any $J\subseteq [n]$, $J^c$ denotes the complement of $J$ in $[n]$ (i.e., $[n]\setminus J$) and $\mathcal{S}_J$ denotes the subspace of $\mathbb{R}^n$ spanned by the columns of the $n\times n$ identity matrix with indices in $J$. In other words, $\mathcal{S}_J$ is the subspace of $\mathbb{R}^n$ which includes all the vectors that have zero entries in $J^c$. We call $\mathcal{S}_J$ the \emph{sparse subspace} of the set $J$.
We now state the main result in the following theorem.
\begin{thm}\label{thm:main}
The following two statements are equivalent.
\begin{gather}
\emph{rank} ([\Lambda_1 B_1 ~~\Lambda_2 B_2~~ ... ~~ \Lambda_K B_K])\overset{a.s.}{\leq}R -\tau.\tag{C1}\label{eq:original}\\
\forall Y_i\subseteq[m_i], i \in [K] \emph{ s.t. }\sum_{i=1}^K |Y_i|=R, \exists J \subseteq [n]: \sum_{i=1}^K \dim(\mathcal{S}_{J} \cap \mathcal{B}_{i,*,Y_i}) \geq |J|+\tau.\tag{C2}\label{eq:result}
\end{gather}
\end{thm}
\begin{ex}\label{ex:result_clr_1}
As a simple example, consider the case where $K=2$ and $m_1=m_2=\frac{n}{2}$. In this case, $B_1$ and $B_2$ will each be of size $n\times \frac{n}{2}$ and the only possible choice for $Y_1$ and $Y_2$ will be $[\frac{n}{2}]$. Theorem \ref{thm:main} implies that $\text{rank} ([\Lambda_1 B_1 ~~\Lambda_2 B_2])\overset{a.s.}{\leq}n -\tau$ if and only if there exists a set $J\subseteq[n]$ such that,
\begin{align}\label{eq:example}
\dim(\mathcal{S}_{J} \cap \mathcal{B}_{1}) + \dim(\mathcal{S}_{J} \cap \mathcal{B}_{2}) \geq |J|+\tau.
\end{align}
\end{ex}
\begin{ex}
Continuing Example \ref{ex:result_clr_1}, assume $n=2m_1=2m_2=4$ and suppose we fix $B_1=\begin{bmatrix}
1 & 1 & 1 & 0\\
1 & 2 & 3 & 0
\end{bmatrix}^T$.
In this case, Theorem \ref{thm:main} implies that the only way for the matrix $[\Lambda_1 B_1 ~~\Lambda_2 B_2]$ to lose rank by $\tau=1$ almost-surely is that both columns of $B_2$ have zeros in their $4^{th}$ entries. The set $J\subseteq [4]$ satisfying (\ref{eq:example}) would be $\{1,2,3\}$ in this case.
\end{ex}
\begin{remk}
The significance of the above result lies in finding a condition on the structure of the matrices $B_i, i\in[K]$, such that the statistical rank loss condition is met. Checking the structural condition of~\eqref{eq:result} does not involve statistical analysis and relies only on the combinatorial structure of $\mathcal{B}_i$'s.
\end{remk}
\begin{remk}
When each element of the matrices $B_i,\ldots,B_K$ is scaled by a random coefficient,~\eqref{eq:original} corresponds to the size of a bipartite graph matching representing rows and columns of $B_D$ as the two partite sets. This condition will be similar to~\eqref{eq:result}, but instead of $\dim(\mathcal{S}_J \cap \mathcal{B}_i)$, the number of columns in $B_i$ inside the subspace $\mathcal{S}_J$ is considered. This is intuitive as in the considered setting, each row is scaled by the same coefficient, and hence, the angles between column vectors are preserved.
Our result can be seen as a generalization of the classical rank-loss results for random matrices.
\end{remk}
\begin{remk}
We show in Section~\ref{sec:tim} how the result of Theorem~\ref{thm:main} can be used to characterize the linear degrees of freedom for the topological interference management problem of a class of networks that had been considered as an open problem before.
\end{remk}
\section{Proof of the Main Result}\label{sec:result}
The proof of Theorem~\ref{thm:main} is composed of four steps, in each of which we present a condition equivalent to condition (\ref{eq:original}).
\subsection{Step 1: Expansion of Determinant}\label{sec:proofone}
We begin by the following equivalence lemma.
\begin{lem}\label{lem:expansion}
Condition (\ref{eq:original}) is equivalent to the following.
\begin{gather}
\forall X\subseteq[n]: |X|>R-\tau,
\forall Y_i\subseteq[m_i], i\in[K] \text{ s.t. }\sum_{i=1}^K |Y_i|=|X|, \nonumber\\
\forall (I_1,I_2,...,I_K)\in P_X^K \text{ s.t. }|I_i|=|Y_i|:
\prod_{i=1}^K \det(B_{i,I_i,Y_i})=0.\tag{C3}\label{eq:condone}
\end{gather}
\end{lem}
\begin{proof}
By the definition of matrix rank, the condition in~\eqref{eq:original} is equivalent to the fact that any square submatrix of $B_D$ with size greater than $(R-\tau) \times (R-\tau)$ should have a zero determinant almost-surely. This means that for any subset of rows $X\subseteq[n]: |X| > R-\tau$ and any subsets of columns $Y_i\subseteq[m_i]$, $i\in[K] \text{ s.t. }\sum_{i=1}^K |Y_i|=|X|$,
\begin{align*}
&\det\left([\Lambda_{1,X,X} B_{1,X,Y_1} ~~\ldots~~\Lambda_{K,X,X} B_{K,X,Y_K}]\right)\overset{a.s.}{=}0.
\end{align*}
It can be shown that this determinant is composed of monomials in the channel gains whose coefficients are in the form of $\prod_{i=1}^K \det(B_{i,I_i,Y_i})$ for some $(I_1,I_2,...,I_K)\in P_X^K \text{ s.t. }|I_i|=|Y_i|$. By the Zippel-Schwartz Lemma \cite{zip,sch}, for the whole multivariate polynomial to be equal to zero for almost all values of the channel gains, each of these coefficients should be equal to zero, which gives~\eqref{eq:condone}.
\end{proof}
\begin{ex}
Consider the case where $K=2$, $n=3$, $m_1=1$, $m_2=2$, and we have
$B_1=\begin{bmatrix}
a_{11} & a_{21} & a_{31}
\end{bmatrix}^T,
B_2=\begin{bmatrix}
a'_{11} & a'_{21} & a'_{31}\\a'_{12} & a'_{22} & a'_{32}
\end{bmatrix}^T.$
Also, let $\lambda_i$, $\lambda'_i$, $i \in \{1,2,3\}$, be the $i^{\textrm{th}}$ diagonal element of $\Lambda_1$ and $\Lambda_2$, respectively.
We then have,
\begin{align*}
B_D=\begin{bmatrix}
\lambda_1 a_{11} & \lambda'_1 a'_{11} & \lambda'_1 a'_{12} \\ \lambda_2 a_{21} & \lambda'_2 a'_{21} & \lambda'_2 a'_{22} \\ \lambda_3 a_{31} & \lambda'_3 a'_{31} & \lambda'_3 a'_{32}
\end{bmatrix}.
\end{align*}
Now, $\text{rank}(B_D)\leq 2$ is equivalent to $\det(B_D)=0$. Note that,
\begin{align*}
\det(B_D)&=\lambda_1 \lambda'_2 \lambda'_3 \left[a_{11} \det\left(\begin{bmatrix}
a'_{21} & a'_{22}\\a'_{31} & a'_{32}
\end{bmatrix}\right)\right]\\
&-\lambda_2 \lambda'_1 \lambda'_3 \left[a_{21} \det\left(\begin{bmatrix}
a'_{11} & a'_{12}\\a'_{31} & a'_{32}
\end{bmatrix}\right)\right]\\
&+\lambda_3 \lambda'_1 \lambda'_2 \left[a_{31} \det\left(\begin{bmatrix}
a'_{11} & a'_{12}\\a'_{21} & a'_{22}
\end{bmatrix}\right)\right].
\end{align*}
Therefore, the Zippel-Schwartz Lemma implies that this determinant is almost-surely equal to zero if and only if each of the products inside the brackets is equal to zero.
\end{ex}
\subsection{Step 2: Sparse Subspaces}\label{sec:prooftwo}
We now express the condition~\eqref{eq:condone} in terms of sparse subspaces. Recall that a sparse subspace is defined by column vectors that have zero entries for a specific set of rows.
\begin{lem}\label{lem:sparse}
Condition (\ref{eq:condone}) is equivalent to the following.
\begin{gather}
\forall X\subseteq[n]: |X|>R-\tau,\nonumber\\
\forall Y_i\subseteq[m_i], i\in[K] \text{ s.t. }\sum_{i=1}^K |Y_i|=|X|, \nonumber\\
\forall (I_1,I_2,...,I_K)\in P_X^K \text{ s.t. }|I_i|=|Y_i|, \forall i\in[K]:\nonumber\\
\sum_{i=1}^K \dim(\mathcal{S}_{I_i^c} \cap \mathcal{B}_{i,*,Y_i})>0.\tag{C4}\label{eq:condtwo}
\end{gather}
\end{lem}
\begin{proof}
First assume that the condition in~\eqref{eq:condone} holds. Then for some $i\in[K]$ there exists a linear combination of the columns in $B_{i,I_i,Y_i}$ which is equal to the zero vector. Therefore, if we apply this linear combination to the entire matrix $B_{i,*,Y_i}$, we end up with a vector in the sparse subspace $\mathcal{S}_{{I_i}^c}$ since $B_{i,*,Y_i}$ is full-column rank. This implies~\eqref{eq:condtwo}.
If the condition in~\eqref{eq:condtwo} holds, then for some $i\in[K]$, there exists a linear combination of the columns in $B_{i,*,Y_i}$ which is zero in the coordinates in $I_i$. This means that for any $I_i: |I_i|=|Y_i|$, $\det (B_{i,I_i,Y_i})=0$, implying that~\eqref{eq:condone} holds.
\end{proof}
\subsection{Step 3: Connection to Rank of Union of Matroids}\label{sec:proofthree}
In this step, we show how to represent condition (\ref{eq:condtwo}) in terms of the rank loss of the union of certain matroids. To this end, for any choice of $X\subseteq[n]$ and $Y_i\subset[m_i],~\forall i \in [K]$ which satisfy $\sum_{i=1}^K |Y_i|=|X|>R-\tau$ and
\begin{align}\label{eq:matroiddefcondition}
\dim(\mathcal{S}_{X^c} \cap \mathcal{B}_{i,*,Y_i} )=0,\forall i\in[K],
\end{align}
define $\mathcal{I}_{i,X,Y_i}$ as
\begin{align}
\mathcal{I}_{i,X,Y_i}=\{I\subseteq X:\dim(\mathcal{S}_{I\cup X^c}\cap \mathcal{B}_{i,*,Y_i})=0\}, i\in[K].
\end{align}
We have the following claim that is proved in Appendix~\ref{sec:matroid_proof}.
\begin{claim}\label{matroid}
$M_{i,X,Y_i}=(X,\mathcal{I}_{i,X,Y_i})$ is a matroid with rank function $r_{i,X,Y_i}(J)=|J|-\dim (\mathcal{S}_{J\cup X^c}\cap \mathcal{B}_{i,*,Y_i}) $, $i\in[K]$.
\end{claim}
For any matroid $M_{i,X,Y_i}$, the dual matroid $M_{i,X,Y_i}^*$ is defined as $M_{i,X,Y_i}^*=(X,\mathcal{I}_{i,X,Y_i}^*)$ whose basis sets are the complements of the basis sets of $M_{i,X,Y_i}$ \cite{schrijver}.\footnote{For a matroid $M=(X,\mathcal{I})$, $J\subseteq X$ is called a \emph{basis set} if $J\in \mathcal{I}$ and there is no $J'\in\mathcal{I}$ such that $J\subset J' \subseteq X$ \cite[Ch. 39]{schrijver}.} Using Theorem 39.2 in \cite{schrijver} and Claim \ref{matroid}, the rank of this dual matroid is
\begin{align}
r_{i,X,Y_i}^*(J)&=|J|-r_{i,X,Y_i}(X)+r_{i,X,Y_i}(X\setminus J)\nonumber\\
&=|J|-(|X|-|Y_i|)+(|X\setminus J|-\dim (\mathcal{S}_{J^c}\cap \mathcal{B}_{i,*,Y_i}))\nonumber\\
&=|Y_i|-\dim (\mathcal{S}_{J^c}\cap \mathcal{B}_{i,*,Y_i}).\label{dualrank}
\end{align}
Now, consider the union of the dual matroids $M_{i,X,Y_i}^*$, $i\in[K]$, denoted by $\bigvee_{i=1}^K M_{i,X,Y_i}^*=\left(X,\bigvee_{i=1}^K \mathcal{I}_{i,X,Y_i}^*\right)$, where
\begin{align*}
\bigvee_{i=1}^K \mathcal{I}_{i,X,Y_i}^*=\left\{\bigcup_{i=1}^K I_i: I_i\in\mathcal{I}_{i,X,Y_i}^*\right\}.
\end{align*}
Let $r_{X,Y_1,...,Y_K}^*(.)$ denote the rank of $\bigvee_{i=1}^K M_{i,X,Y_i}^*$. Then we have the following lemma.
\begin{lem}\label{lem:matroidunion}
For any $X\subseteq [n], Y_i\subseteq[m_i], i\in[K]$ s.t. $\sum_{i=1}^K |Y_i|=|X|> R -\tau$ and (\ref{eq:matroiddefcondition}) is satisfied, the following are equivalent
\begin{gather}
\forall (I_1,I_2,...,I_K)\in P_X^K \text{ s.t. }|I_i|=|Y_i|:
\sum_{i=1}^K \dim(\mathcal{S}_{I_i^c} \cap \mathcal{B}_{i,*,Y_i})>0.\label{eq:c4prime}
\end{gather}
\begin{gather}
r_{X,Y_1,...,Y_K}^*(X)<|X|.\label{eq:condunionmatroid}
\end{gather}
\end{lem}
\begin{proof}
Condition~\eqref{eq:c4prime} is equivalent to the following for any valid choice of $X$ and $Y_i$'s:
\begin{gather}
\nexists (I_1,I_2,...,I_K)\in P_X^K:|I_i|=|Y_i|\text{ and } X\setminus I_i \in \mathcal{I}_{i,X,Y_i}, \forall i\in[K].\label{equiv}
\end{gather}
Now, let us focus on a specific $i\in[K]$ and the corresponding set $I_i$ whose size satisfies $|I_i|=|Y_i|$,
\begin{align}
|X\setminus I_i|&=|X|-|Y_i|\nonumber\\
&=|X|-\dim(\mathcal{S}_{X\cup X^c}\cap \mathcal{B}_{i,*,Y_i})\nonumber\\
&=r_{i,X,Y_i}(X),\label{eq:rank}
\end{align}
where \eqref{eq:rank} follows from Claim \ref{matroid}. By the definition of the rank of a matroid (see, e.g. \cite[Ch. 39]{schrijver}), \eqref{eq:rank} implies that all the members of $\mathcal{I}_{i,X,Y_i}$ (i.e., the independent sets of $M_{i,X,Y_i}$) are of size at most $|X\setminus I_i|$. This means that if $X\setminus I_i\in \mathcal{I}_{i,X,Y_i}$, then it is a basis for $M_{i,X,Y_i}$. This, in turn, is equivalent to $I_i$ being a basis for the dual matroid $M_{i,X,Y_i}^*$.
On the other hand, from (\ref{dualrank}) we have that the rank of the dual matroid $M_{i,X,Y_i}^*$ is upper bounded by $|Y_i|$, implying that all the members of $\mathcal{I}_{i,X,Y_i}^*$ have size at most $|Y_i|$. Thus, under the constraint $|I_i|=|Y_i|$, $I_i$ being a basis for the dual matroid $M_{i,X,Y_i}^*$ is equivalent to $I_i\in \mathcal{I}_{i,X,Y_i}^*$. Consequently, \eqref{equiv} (hence ~\eqref{eq:c4prime}) is equivalent to the following for any valid choice of $X$ and $Y_i$'s:
\begin{gather}
\nexists (I_1,I_2,...,I_K)\in P_X^K:|I_i|=|Y_i|\text{ and } I_i \in \mathcal{I}_{i,X,Y_i}^*, \forall i\in[K].\label{equiv2}
\end{gather}
Finally, it is easy to verify that (\ref{equiv2}) is equivalent to $r_{X,Y_1,...,Y_K}^*(X)<|X|$, since there cannot exist any $(I_1,I_2,...,I_K)\in P_X^K$ such that $I_i\in \mathcal{I}_{i,X,Y_i}^*$ at the same time for all $i\in[K]$. This completes the proof.
\end{proof}
Note that using this lemma, we can now replace the last two lines of condition~\eqref{eq:condtwo} by simply $r_{X,Y_1,...,Y_K}^*(X)<|X|$.
\subsection{Step 4: Matroid Union Theorem}\label{sec:proofunionmatroid}
In this step of the proof, we make use of the Matroid Union Theorem \cite{schrijver} to characterize an equivalent condition to (\ref{eq:condunionmatroid}). We start by stating the Matroid Union Theorem.
\begin{thm}(Matroid Union Theorem \cite[Chapter 42]{schrijver})\label{mut}
Let $M_{i}=(E_{i},\mathcal{I}_{i})$, $i\in[K]$, be $K$ matroids with rank functions $r_i(.)$. Then $\bigvee_{i=1}^K M_{i}$ is also a matroid with rank function
\begin{align}
r(U)=\min_{T\subseteq U} \left(|U\setminus T|+\sum_{i=1}^K r_{i}(T\cap E_{i})\right).
\end{align}
\end{thm}
Equipped with the above theorem, we can now state the following equivalence lemma.
\begin{lem}
Condition (\ref{eq:condtwo}) is equivalent to the following.
\begin{gather}
\forall X\subseteq[n]: |X|>R-\tau,\nonumber\\
\forall Y_i\subseteq[m_i], i\in[K] \text{ s.t. }\sum_{i=1}^K |Y_i|=|X|, \nonumber\\
\exists J\subseteq X:
\sum_{i=1}^K \dim (\mathcal{S}_{J\cup X^c}\cap \mathcal{B}_{i,*,Y_i})>|J|.\tag{C5}\label{eq:condthree}
\end{gather}
\end{lem}
\begin{proof}
If for a specific choice of $X$ and $Y_i$'s, (\ref{eq:matroiddefcondition}) is not satisfied (i.e., $\dim(\mathcal{S}_{X^c} \cap \mathcal{B}_{i,*,Y_i} )>0$ for some $i\in[K]$), then it is clear that both statements in~\eqref{eq:condtwo} and~\eqref{eq:condthree} hold. Hence, w.l.o.g. we assume that~\eqref{eq:matroiddefcondition} is satisfied. For any $X\subseteq [n], Y_i\subseteq[m_i], i\in[K] \text{ s.t. }\sum_{i=1}^K |Y_i|=|X|> R -\tau$ and (\ref{eq:matroiddefcondition}) is satisfied, we will get from Lemma~\ref{lem:matroidunion} that the condition in~\eqref{eq:condtwo} is equivalent to,
\begin{gather*}
r_{X,Y_1,...,Y_K}^*(X)<|X|\nonumber\\
\overset{(a)}{\Leftrightarrow}\underset{J\subseteq X}{\min} \left((|X|-|J|)+\sum_{i=1}^K (|Y_i|-\dim (\mathcal{S}_{J^c}\cap \mathcal{B}_{i,*,Y_i}))\right)<|X|\nonumber\\
\Leftrightarrow |X|+\sum_{i=1}^K |Y_i|-\underset{J\subseteq X}{\max} \left(|J|+\sum_{i=1}^K \dim (\mathcal{S}_{J^c}\cap \mathcal{B}_{i,*,Y_i})\right)<|X|\nonumber\\
\overset{(b)}{\Leftrightarrow} ~\max_{J\subseteq X} \bigg(|J|+\sum_{i=1}^K \dim (\mathcal{S}_{J^c}\cap \mathcal{B}_{i,*,Y_i})\bigg)>|X|\\
\Leftrightarrow ~\exists J\subseteq X : \sum_{i=1}^K \dim (\mathcal{S}_{J^c}\cap \mathcal{B}_{i,*,Y_i})>|X\setminus J|\\
\overset{(c)}{\Leftrightarrow} ~\exists J\subseteq X : \sum_{i=1}^K \dim (\mathcal{S}_{J\cup X^c}\cap \mathcal{B}_{i,*,Y_i})>|J|,
\end{gather*}
where (a) follows from (\ref{dualrank}) and Theorem \ref{mut}, (b) follows from the assumption that $\sum_{i=1}^K |Y_i|=|X|$ and (c) follows by changing $J$ to $X\setminus J$. This completes the proof.
\end{proof}
\subsection{Step 5: Hall's Marriage Theorem}\label{sec:prooffour}
In the final step of the proof, we prove the equivalence between~\eqref{eq:condthree} and~\eqref{eq:result} by constructing an appropriate bipartite graph. One partite set represents the column vectors of a carefully chosen set of bases for the subspaces $\mathcal{B}_{i,*,Y_i}$ and the other partite set has $n$ elements, each corresponding to a row. A column vector vertex is connected to a row vertex if and only if the vector has a non-zero entry in the corresponding row. We then use the following variation of Hall's marriage theorem to complete the proof by representing the final condition~\eqref{eq:result} in terms of a matching size on the constructed bipartite graph.
\begin{thm}\label{thm:hull}
Let $G=(A\cup B,E)$ be a bipartite graph. $G$ has a matching of size $k$ if and only if the size of the neighboring set $|N(I)|\geq |I|-|B|+k$ for any $I\subseteq B$.
\end{thm}
We first show that \eqref{eq:result} $\Rightarrow$ \eqref{eq:condthree}. Assume that \eqref{eq:result} holds and fix $Y_i \subseteq [m_i], i\in[K]: \sum_{i=1}^K |Y_i|=R$. Let $J^* \subseteq [n]$ be such that $\sum_{i=1}^K \dim(\mathcal{S}_{J^*} \cap \mathcal{B}_{i,*,Y_i}) \geq |J^*|+\tau$. For each $i \in [K]$, let $c_1^{(i)},...,c_{|Y_i|}^{(i)}$ be $|Y_i|$ column vectors that form a basis for $\mathcal{B}_{i,*,Y_i}$ and a subset of these vectors form a basis for $\mathcal{S}_{J^*} \cap \mathcal{B}_{i,*,Y_i}$. Now, let $C_i=\left\{c_1^{(i)},...,c_{|Y_i|}^{(i)}\right\}$ and construct a bipartite graph $G=(A \cup C,E)$ where $A=\{1,2,\ldots,n\}$ and $C=\{C_1,\ldots,C_K\}$ is a multiset consisting of the elements of the sets $C_i, i\in [K]$. $A$ and $C$ are the two partite vertex sets of the graph $G$. For any $i\in [n]$ and $c\in C$, $(i,c)\in E$ if and only if the vector $c$ has a non-zero entry in the $i^{\textrm{th}}$ position, i.e., if $c=(c_1,\ldots,c_n)$ then $(a_i,c)\in E \Leftrightarrow c_i \neq 0$.
Now, note that all the vertices in $C$ that correspond to vectors in $\mathcal{S}_{J^*}$ can only be connected to vertices in $A$ that correspond to the set $J^*$. Since $\sum_{i=1}^K \dim(\mathcal{S}_{J^*} \cap \mathcal{B}_{i,*,Y_i}) \geq |J^*|+\tau$, it follows that there is a subset of the partite set $C$ of size at least $|J^*|+\tau$ whose neigboring set has size at most $|J^*|$. It follows from Theorem~\ref{thm:hull} that there is no matching of size $R-\tau+1$ in the bipartite graph $G$. It follows that for any $X \subseteq [n]: |X| > R-\tau$ and any $K$ sets $Y'_i \subseteq C,i \in [K]: \sum_{i=1}^K |Y'_i|=|X|$, there is no matching between the vertices corresponding to the set $X$ in the partite set $A$ and the vertices in $\{Y'_1,\ldots,Y'_K\}$, and hence, there is a set $J \subseteq X$ such that the vertices corresponding to the set $X \backslash J$ are connected to less than $|X \backslash J|$ vertices in $\bigcup_{i=1}^K Y'_i$. It follows that for this choice of the set $J$, there are more than $|J|$ vertices in $\bigcup_{i=1}^K Y'_i$ that are not connected to any vertex in the set $X \backslash J$ in the partite set $A$. Now, if $\forall i\in[K], Y'_i \subseteq Y_i$, then $\sum_{i=1}^K \dim(\mathcal{S}_{J \cup \bar{X}} \cap \mathcal{B}_{i,*,Y'_i}) >|J|$. We now use the above argument to prove that \eqref{eq:condthree} holds as follows. For all $ Y'_i\subseteq[m_i], i\in[K]$ such that $\sum_{i=1}^K |Y'_i| > R-\tau$, we find $Y_i, i\in[K]$ such that $Y'_i \subseteq Y_i, \forall i \in [K]$ and $\sum_{i=1}^K |Y_i|=R$, and then use the above argument to show that,
\begin{gather*}
\exists J \subseteq [n]: \sum_{i=1}^K \dim(\mathcal{S}_{J} \cap \mathcal{B}_{i,*,Y_i}) \geq |J|+\tau \Rightarrow\\ \forall X \subseteq [n]: |X| = \sum_{i=1}^K |Y'_i|, \\\exists J\subseteq X:
\sum_{i=1}^K \dim(\mathcal{S}_{J \cup \bar{X}} \cap \mathcal{B}_{i,*,Y'_i}) >|J|,
\end{gather*}
and hence, \eqref{eq:condthree} follows.
We now show that \eqref{eq:condthree} $\Rightarrow$ \eqref{eq:result} by contradiction. Suppose that \eqref{eq:result} does not hold and fix $Y_i \subseteq [m_i], i\in[K]: \sum_{i=1}^K |Y_i|=R$ such that for any $J\subseteq [n]$, $\sum_{i=1}^K \dim(\mathcal{S}_{J} \cap \mathcal{B}_{i,*,Y_i}) < |J|+\tau$. For each $i \in [K]$, let $c_1^{(i)},...,c_{|Y_i|}^{(i)}$ be $|Y_i|$ column vectors that form a basis for $\mathcal{B}_{i,*,Y_i}$ and define $C_i=\left\{c_1^{(i)},...,c_{|Y_i|}^{(i)}\right\}$. Now, we construct a bipartite graph $G=(A \cup C,E)$ where $A=\{1,2,\ldots,n\}$ and $C=\{C_1,\ldots,C_K\}$ is a multiset consisting of the elements of the sets $C_i, i\in [K]$. Also, for any $i \in [n], c\in C$, $(i,c) \in E \Leftrightarrow c_i \neq 0$.
From the selection of the sets $Y_i, i\in[K]$, we have that for any $I\subseteq C$, the neighboring set $N(I) \subseteq A$ has size $|N(I)| > |I|-\tau$. It follows from Theorem \ref{thm:hull} that there is a matching in $G$ of size $|C|-\tau+1$. Such a matching will include $R-\tau+1$ edges incident on $R-\tau+1$ nodes in $A$ (which we denote by $X$) and $R-\tau+1$ nodes in $C$ (which we denote by $Y'$). For each $i \in [K]$, let $Y'_i=\{j:c_j^{(i)}\in Y'\}$. Now, in order to show that \eqref{eq:condthree} is not true, it suffices to show that for the specific choice of $X$, $Y'_i, i\in[K]$ mentioned above, the following holds:
\begin{align}\label{contradict}
\forall J\subseteq X:
\sum_{i=1}^K \dim(\mathcal{S}_{J \cup X^c} \cap \mathcal{B}_{i,*,Y'_i}) \leq |J|
\end{align}
Fix any $J\subseteq X$. We know that there exist $|X\setminus J|$ nodes in $Y'$ that are connected to the vertices corresponding to elements in $X\setminus J$ through the edges in the matching. This implies that there exist $|X\setminus J|$ columns in $\left[B_{1,*,Y'_1}~\ldots~B_{K,*,Y'_K}\right]$ which have at least one non-zero entry in $X\setminus J$ and therefore cannot belong to $\mathcal{S}_{J\cup X^c}$. This means that $\sum_{i=1}^K \dim(\mathcal{S}_{J \cup X^c} \cap \mathcal{B}_{i,*,Y'_i})$ cannot be greater than the dimension of the span of the remaining columns, which is at most $\left(\sum_{i=1}^K |Y'_i|\right)-|X\setminus J|=|J|$, verifying (\ref{contradict}). Hence \eqref{eq:condthree} does not hold.
\section{Application to Topological Interference Management}\label{sec:tim}
\noindent In this section, we study an application of Theorem~\ref{thm:main} to characterize the linear symmetric degrees of freedom for a class of topological interference management problems as defined next.
\subsection{Topological Interference Management: System Model and Problem Overview}
We consider $K$-user interference networks composed of $K$ transmitter nodes $\{\text{T}_i\}_{i=1}^K$ and $K$ receiver nodes $\{\text{D}_i\}_{i=1}^K$. Each transmitter $\text{T}_i$ intends to deliver a message $W_i\in\mathcal{W}_i$ to its corresponding receiver $\text{D}_i$. We assume that each receiver is subject to interference only from a specific subset of the other transmitters and the interference power that it receives from the other transmitters is below the noise level. This leads to a \emph{network topology} indicating the network interference pattern.
Each transmitter $\text{T}_i$ intends to send a vector $\mathbf{w}_i\in\mathbb{R}^{m_i}$ of $m_i$ symbols to its desired receiver $\text{D}_i$ over $n$ time slots. This message is encoded to the transmit vector $\mathbf{x}_i=B_i \mathbf{w}_i$, where $B_i$ denotes the linear beamforming \emph{precoding} matrix of transmitter $i$, which is of size $n\times m_i$. The received signal of receiver $j$ over $n$ time slots is given by,
\begin{align*}
\mathbf{y}_j=(\Lambda_{jj} B_j) \mathbf{w}_j+\sum_{i\in I_j} (\Lambda_{ij} B_i) \mathbf{w}_i+\mathbf{z}_j,
\end{align*}
where $I_j$ is the set of transmitters interfering at receiver $j$, $\Lambda_{ij}$ is the $n\times n$ diagonal matrix with the $k^{\textrm{th}}$ diagonal element being equal to the value of the channel coefficient between transmitter $i$ and receiver $j$ in time slot $k$, and $\mathbf{z}_j$ is the noise vector at receiver $j$ where each of its elements is an i.i.d. $\mathcal{N}(0,N)$ random variable, $N$ being the noise variance. The channel gain values are assumed to be identically distributed and drawn from a continuous distribution at each time slot. We assume that transmitters have no knowledge about the realization of the channel gains except for the topology of the network. However, the receivers have full channel state information. We refer to this assumption as no CSIT (channel state information at the transmitters) beyond topology. Each precoding matrix $B_i$ is an $n \times m_i$ matrix that can only depend on the knowledge of topology.
At receiver $j$, the interference subspace denoted by $\mathcal{I}_j$ can be written as
\begin{align*}
\mathcal{I}_j=\bigcup_{i\in I_j} \mathsf{colspan}(\Lambda_{ij} B_i).
\end{align*}
In order to decode its desired symbols, receiver $j$ projects its received signal subspace given by $\mathsf{colspan} (\Lambda_{jj} B_j)$ onto the subspace orthogonal to $\mathcal{I}_j$, and its successful decoding condition can be expressed as
\begin{align}\label{eq:decodability}
\text{dim}\left(\mathsf{Proj}_{{\cal I}_{j}^c} \mathsf{colspan} \left(\Lambda_{jj} B_j\right)\right)=m_j.
\end{align}
If the above decodability condition is satisfied at all the receivers $\{\text{D}_j\}_{j=1}^K$ for almost all realizations of channel gains, then the linear degrees of freedom (LDoF) tuple $(\frac{m_1}{n},...,\frac{m_K}{n})$ is achievable under the aforementioned linear scheme. The linear symmetric degrees of freedom $\text{LDoF}_{sym}$ is defined as the supremum $d$ for which the LDoF tuple $(d,...,d)$ is achievable.
In this setting, the main goal is to characterize the linear symmetric degrees of freedom for general network topologies. There have been multiple attempts in the literature to resolve this problem. In particular, in \cite{jafar} it was shown that if the channel gains in the network remain constant for a sufficiently large time, then topological interference management is closely connected to the classical index coding problem (see, e.g., \cite{birk,baryossef,rouayheb}), and via this connection, a class of linear interference management schemes has been introduced which rely on the network topology knowledge to align the interference over time at unintended receivers. Furthermore, in~\cite{topology,topology_isit}, the authors considered a class of retransmission-based schemes that only allows transmitters to resend their symbols in order to assist with the neutralization of interference at the receivers. Besides their simplicity, these schemes are robust to channel variations over time and were shown to be optimal in terms of the symmetric degrees-of-freedom (DoF) in many classes of topologies via the outer bounds developed in \cite{topology,topology_isit}. More recently, the authors in \cite{avoidance_jafar} have characterized a necessary and sufficient condition for a topology under which interference avoidance can achieve the whole DoF region.
In the following two sections, we use our derived rank loss condition in order to address two problems in the topological interference management framework. First, in Section \ref{sec:best}, we characterize the ``best'' topologies for topological interference management; i.e., the topologies for which half linear symmetric DoF is achievable for the case of time-varying channels. For the case where the channel gains in the network are assumed to remain constant for a long-enough period of time, the necessary and sufficient condition on the network topology for achieving half linear symmetric DoF was characterized in \cite{jafar}. For the case of case of time-varying channels, a sufficient condition for the achievability of half linear symmetric DoF was derived in \cite{multialign}. In the next section, we close the gap by introducing a necessary and sufficient condition under which the half linear symmetric DoF is achievable for the case of time-varying channels (i.e., without requiring the channels to remain fixed for a long enough time). As for the second problem, in Section \ref{sec:nonoverlapping} we characterize the linear symmetric degrees of freedom for a class of network topologies with exclusive interference alignment sets. This helps to resolve the characterization of the linear symmetric DoF for a set of previously open problems considered in~\cite{topology}.
\subsection{Identifying the ``Best'' Topologies}\label{sec:best}
In this section, we use Theorem~\ref{thm:main} to identify network topologies where a linear symmetric DoF of $\frac{1}{2}$ is achievable. It is easy to see that in topologies where at least one transmitter is interfering at an undesired receiver, the symmetric DoF is upper bounded by $\frac{1}{2}$. Thus, the topologies in which half symmetric DoF is achievable represent the ``best'' topologies that one can hope for (from the degrees-of-freedom perspective).
We start by making the following definition of \emph{reduced conflict graphs}.
\begin{defn}\label{redgraph}
The reduced conflict graph of a $K$-user interference network is a directed graph $G=(V,A)$ with $V=\{1,2,...,K\}$. As for the edges, vertex $i$ is connected to vertex $j$ (i.e., $(i,j)\in A$) if and only if $i \neq j$ and the following two conditions hold:
\begin{enumerate}
\item Transmitter $i$ is connected to receiver $j$.
\item $\exists s,k \in \{1,2,\ldots,K\}\backslash\{i\}, s\neq k$ such that both transmitter $i$ and transmitter $s$ are connected to receiver $k$.
\end{enumerate}
\end{defn}
Clearly, the difference between the reduced conflict graph defined above and the regular conflict graph (used e.g. in \cite{topology}) is the additional condition 2 in Definition \ref{redgraph}, which implies that the set of edges in the reduced conflict graph is a subset of the edges in the regular conflict graph, hence the name ``reduced'' conflict graph.
Having the above definition, we characterize the best topologies in the following theorem.
\begin{thm}\label{thm:halfdof}
For a $K$-user interference network with arbitrary topology, half linear symmetric DoF can be achieved if and only if the reduced conflict graph of the network is bipartite.
\end{thm}
\begin{remk}\label{remk:avoidance}
It is easy to see that the interference avoidance scheme, which schedules the users in an independent set of the regular conflict graph of the network, can achieve half symmetric DoF if and only if the regular conflict graph of the network is bipartite. Since the reduced conflict graph is the same as the regular conflict graph with the removal of edges that do not satisfy condition $2$ in Definition~\ref{redgraph}, there exist topologies where the reduced conflict graph is bipartite but the regular conflict graph is not bipartite. One such example topology is illustrated in Figure \ref{ex_gain} together with its regular and reduced conflict graphs in Figure \ref{conf_graph}.
\begin{figure}[h]
\centering
\begin{subfigure}{0.3\textwidth}
\centering
\includegraphics[trim = 2.68in 2.3in 5.4in 2.2in, clip,width=0.7\textwidth]{ex_top_avoidance}
\caption{}
\label{ex_gain}
\end{subfigure}
~
\begin{subfigure}{0.3\textwidth}
\centering
\includegraphics[trim = 5.4in 2.3in 1.9in 2.2in, clip,width=0.9\textwidth]{ex_top_avoidance}
\caption{}
\label{conf_graph}
\end{subfigure}
\caption{(a) A topology in which half symmteric DoF is achievable, but not by interference avoidance, and (b) the corresponding regular and reduced conflict graph. The dashed red lines exist in the conflict graph and are absent in the reduced conflict graph.}
\end{figure}
The black edges are the edges in the reduced conflict graph and the dashed red edges are the ones that exist in the regular conflict graph but are removed in the reduced conflict graph. It is clear that the reduced conflict graph is bipartite (with the two partite sets being $\{1,3,4\}$ and $\{2,5,6\}$), hence $d_{sym}=\frac{1}{2}$ is achievable in this topology. However the addition of the dashed red edges in the regular conflict graph removes the bipartiteness property of the graph and therefore interference avoidance cannot achieve half symmetric DoF in this topology. In fact, since the chromatic number of the regular conflict graph in Figure \ref{conf_graph} is 3, interference avoidance can only achieve a symmetric DoF of $\frac{1}{3}$. The suboptimality of interference avoidance can also be seen as a result of the network topology in Figure \ref{ex_gain} not being chordal \cite{avoidance_jafar}.
\end{remk}
\begin{remk}
In \cite{jafar}, a linear scheme based on aligning the interference at unintended receivers was considered. The alignment scheme relies on the channel remaining constant for a long time (large coherence time). It was shown that half symmetric DoF is achievable if and only if there is no conflict between any two nodes that cause interference at a third receiver, i.e., there is no internal conflict within an alignment set. It is easy to show that the condition in Theorem \ref{thm:halfdof} implies the absence of internal conflicts but the opposite is not true; this is consistent with intuition as we make no assumption on the coherence time of the channel. Interestingly, it is shown in~\cite{ICC} that the topologies for which interference alignment can achieve half symmetric DoF while retransmission cannot achieve it comprise a negligible fraction of possible topologies in a heterogeneous network scenario of practical interest.
\end{remk}
\begin{proof}[Proof of Theorem \ref{thm:halfdof}]
We first prove the converse. In order to achieve $\frac{n}{2}$ symmetric DoF, there has to be a sequence $(\epsilon_n, n\in{\bf Z}^+)$, such that $\epsilon_n \rightarrow 0$ and $\frac{n}{2}-\epsilon_n$ symmetric DoF is achievable by coding over $n$ time slots. Consider some generic receiver $l$ in the network subject to interference from transmitters $\{\text{T}_i\}_{i=1}^K$. Then the decodability condition is equivalent to the dimension of the interference subspace being at most $\frac{n}{2}+\epsilon_n$ almost surely; i.e.,
\begin{align}\label{eq:halfdofrankloss}
\emph{rank} ([\Lambda_{1l} B_1 ~~\Lambda_{2l} B_2~~ ... ~~ \Lambda_{Kl} B_K])\overset{a.s.}{\leq} \frac{n}{2}+\epsilon_n.
\end{align}
Then, we know from Theorem~\ref{thm:main} that \eqref{eq:halfdofrankloss} implies the following for $K \geq 2$ and $m_i=\frac{n}{2}-\epsilon_n, \forall i\in[K]$.
\begin{align}\label{eq:halfcondone}
\forall Y_i\subseteq\left[\frac{n}{2}\right], i \in [K] \emph{ s.t. }\sum_{i=1}^K |Y_i|=\min\left(n, K\left(\frac{n}{2}-\epsilon_n\right)\right),
\exists J \subseteq [n]: \sum_{i=1}^K \dim(\mathcal{S}_{J} \cap \mathcal{B}_{i,*,Y_i}) \geq |J|+\frac{n}{2} - 3\epsilon_n.
\end{align}
Now, we can state the following lemma that simplifies the condition~\eqref{eq:halfcondone} resulting from Theorem~\ref{thm:main} for this particular setting.
\begin{lem}\label{lem:halfdofcondition}
For the case where $K \geq 2, m_i=\frac{n}{2}-\epsilon_n, \forall i\in[K]$, the statement in~\eqref{eq:halfcondone} implies the following.
\begin{gather}\label{eq:halfcondtwo}
\exists \tilde{J} \subseteq [n]: \left|\tilde{J}\right|\geq\frac{n}{2}-c K \epsilon_n, \dim\left(\cap_{i=1}^K \mathcal{B}_i \cap \mathcal{S}_{\tilde{J}}\right) \geq \frac{n}{2}-c K \epsilon_n,
\end{gather}
where $c$ is a constant integer.
\end{lem}
\begin{proof}
Assume that~\eqref{eq:halfcondone} holds, and set $Y_1=Y_2=\left[\frac{n}{2}-\epsilon_n\right]$. It then follows that,
\begin{equation}\label{eq:halfproofone}
\exists J\subseteq[n]: \dim(\mathcal{S}_{J} \cap \mathcal{B}_{1})+\dim(\mathcal{S}_{J} \cap \mathcal{B}_{2}) \geq |J| + \frac{n}{2}-5 \epsilon_n.
\end{equation}
Since $\dim(\mathcal{S}_{J} \cap \mathcal{B}_{i})\leq \min\left(|J|,m_i\right), \forall J\subseteq[n], i\in[K]$, it follows from~\eqref{eq:halfproofone} that there exists $J\subseteq[n], |J| \geq \frac{n}{2}-5\epsilon_n, \dim\left(\mathcal{B}_1 \cap \mathcal{B}_2 \cap S_J\right) \geq \frac{n}{2}-9\epsilon_n$; let this set $J$ be called $J_{1,2}$. Similarly, by taking $Y_2=Y_3=\left[\frac{n}{2}-\epsilon_n\right]$, there exists $J \subseteq[n],|J| \geq \frac{n}{2}-5\epsilon_n, \dim\left(\mathcal{B}_2 \cap \mathcal{B}_3 \cap S_J\right) \geq \frac{n}{2}-9\epsilon_n$; call this set $J$ as $J_{2,3}$. Since $\dim\left(\mathcal{B}_2 \cap S_{J_{1,2}}\right) \geq \frac{n}{2}-9\epsilon_n$ and $\dim\left(\mathcal{B}_2 \cap S_{J_{2.3}}\right) \geq \frac{n}{2}-9\epsilon_n$, then $\left|J_{1,2} \cap J_{2,3}\right| \geq \frac{n}{2}-17\epsilon_n$. Also, let $J_{1,2,3}=J_{1,2} \cap J_{2,3}$, then $\dim\left(\mathcal{B}_1 \cap \mathcal{B}_2 \cap \mathcal{B}_3 \cap S_{J_{1,2,3}}\right) \geq \frac{n}{2}-17 \epsilon_n$. Proceeding in the same way, we can show that there exists a constant integer $c$ such that if we let $\tilde{J}=J_{1,2}\cap J_{2,3} \cap \ldots \cap J_{K-1,K}$, then $\left|\tilde{J}\right| \geq c K \epsilon_n$ and $\dim\left(\mathcal{B}_1 \cap \ldots \cap \mathcal{B}_K \cap S_{\tilde{J}}\right) \geq c K \epsilon_n$, and hence,~\eqref{eq:halfcondtwo} follows.
\end{proof}
We use the result of Lemma~\ref{lem:halfdofcondition} together with the following lemma to prove the converse of Theorem~\ref{thm:halfdof}.
\begin{lem}\label{lem:conflict}
If transmitter $i$ is connected to receiver $k$, $k \neq i$ and there exist $J_i, J_k \subseteq [n]$ such that $\dim\left(\mathcal{B}_i \cap S_{J_i}\right) \geq \frac{n}{2}-cK\epsilon_n$ and $\dim\left(\mathcal{B}_k \cap S_{J_k}\right) \geq \frac{n}{2}-c K \epsilon_n$, then the decodability condition of~\eqref{eq:decodability} is satisfied at receiver $k$ only if $|J_i \cap J_k| \leq 2 c K \epsilon_n$.
\end{lem}
\begin{proof}
Since $\dim\left(S_{J_i} \cap \mathcal{B}_i\right) \geq \frac{n}{2}-cK\epsilon_n$, then $\dim(S_{J_i} \cap \mathsf{colspan}(\Lambda_{ik}B_i)) \geq \frac{n}{2}-cK\epsilon_n$ almost surely. Similarly, if $\dim\left(S_{J_k} \cap \mathcal{B}_k\right) \geq \frac{n}{2}-cK\epsilon_n$, then $\dim(S_{J_k} \cap \mathsf{colspan}(\Lambda_{kk}B_k)) \geq \frac{n}{2}-cK\epsilon_n$ almost surely. If $|J_i \cap J_k| > 2cK\epsilon_n$, then $\dim(\mathsf{colspan}(\Lambda_{ik}B_i) \cap \mathsf{colspan}(\Lambda_{kk}B_k)) >0$, and hence, it follows that the dimension of the projection of the received signal at receiver $k$ on the complement of the interference subspace is less than $\dim(\mathcal{B}_k)$, i.e., $\mathsf{dim}\left(\mathsf{Proj}_{{\cal I}_{k}^c} \mathsf{colspan} \left(\Lambda_{kk} B_k\right)\right) < \frac{n}{2}-\epsilon_n$, and therefore, violating~\eqref{eq:decodability}.
\end{proof}
Now, assume that the reduced conflict graph $G$ is not bipartite, then its chromatic number is at least $3$. In this case, we know from Lemma~\ref{lem:halfdofcondition} and Lemma~\ref{lem:conflict} that there exist three users (say users $1$, $2$ and $3$) and three sets $J_i \subseteq [n], i\in\{1,2,3\}$ such that the following holds.
\begin{equation}\label{eq:convcondone}
|J_i| \geq \frac{n}{2} - cK\epsilon_n, \forall i\in\{1,2,3\},
\end{equation}
\begin{equation}\label{eq:convcondtwo}
|J_1 \cap J_2| \leq 2cK\epsilon_n, |J_2 \cap J_3| \leq 2cK\epsilon_n, |J_1 \cap J_3| \leq 2cK\epsilon_n.
\end{equation}
It is easy to see that if $7cK\epsilon_n < \frac{n}{2}$, then one of the conditions in~\eqref{eq:convcondone} and~\eqref{eq:convcondtwo} is violated. It then follows that it cannot be the case that $\epsilon_n \rightarrow 0$, and hence, $\frac{1}{2}$ symmetric DoF cannot be achieved.
Conversely, if $G$ is bipartite, we show a linear coding scheme achieving half symmetric DoF. Consider a coding scheme over two time slots and suppose that each transmitter uses a point-to-point capacity achieving code and whenever it is activated in any of the two time slots, it transmits the selected codeword and otherwise it remains silent. The activation of transmitters is determined by the graph $G$ as follows. Let $P_1$ and $P_2$ be the two partite sets constituting $G$. If vertex $i$ has no outgoing edges, then transmitter $i$ is activated in both time slots. For any remaining user $j$ (corresponding to the nodes that have least one outgoing edge in $G$), if vertex $j$ is in $P_k$, $k\in\{1,2\}$, then transmitter $j$ is active in time slot $k$ and inactive in the other time slot. As an example, for the 6-user topology of Figure \ref{ex_gain}, the retransmission pattern can be written as follows,
\begin{align}\label{code}
\setlength{\arraycolsep}{4pt}
\begin{bmatrix}
B_1 & B_2 & B_3 & B_4 & B_5 & B_6
\end{bmatrix}
=
\begin{bmatrix}
0 & 1 & 0 & 0 & 1 & 1\\
1 & 0 & 1 & 1 & 0 & 1
\end{bmatrix}.
\end{align}
Each column in~\eqref{code} corresponds to a user and each row corresponds to a time slot. For instance, transmitter 2 sends its codeword in time slot 1 and remains silent in time slot 2, whereas transmitter 6 repeats its codeword in both time slots.
We now show that successful decoding is possible for almost all realizations of the channel coefficients, and hence, $1$ DoF is achieved for each user over two time slots. For each receiver $i$ with more than two interfering links, transmitter $i$ is activated in a time slot where all interfering transmitters are silent, and hence, the successful decoding condition of~\eqref{eq:decodability} is guaranteed. In the example of Figure~\ref{ex_gain}, receivers $4$ and $5$ have two interfering links. Transmitters $2$ and $5$ are interfering at receiver $4$, and hence, in~\eqref{code}, transmitters $2$ and $5$ are silent in the second time slot where transmitter $4$ is active. Similarly, transmitters $3$ and $4$ are interfering at receiver $5$, and hence, both are silent in the first time slot where transmitter $5$ is active. For each receiver $i$ with one interfering link, it is either the case that transmitter $i$ is active in a time slot for which the interfering transmitter is silent (as is the case for receiver $6$ in the example), or it is the case that at least one of transmitter $i$ and the interfering transmitter is active in both time slots (as is the case for receivers $1$, $2$ and $3$ in the example); in both cases, the condition in~\eqref{eq:decodability} is satisfied. This completes the proof.
\end{proof}
\subsection{Characterizing the Linear Symmetric DoF for Network Topologies with Exclusive Alignment Sets}\label{sec:nonoverlapping}
In this section, we characterize the linear symmetric degrees of freedom for a broader class of topologies; namely the topologies with ``exclusive alignment sets'', to be defined shortly. This generalizes the result that we proved in the previous section for half linear symmetric DoF.
Consider a coding scheme achieving a linear symmetric DoF $d$ over $n$ time slots. At each receiver $j$, the decodability condition~\eqref{eq:decodability} implies that
$\text{dim} \left({\cal I}_j\right) \leq n\left(1-d\right)$.
Using Theorem~\ref{thm:main}, we obtain the following equivalent condition on the design of beamforming matrices corresponding to interfering signals,
\begin{gather}
\forall Y_i\subseteq[nd], i \in I_j \emph{ s.t. }\sum_{i \in I_j} |Y_i|=\min\left(nd|I_j|,n\right)\nonumber,\\
\exists J \subseteq [n]: \sum_{i \in I_j} \dim(\mathcal{S}_{J} \cap \mathcal{B}_{i,*,Y_i}) \geq |J|+\tau,\label{eq:finalalignment}
\end{gather}
where $\tau=n\left(3d-1\right)$. Using the condition in~\eqref{eq:finalalignment} to reach a converse for the achievable linear symmetric DoF of arbitrary network topologies is a difficult problem, as it is not clear what the required number of time slots $n$ is to achieve $d_{\text{sym}}$. Further, for each value $n$, reducing the complexity of the search for the optimal design of the beamforming matrices by converting it to a problem of combinatorial optimization does not seem straightforward. However, under the restriction to a certain class of topologies, the task becomes easier as it reduces to a problem independent of the value of $n$, and can be described directly in terms of the interference conflict pattern between network users. We introduce the properties of the considered network topologies through the interference sets $I_j, j \in [K]$.
We consider topologies where the following properties hold,
\begin{itemize}
\item {\bf (P1) Maximum Degree:} For all $j \in [K]$, $|I_j| \leq 2$,
\item {\bf (P2) Exclusive Alignment Sets:} For all $j,k \in [K]$ such that $\max\left(|I_j|,|I_k|\right)=2$, $I_j \cap I_k = \phi$.
\end{itemize}
The property (P1) simplifies the problem because for any network with at least one interfering link, $d_{\text{sym}} \leq \frac{n}{2}$. (P1) then implies that $nd|I_j| \leq n$. Therefore,~\eqref{eq:finalalignment} reduces to,
\begin{align}\label{eq:reducedalignment}
\exists J \subseteq [n]: \sum_{i \in I_j} \dim(\mathcal{S}_{J} \cap \mathcal{B}_{i}) \geq |J|+\tau.
\end{align}
Moreover, (P2) simplifies the problem as in this case we can conclude from~\eqref{eq:reducedalignment} that there is no loss in generality in assuming that $\mathcal{S}_J \subseteq \mathcal{B}_i, \forall i \in I_j$ (see Lemma~\ref{lem:non_overlapping} below). The problem then becomes that of finding sparse subspaces for each interference set of size $2$ such that the subspaces corresponding to conflicting interference sets do not overlap. This is captured through the chromatic number of a reduced conflict graph that captures only conflicts between interference sets.
Having the aforementioned properties, we can now state the result on the linear symmetric DoF for the considered class of topologies. We call any network topology with at least one interference link an interference network topology.
\begin{thm}\label{thm:dsym}
For any interference network topology satisfying (P1) and (P2), the linear symmetric DoF is given by,
\begin{equation}\label{eq:ldof}
\emph{LDoF}_{\text{sym}} = \min\left(\frac{1}{2},\frac{\chi(G)+1}{3\chi(G)}\right),
\end{equation}
where $G$ is the reduced conflict graph of the topology (as defined in Definition \ref{redgraph}) and $\chi(.)$ denotes the chromatic number.
\end{thm}
\begin{remk}
In~\cite{topology}, several examples of topologies have been discussed for which the converse for $d_{\text{sym}}$ has remained open. Using Theorem \ref{thm:dsym} , we can now characterize the linear symmetric DoF of all those topologies which satisfy (P1) and (P2). In Figure~\ref{fig:example}, we plot one example where the chromatic number of the reduced conflict graph is 3. The symmetric DoF of $\frac{4}{9}$ is achievable through the structured repetition coding scheme introduced in~\cite{topology}. Here, the converse follows from Theorem~\ref{thm:dsym}.
\end{remk}
\begin{figure}[h]
\centering
\begin{subfigure}{0.3\textwidth}
\centering
\includegraphics[trim = 1.48in 2.3in 6.3in 2.5in, clip,width=.8\textwidth]{4_9_nonoverlapping}
\caption{}
\label{fig:example_topology}
\end{subfigure}
~
\begin{subfigure}{0.3\textwidth}
\centering
\includegraphics[trim = 5.6in 2.2in 1.0in 2.1in, clip,width=\textwidth]{4_9_nonoverlapping}
\caption{}
\label{conf_graph_4_9}
\end{subfigure}
\caption{(a) A topology in which the symmetric DoF $d_{\text{sym}}=\frac{4}{9}$ and (b) the corresponding reduced conflict graph with chromatic number $\chi(G)=3$.}
\label{fig:example}
\end{figure}
\begin{proof}[Proof of Theorem \ref{thm:dsym}]
We know from~\cite{jafar} that for any interference network topology, $d_{\text{sym}} \leq \frac{1}{2}$. For the case where the reduced conflict graph has no edges, i.e., $\chi(G) = 1$, $d_{\text{sym}} = \frac{1}{2}$ is achievable by having each of the column vectors in each matrix $B_i, i\in[K]$ to have no zero entries. Hence, we consider the case where $\chi(G) \geq 2$ and show that $\text{LDoF}_{\text{sym}}=\frac{\chi(G)+1}{3\chi(G)}$ in this case.
We first show that for any receiver with two or more interfering signals, the sparse subspace $\mathcal{S}_J$ of~\eqref{eq:result} is \emph{fully occupied} by the interference, almost surely. More precisely, we prove the following corollary of the equivalent condition of Theorem~\ref{thm:main}.
\begin{lem}\label{lem:minimal}
If~\eqref{eq:result} holds, then for a minimal set $J$ satisfying~\eqref{eq:result}, $\mathcal{S}_J\overset{a.s.}{\subseteq} \mathcal{B}_D$. More precisely,
\begin{gather}
\exists Y_i\subseteq[m_i], i \in [K] \text{ s.t. }\sum_{i=1}^K |Y_i|=\min \left(\sum_{i=1}^K m_i,n\right),\nonumber\\ J\subseteq [n]: \sum_{i=1}^K \dim(\mathcal{S}_{J} \cap \mathcal{B}_{i,*,Y_i}) \geq |J|+x, x \geq \tau, \nonumber\\
\bigwedge \nexists L \subset J: \sum_{i=1}^K \dim(\mathcal{S}_{L} \cap \mathcal{B}_{i,*,Y_i}) \geq |L|+x, \nonumber\\ \Rightarrow \mathcal{S}_J\overset{a.s.}{\subseteq}\mathsf{colspan}\left(\left[\Lambda_1 B_1 ~ \ldots ~ \Lambda_K B_K\right]\right).\label{eq:minimal}
\end{gather}
\begin{proof}
Let $J^*$ be a set satisfying the condition in~\eqref{eq:minimal}. For each $i\in[K]$, let $c_1^{(i)},\ldots,c_{n_i}^{(i)}$ be $n_i$ vectors that form a basis for $\mathcal{S}_{J^*} \cap \mathcal{B}_{i,*,Y_i}$, where $\sum_{i=1}^K n_i = |J^*|+x$ and let $C_i=\left\{c_1^{(i)},\ldots,c_{n_i}^{(i)}\right\}$. Let $C=\{C_1,\ldots,C_K\}$ be the multiset consisting of the elements of $C_i, i\in[K]$ and let $G=(J^* \cup C, E)$ be the bipartite graph whose left partite set consists of vertices corresponding to elements in $J^*$ and right partite consists of vertices corresponding to the elements in $C$, and $\forall i\in J, c \in C, (i,c) \Leftrightarrow c_i \neq 0$.
Since $\nexists L \subseteq J: \sum_{i=1}^K \dim(\mathcal{S}_{L} \cap \mathcal{B}_{i,*,Y_i}) \geq |L|+x+1$, we know that for any subset of vertices $I \subseteq C$, the neighboring set $N(I)$ satisfies the condition $|N(I)| \geq |I|-x$. It follows from Theorem~\ref{thm:hull} that there is a matching in $G$ of size $|C|-x$. Also, since $|J^*|=|C|-x$, we know that there is a matching in $G$ covering all elements of the left partite set. For each $i \in [K]$, let $c_1^{*(i)},\ldots,c_{n^*_i}^{*(i)}$ be the elements of $\{c_1^{(i)},\ldots,c_{n_i}^{(i)}\}$ in the matching in the right partite set $C$, where $\sum_{i=1}^K n^*_i=|J^*|$.
For each $i\in[K]$, let $C_i^*=\left\{c_1^{*(i)},\ldots,c_{n^*_i}^{*(i)}\right\}$. Let $C^*$ be the multiset $C^*=\{C_1^*,\ldots,C_K^*\}$, and consider the bipartite graph $G^*=(J^* \cup C^*, E^*)$, where $\forall i\in J^*, c\in C^*, (i,b) \in E^* \Leftrightarrow c_i \neq 0$. Since $G^*$ has a perfect matching, it follows that $\nexists J \subseteq J^*: \sum_{i=1}^K \dim(\mathcal{S}_{J} \cap \mathcal{B}_{i,*,C_i^*}) > |J|$, and hence, from~\eqref{eq:result} we know that $[\Lambda_1 B_{1,*,C_1^*} ~ \ldots ~ \ldots ~ \Lambda_K B_{K,*,C_K^*}]$ is full rank almost surely. Now, since $\mathsf{colspan}\left(\left[\Lambda_1 B_{1,*,C_1^*} ~ \ldots ~ \Lambda_K B_{K,*,C_K^*}\right]\right)$ has $|J^*|$ linearly independent column vectors almost surely, it follows that $\mathcal{S}_{J^*} \overset{a.s.}{\subseteq} \mathsf{colspan}\left(\left[\Lambda_1 B_{1,*,C_1^*} ~ \ldots ~ \Lambda_K B_{K,*,C_K^*}\right]\right)$, and hence, $\mathcal{S}_{J^*} \overset{a.s.}{\subseteq} \mathcal{B}_D$.
\end{proof}
\end{lem}
From the condition in~\eqref{eq:result}, Lemma~\ref{lem:minimal} and the decodability condition~\eqref{eq:decodability}, we obtain the following condition,
\begin{gather}
\forall r \in [K]: I_r=\{r_1,r_2\}, \exists J_r \subseteq [n]: \nonumber\\\dim(S_{J_r} \cap \mathcal{B}_{r1}) + \dim(S_{J_r} \cap \mathcal{B}_{r2}) \geq |J_r|+\tau, \nonumber\\\dim(S_{J_r} \cap \mathcal{B}_r)=0.\label{eq:alignment}
\end{gather}
We now use the following lemma to restrict our attention to a simpler condition for the considered class of topologies. The proof of the lemma is in Appendix~\ref{app:non_overlapping}.
\begin{lem}\label{lem:non_overlapping}
For the case where alignment sets are exclusive, if there exist $B_1,\ldots,B_K$ such that \eqref{eq:alignment} holds, then there exist $B_1,\ldots,B_K$ such that the following condition holds,
\begin{gather}
\forall r \in [K]: I_r=\{r_1,r_2\}, \exists J_r \subset [n]: \nonumber\\
|J_r|=\tau, \mathcal{S}_{J_r} \subseteq \mathcal{B}_{r1} \cap \mathcal{B}_{r2}, \dim(\mathcal{S}_{J_r} \cap \mathcal{B}_r)=0.\label{eq:newalignment}
\end{gather}
\end{lem}
We now complete the proof by arguing that there exists a design of the beamforming matrices that satisfies~\eqref{eq:newalignment} if and only if $\tau \leq \frac{n}{\chi(G)}$. This follows directly by observing that the decodability condition in~\eqref{eq:newalignment} $\left(\dim(\mathcal{S}_{J_r} \cap \mathcal{B}_r)=0\right)$ is satisfiable if and only if there is no overlap between sparse subspaces corresponding to conflicting interference sets; more precisely,
\begin{gather}
r,d \in [K]: r\in I_d, |I_r|=|I_d|=2 \nonumber\\
\Rightarrow J_r \cap J_d = \phi.\label{eq:newdecode}
\end{gather}
\end{proof}
In order to prove the converse part of Theorem~\ref{thm:dsym}, we showed that the property (P2) allows us to assume that for the optimal coding scheme, whenever the interference alignment condition in~\eqref{eq:reducedalignment} is satisfied for an alignment set, it is the case that $|J|=\tau$ and $S_J \subseteq {\cal B}_i, \forall i\in I_j$. We end this section with two remarks on the generalization of the converse for arbitrary network topologies; in particular, topologies that do not satisfy the property (P2).
\begin{remk}
If the exclusive alignment set property (P2) is not satisfied, then the symmetric linear degrees of freedom can be larger than the value in the statement of Theorem~\ref{thm:dsym}. Consider the $9$-user network depicted in Figure~\ref{fig:ex_top} and note that transmitter $7$ is part of the alignment set at receiver $9$, and is also causing interference at receivers $2$, $4$ and $6$.
While $\frac{\chi(G)+1}{3\chi(G)}=\frac{5}{12}$, we show a scheme that achieves a symmetric DoF of $\frac{3}{7}$. This value of the symmetric DoF can be achieved by using a symbol extension $n=7$ and assigning the sparse subspaces as specified in Figure~\ref{fig:design}.
We explain why almost surely, the decodability condition of~\eqref{eq:decodability} is met at all receivers with one or more interfering transmitter. Since the sparse subspace for the alignment set at receiver $9$ intersects with any of $S_{\{1,2\}}, S_{\{3,4\}}$ and $S_{\{5,6\}}$ in a subspace of dimension at most $1$, the two column vectors in ${\cal B}_7 \cap S_{\{1,3,5\}}$ will almost surely lie outside any of these subspaces, and hence, the decodability conditions at receivers $2$, $4$ and $6$ will be met. Since $S_{\{1,2\}}$, $S_{\{3,4\}}$ and $S_{\{5,6\}}$ do not overlap, the desired signal subspace does not overlap with the interference subspace at each of receivers $1$, $3$ and $5$. Finally, at each of the receivers $1$, $3$, $5$ and $9$, the rank of the interference subspace will be at most $4$ almost surely because~\eqref{eq:reducedalignment} is satisfied with $\tau=2$.
\end{remk}
\begin{figure}[htb]
\centering
\includegraphics[width=0.5\columnwidth]{exampletopology}
\caption{An example of a 9-user network where $\frac{3}{7}$ symmetric DoF are achieved but $\frac{\chi(G)+1}{3\chi(G)}=\frac{5}{12}$.}
\label{fig:ex_top}
\end{figure}
\begin{figure}[htb]
\centering
\includegraphics[width=0.8\columnwidth]{beamformingmat}
\caption{The design of beamforming matrices for the example topology of Figure~\ref{fig:ex_top}. Each figure designates the places of the non-zero entries in the corresponding matrices by crossed squares. Each non-zero entry is drawn independently from a continuous distribution, and successful decoding is guaranteed almost surely.}
\label{fig:design}
\end{figure}
\begin{remk}
We believe that the problem of deriving a converse on the linear symmetric DoF for general network topologies can be simplified by describing each beamforming matrix $B_i$ by the number of vectors it has in each possible sparse subspace. For any set $J \subseteq [n]$, we let $\mu_i(J)$ be the number of vectors that $\mathcal{B}_i$ has in the sparse subspace $S_J$. The numbers $\left\{\mu_i(J), J \subseteq [n]\right\}$ have to satisfy the following constraints for each $i \in [K]$:
$\mu_i([n])=nd$, $\mu_i(J_1\cup J_2)\geq \mu_i(J_1)+\mu_i(J_2)-\mu_i(J_1\cap J_2),~ \forall J_1,J_2\subseteq[n]$, $\mu_i(J_1)\geq \mu_i(J_2)-\left(|J_2|-|J_1|\right). ~ \forall J_1\subseteq J_2\subseteq [n]$.
We can then derive an upper bound on the linear symmetric DoF for each value of $n$ through the solution of an optimization problem that relies on this description for beamforming matrices.
We believe that the upper bound obtained through this method is tight for any network topology. The key question here is to validate whether it is true that for any set of numbers $\left\{\mu_i(J), J\subseteq [n]\right\}$ satisfying the stated conditions, there exists an $n \times d$ matrix $B_i$ such that the number of vectors that $\mathcal{B}_i$ occupies in any sparse subspace $S_J, J \subseteq [n]$ is given by $\mu_i(J)$.
\end{remk}
\section{Conclusion}\label{sec:conclusion}
We characterized necessary and sufficient conditions for almost sure rank loss of a concatenation of full rank matrices with randomly scaled rows. The characterized condition is in terms of the combinatorial structure of the individual matrices with respect to sparse column subspaces. We showed that an almost sure rank loss by a factor $\tau$ is possible if and only if the total number of column vectors in the ensemble in a sparse subspace exceeds the size of the sparse subspace by the same factor $\tau$.
We then used the result to characterize the linear symmetric DoF for a class of topological interference management problems that was previously studied in~\cite{topology}. In particular, we could identify necessary and sufficient conditions on the network topology, under which, a linear symmetric DoF of $\frac{1}{2}$ is achievable. Further, we characterized the linear symmetric DoF for any network topology with a maximum receiver degree of $3$ and exlusive alignment sets. In general, our result solves an underlying fundamental problem in topological interference management; we are considering for future work how it can be used as a cornerstone to characterize the linear degrees of freedom region for arbitrary network topologies.
\appendices
\section{Proof of Claim~\ref{matroid}}\label{sec:matroid_proof}
Without loss of generality, we consider the case of $i=1$. We first show that $M_{1,X,Y_1}$ is a matroid. To this end, we need to prove the following properties.
\begin{itemize}
\item If $I\subseteq J$ and $J\in \mathcal{I}_{1,X,Y_1}$, then $I\in \mathcal{I}_{1,X,Y_1}$: This is clear since $\mathcal{S}_{I\cup X^c}\subseteq \mathcal{S}_{J\cup X^c}$ and we therefore have
\begin{gather*}
\dim (\mathcal{S}_{I\cup X^c}\cap \mathcal{B}_{1,*,Y_1}) \leq \dim (\mathcal{S}_{J\cup X^c}\cap \mathcal{B}_{1,*,Y_1}) =0 \nonumber\\
\Rightarrow \dim (\mathcal{S}_{I\cup X^c}\cap \mathcal{B}_{1,*,Y_1}) =0.
\end{gather*}
\item If $I\in\mathcal{I}_{1,X,Y_1}$, $J\in\mathcal{I}_{1,X,Y_1}$ and $|I|<|J|$, then $\exists j\in J \setminus I$ s.t. $I\cup \{j\}\in\mathcal{I}_{1,X,Y_1}$:
Assume $I=\{i_1,...,i_k\}$ and $J=\{i_1,...,i_{k'},j_1,...,j_l\}$ where we have $k'\leq k$ and $k'+l>k$. Using this notation, we have that $J\setminus I=\{j_1,...,j_l\}$. Suppose that this property is not true. This implies that adding any of the $e_{j_n}$'s to $I$ will force it to lie outside the set $\mathcal{I}_{1,X,Y_1}$ (where $e_{j_n}$ denotes the $j_n^{th}$ standard basis vector). If $ X^c=\{x_1,...,x_m\}$, this is equivalent to the fact that there exist coefficients $\lambda_{ij}$ ($i\in [l],j\in[k]$), $\beta_{ij}$ ($i\in[l],j\in[m]$) and $\mu_i\neq 0$ ($i\in[l]$) such that
\begin{align}
\lambda_{11}e_{i_1}+...+\lambda_{1k}e_{i_k}+\beta_{11}e_{x_1}+...+\beta_{1m} e_{x_m}+ mu_{1}e_{j_1}&=v_1\label{coeffs1}\\
\lambda_{21}e_{i_1}+...+\lambda_{2k}e_{i_k}+\beta_{21}e_{x_1}+...+\beta_{2m} e_{x_m}+mu_{2}e_{j_2}&=v_2\\
\vdots~~~~~~~~~~~~~~~&\nonumber\\
\lambda_{l1}e_{i_1}+...+\lambda_{lk}e_{i_k}+\beta_{l1}e_{x_1}+...+\beta_{lm} e_{x_m}+\mu_{l}e_{j_l}&=v_l\label{coeffs3},
\end{align}
where $v_1,...,v_l$ belong to $\mathcal{B}_{1,*,Y_1}$. Now, consider the following vectors
\begin{align}
\overrightarrow{\lambda_i}=[\lambda_{i,k'+1}~ \lambda_{i,k'+2}~ ...~ \lambda_{ik}]^T,~i\in[l].
\end{align}
There are $l$ of these vectors in $\mathbb{R}^{k-k'}$ and since we assumed that $k'+l>k$, these vectors should be linearly dependent. This implies that there exist not-all-zero coefficients $a_i$, $i\in\{1,...,l\}$ such that $\sum_{i=1}^l a_i \overrightarrow{\lambda_i}=\overrightarrow{0}$. Multiplying each $a_i$ by the corresponding equation in (\ref{coeffs1})-(\ref{coeffs3}) and then adding the resulting equations yields
\begin{align}\label{lincomb}
\sum_{n=1}^{k'}\left(\sum_{i=1}^l a_i \lambda_{in}\right) e_{i_n} + \sum_{n=1}^{m}\left(\sum_{i=1}^l a_i \beta_{in}\right) e_{x_n} + \sum_{i=1}^l (a_i \mu_i) e_{j_i}=\sum_{i=1}^l a_i v_i,
\end{align}
where the vectors in (\ref{lincomb}) are non-zero since the coefficients $a_i$ are not all zeros and all the coefficients $\mu_i$ are non-zero. The RHS of (\ref{lincomb}) is a vector in $\mathcal{B}_{1,*,Y_1}$, which implies that there exists a linear combination of the bases of $S_{J\cup X^c}$ which lies in $\mathcal{B}_{1,*,Y_1}$, therefore contradicting the fact that $J\in\mathcal{I}_{1,X,Y_1}$. Hence, this property is also true.
\end{itemize}
Having proven the above properties, it is now verified that
$M_{1,X,Y_1}=(X,\mathcal{I}_{1,X,Y_1})$ is a matroid.
To complete the proof of Claim \ref{matroid}, we need to show that the rank function of $M_{1,X,Y_1}$ is equal to $r_{1,X,Y_1}(J)=|J|-\dim (\mathcal{S}_{J\cup X^c}\cap \mathcal{B}_{1,*,Y_1}) , J\subseteq X$. We do so through the following two steps.
\begin{itemize}
\item For any $I\subseteq J$ such that $\dim (\mathcal{S}_{I\cup X^c} \cap \mathcal{B}_{1,*,Y_1}) =0$, we have that $(\mathcal{S}_{J\cup X^c} \cap \mathcal{S}_{I\cup X^c} ) + (\mathcal{S}_{J\cup X^c} \cap \mathcal{B}_{1*,Y_1}) \subseteq \mathcal{S}_{J\cup X^c}$.\footnote{For two subspaces $\mathcal{U}$ and $\mathcal{V}$, $\mathcal{U}+\mathcal{V}$ stands for $\text{span}(\mathcal{U}\cup\mathcal{V})$.} Therefore, since $\dim (\mathcal{S}_{I\cup X^c} \cap \mathcal{B}_{1,*,Y_1}) =0$, we will get $\dim (\mathcal{S}_{J\cup X^c} \cap \mathcal{S}_{I\cup X^c} ) + \dim (\mathcal{S}_{J\cup X^c} \cap \mathcal{B}_{1,*,Y_1}) \leq \dim(\mathcal{S}_{J\cup X^c} ) $. But $\mathcal{S}_{J\cup X^c} \cap \mathcal{S}_{I\cup X^c} =\mathcal{S}_{I\cup X^c} $. Hence, $\dim (\mathcal{S}_{I\cup X^c} ) \leq \dim(\mathcal{S}_{J\cup X^c} ) - \dim (\mathcal{S}_{J\cup X^c} \cap \mathcal{B}_{1,*,Y_1}) $, implying that $r_{1,X}(J)\leq |J|-\dim (\mathcal{S}_{J\cup X^c} \cap \mathcal{B}_{1,*,Y_1}) $.
\item
Now, we only need to show that there exists some $K\subseteq J $ for which $|K| = |J |- \dim (\mathcal{S}_{J\cup X^c} \cap \mathcal{B}_{1,*,Y_1}) $ and $\dim (\mathcal{S}_{K\cup X^c} \cap \mathcal{B}_{1,*,Y_1}) =0$. By definition, vectors $\{e_i: i\in J\cup X^c\}$ form a basis for the vector space $\mathcal{S}_{J\cup X^c} $. Suppose that the vectors in $\{f_1,...,f_d\}$ form a basis for the space $\mathcal{S}_{J\cup X^c} \cap \mathcal{B}_{1,*,Y_1}$. Therefore, since $\dim(\mathcal{S}_{X^c}\cap \mathcal{B}_{1,*,Y_1})=0$, the vectors in $\{f_1,...,f_d\} \cup\{e_i:i\in X^c\}$ form a basis for
$\mathcal{S}_{J\cup X^c} \cap \mathcal{B}_{1,*,Y_1} + \mathcal{S}_{X^c}$. Then, by the Steinitz exchange lemma, there exists a subset $K\subseteq J$ such that the vectors in $\{f_1,...,f_d\}\cup\{e_i:i\in K\cup X^c\}$ form a basis for the space $\mathcal{S}_{J\cup X^c}$. This implies that $\mathcal{S}_{K}$ has no intersection with $\mathcal{S}_{J\cup X^c} \cap \mathcal{B}_{1,*,Y_1} + \mathcal{S}_{X^c}$ and therefore with $\mathcal{B}_{1,*,Y_1}$ (except for the zero vector) and $|K| = |J\cup X^c|- \dim (\mathcal{S}_{J\cup X^c} \cap \mathcal{B}_{1,*,Y_1} + \mathcal{S}_{X^c})=|J|-\dim (\mathcal{S}_{J\cup X^c}\cap \mathcal{B}_{1,*,Y_1}) $, suggesting that $K$ is the actual desired subset of $J$.
\end{itemize}
The above two steps imply that $r_{1,X,Y_1}(J)=|J|-\dim (\mathcal{S}_{J\cup X^c}\cap \mathcal{B}_{1,*,Y_1}) , J\subseteq X$. The same arguments hold for any $M_{i,X,Y_i}=(X,\mathcal{I}_{i,X,Y_i}), i\in[K]$ as well. This completes the proof.
\section{Proof of Lemma~\ref{lem:non_overlapping}}\label{app:non_overlapping}
Fix a design for the transmit beamforming matrices $B_i: i \in [K]$ such that~\eqref{eq:alignment} holds. For each $r \in [K]: |I_r| = 2$, fix a subset $J_r \subseteq [n]$ such that the condition in~\eqref{eq:alignment} is satisfied for the chosen subsets; it is easy to verify that $|J_r| \geq \tau$ for any selected subset. We then set $J_r = \phi$ for every $r \in [K]$ such that $|I_r| < 2$.
For each $r \in [K]$ such that $I_r=\{r_1,r_2\}$, we choose the new beamforming matrices $B_{r1}^{(\text{new})}$ and $B_{r2}^{(\text{new})}$ as follows. let $J'_r=J_r \backslash \left(J_{r1} \cup J_{r2}\right)$ and $\mathcal{S}'_r=\left(\mathcal{S}_{J_r} \backslash \mathcal{S}_{J_{r1}} \cup \mathcal{S}_{J_{r2}}\right)$; since \eqref{eq:alignment} is satisfied, we know that the following holds,
\begin{equation}
\dim(\mathcal{S}'_r \cap \mathcal{B}_{r1})+\dim(\mathcal{S}'_r \cap \mathcal{B}_{r2}) \geq |J'_r|+\tau,
\end{equation}
and hence, we also know that $|J'_r| \geq \tau$. For $i \in \{1,2\}$, let $B'_{ri}$ be an $n \times nd$ matrix with $nd$ columns forming a basis for $B_{ri}$ and a subset of these columns form a basis for $\mathcal{S}'_r \cap \mathcal{B}_{ri}$; we then fix $J_r^{(\text{new})} \subseteq J'_r: |J_r^{(\text{new})}|=\tau$ and replace the columns that form a basis for $\mathcal{S}'_r \cap \mathcal{B}_{ri}$ with an equal number of linearly independent columns in $\mathcal{S}'_r$ and $\tau$ of the new columns form a basis for $\mathcal{S}_{J_r^{(\text{new})}}$ to construct the matrix $B_{ri}^{(\text{new})}$ from $B'_{ri}$.
After performing the above step, it is straightforward to verify that the following holds,
\begin{gather}
\forall r \in [K]: I_r=\{r_1,r_2\}, \nonumber\\\mathcal{S}_{J_r^{(\text{new})}} \subseteq \mathcal{B}_{r1}^{(\text{new})} \cap \mathcal{B}_{r2}^{(\text{new})}, \dim\left(\mathcal{S}_{J_r^{(\text{new})}} \cap \mathcal{B}_r^{(\text{new})}\right)=0.\label{eq:alignmentsatisfied}
\end{gather}
and hence, the new beamforming matrices satisfy the condition in~\eqref{eq:newalignment}.
\bibliographystyle{IEEEtran}
{\footnotesize
|
\section{INTRODUCTION}
Galaxy interactions and mergers play important roles in triggering star formation and/or fueling the nuclear activity in the merging host galaxies \citep{hop06}. Recent high resolution simulations of major mergers show that large scale tidal forces as well as small scale turbulence and stellar feedback can significantly influence the distribution of gas, forming massive clumps of dense gas with $M_{\rm{H_2}}$ = 10$^6$ -- 10$^8$~M$_{\odot}$ \citep[e.g.,][]{tey10, hop13}. These simulations also predict that the star formation not only increases as the galaxies first collide, but it also persists at a higher rate throughout the merger process, peaking at the final coalescence.
(Ultra-)Luminous Infrared Galaxies \citep[U/LIRGs;][]{soi87} at low redshifts are almost exclusively strongly interacting and merging systems \citep{kar10}, often found at the mid to final stages of the merger. The elevated level of infrared luminosity originates from the reprocessed emission from the dust particles surrounding the starburst or the Active Galactic Nuclei (AGNs), both of which are likely triggered by the tidal interaction. The highest gas surface densities ($\Sigma_{\rm{H_2}}$ = 5.4 $\times$ 10$^4$ -- 1.4 $\times$ 10$^5$~M$_{\odot}$~pc$^{-2}$) and consequently the highest star formation activities ($\Sigma_{\rm{SFR}}$ = $\sim$ 1000~M$_{\odot}$~yr$^{-1}$~kpc$^{-2}$) are usually found near the compact nuclear region \citep[e.g., Arp220, NGC 6240;][]{d&s98, eng10, wil14}. Dense molecular gas ($n$ $\sim$ 10$^5$ -- 10$^7$~cm$^{-3}$) in U/LIRGs directly shows nuclear gas distribution and kinematics \citep[e.g.,][]{ion04, sak14}. They are often surrounded by diffuse gas ($n$ $\sim$ 10$^2$ -- 10$^3$~cm$^{-3}$) that may or may not be directly associated with star formation activities.
It has been demonstrated that the HCN~(4--3) and HCO$^+$~(4--3) emission lines, whose critical densities are 8.5 $\times$ 10$^6$ and 1.8 $\times$ 10$^6$ cm$^{-3}$, respectively, can be used as tracers of the dense gas \citep[e.g.,][]{ion13, gar14, ima14}. On the other hand, CO~(1--0) and $^{13}$CO~(1--0) line emission, whose critical densities are 4.1 $\times$ 10$^2$ and 1.5 $\times$ 10$^3$ cm$^{-3}$, respectively, have been used extensively for tracing the global gas distribution and kinematics in merging U/LIRGs \citep[e.g.,][]{yun94, ion04, ued14}. In addition, the ratio of these lines (e.g. $^{12}$CO/$^{13}$CO and HCN/HCO$^+$) have been used to investigate the properties of the ISM \citep{cas92, aal97} or to search for buried AGNs \citep[e.g.,][]{ima07, ima14}. Limitations in sensitivity and angular resolution have been the major obstacles in understanding the detailed distribution and kinematics of both dense and diffuse gas, and investigating the spatial variation of the line ratios and the physical condition of gas.
\begin{figure}
\begin{center}
\includegraphics[scale=.35]{HST_ALMA_VV114.eps}
\caption{The HST/ACS image of VV~114 \citep{eva08}. There is a dust lane from north to south in front of the eastern galaxy. The red crosses show the positions of the nuclei defined by the peak positions of the Ks-band observation \citep{tat12}. The white ellipse shows a field of view of 3-point mosaic observation with band 7, while the white circle shows a field of view of 7-point mosaic observation with band 7 (see \S\ref{obs}).
}
\label{fig_HST}
\end{center}
\end{figure}
In this paper, we present \textit{Atacama Large Millimeter/submillimeter Array} (ALMA) cycle 0 observations of the IR-bright merging galaxy VV~114. VV~114 is one of the best samples for studying the gas response during the critical stage when the two gas disks merge \citep{ion05, wil08}. The target molecular lines include $^{12}$CO~(1--0), $^{13}$CO~(1--0), $^{12}$CO~(3--2), HCN~(4--3) and HCO$^+$~(4--3), and we also present the maps of CH$_3$OH~(2$_k$--1$_k$), CS~(2--1), CN~(1$_{1/2}$--0$_{1/2}$), CN~(1$_{3/2}$--0$_{1/2}$), and CS~(7--6) lines which were observed simultaneously within the same band. The main aim of this study is to investigate the distribution and kinematics of the diffuse and dense molecular gas and to quantify the spatial variation of the excitation conditions across the two merging disks.
VV~114 is a gas-rich \citep[$M_{\rm{H_{2}}}$ = 5.1 $\times$ 10$^{10}$~M$_{\odot}$;][]{yun94} nearby (D = 82~Mpc; 1\farcs0 = 400~pc) interacting system (Figure~\ref{fig_HST}) with high-infrared luminosity \citep[$L_{\rm{IR}}$ = 4.7 $\times$ 10$^{11}$~$L_{\odot}$;][]{arm09}. The projected nuclear separation between the two optical galaxies (VV 114E and VV 114W) is about 6~kpc. \citet{fry99} found a large amount of dust ($M_{\rm{dust}}$ = 1.2 $\times$ 10$^8$~M$_{\odot}$) distributed across the system with a dust temperature of 20 -- 25~K. About half of the warmer dust traced in the mid-IR (MIR) is associated with the eastern galaxy, where both compact (nuclear region) and extended emission is found \citep{lfl02}. \citet{ric11} found a bimodal distribution of velocity dispersions of several atomic forbidden lines and emission line ratios indicative of composite activity explained by a combination of wide-spread shocks and star formation. The wide-spread star formation is also revealed by Pa$\alpha$ observation using ANIR camera mounted on miniTAO \citep[see also Appendix~\ref{A1}]{tat12}. \citet{ion13} (hereafter \citetalias{ion13}) identified a highly obscured AGN and compact starburst clumps using sub-arcsecond resolution ALMA cycle 0 observations of HCN~(4--3) and HCO$^+$~(4--3) emission.
This paper is organized as follows. We describe our observations and data reduction in \S\ref{obs}, and results in \S\ref{result}. In \S\ref{ratio} and \S\ref{der}, we provide molecular line ratios and physical parameters, such as the gas/dust mass, the gas temperature, and the gas density. In \S\ref{dis}, we present the properties of ``dense" gas (\S\ref{dense}), the comparison between molecular gas and star formation (\S\ref{SFR}), the discussions of the CO isotope enhancement (\S\ref{isotope}), the gas-to-dust mass ratio (\S\ref{G/D}), the fractional abundances of CS, CH$_3$OH, and CN relative to H$_2$ (\S\ref{chemical}), and a potential tidal dwarf galaxy formations at the tip of the tidal arms of VV~114 (\S\ref{TDG}). We summarize and conclude this paper in \S\ref{conclusion}. Throughout this paper, we adopt H$_0$ = 73 km s$^{-1}$ Mpc$^{-1}$, $\Omega_{\rm{M}}$ = 0.27, and $\Omega_{\rm{\Delta}}$ = 0.73.
\section{OBSERVATIONS AND DATA REDUCTION} \label{obs}
Observations toward VV~114 were carried out as an ALMA cycle 0 program (ID = 2011.0.00467.S; PI = D. Iono) using fourteen -- twenty 12~m antennas. The band~3 and band~7 receivers were tuned to the $^{12}$CO~(1--0), $^{13}$~CO~(1--0), $^{12}$CO~(3--2), HCN~(4--3), and HCO$^+$~(4--3) line emissions in the upper side band (see Table~\ref{table_obs}). The $^{12}$CO~(1--0) data were obtained on November 6, 2011 and May 4, 2012 in the compact and extended configurations, respectively. The $^{13}$CO~(1--0) data were obtained on May 27 and July 2, 2012 in the compact configuration. The $^{12}$CO~(3--2) emission was observed on November 5, 2011 in the compact configuration (7-point mosaic). The HCN~(4--3) and HCO$^+$~(4--3) data were obtained on July 1, 2, and 3, 2012 in the extended configuration (3-point mosaic), simultaneously. Each spectral window had a bandwidth of 1.875~GHz with 3840~channels, and two spectral windows were set to each sideband to achieve a total frequency coverage of $\sim$ 7.5~GHz in these observations. The spectral resolution was 0.488~MHz per channel. J1924-292, J0132-169, Uranus (Neptune for band 3 observations) were used for bandpass, phase, and flux calibrations. Detailed observational parameters are shown in Table~\ref{table_obs}.
\begin{figure*}[tbh]
\begin{center}
\includegraphics[scale=.5]{fig_mom1.eps}
\caption{(a) $^{12}$CO~(1--0) integrated intensity image overlaid on the HST/ACS/F435W image of VV~114. The contours are 0.2, 0.4, 0.8, 1.6, 3.2, 6.4, 12.8, 25.6, and 33.0~Jy~km~s$^{-1}$. The dashed green box shows an imaging field of other lines and continuum of this work except for the CO and $^{13}$CO lines. (b) $^{12}$CO~(1--0) velocity field image. The velocity field image in color scale ranges from 5600~km~s$^{-1}$ to 6200~km~s$^{-1}$. The dashed black lines represent tidal arms of VV~114. The dashed red line tracks the filamentary structure detected in images of other lines and dust continuum, and the dashed circle shows the overlap region. (c) $^{12}$CO~(1--0) velocity dispersion image. The velocity dispersion image in color scale ranges from 0~km~s$^{-1}$ to 120~km~s$^{-1}$. (d) The same as (a) but for $^{13}$CO~(1--0). The contours are 0.02, 0.04, 0.08, 0.16, 0.32, and 0.64~Jy~km~s$^{-1}$. (e/f) The same as (b/c), respectively, but for $^{13}$CO~(1--0). (g) The same as (a) but for CS~(2--1). The contours are 0.04, 0.08, 0.16, and 0.28~Jy~km~s$^{-1}$. (h/i) The same as (b/c), respectively, but for CS~(2--1). The beam size of each line is shown in the bottom-left of the images (Table~\ref{table_data}). The red crosses show the positions of the nuclei defined by the peak positions of the Ks-band observation \citep{tat12}.
}
\label{fig_mom1}
\end{center}
\end{figure*}
We used the delivered calibrated data and mapping was accomplished using the {\tt clean} task in {\tt CASA} \citep{CASA}. We made the data cubes with a velocity width of 5 km s$^{-1}$ for the $^{12}$CO line and 30 km s$^{-1}$ for the other lines. All maps in this paper, except for $^{12}$CO~(3--2), are reconstructed with a Briggs weighting \citep[robust = 0.5;][]{b&c92} and analyzed with {\tt MIRIAD} and {\tt AIPS}. The $^{12}$CO~(3--2) images are created with uniform weighting (see \S\ref{co32}). The synthesized beam size of the $^{12}$CO~(1--0), $^{13}$CO~(1--0), $^{12}$CO~(3--2), and HCN~(4--3) were 1\farcs97 $\times$ 1\farcs35 (P.A. = 82.3~deg.), 1\farcs77 $\times$ 1\farcs20 (P.A. = 85.8~deg.), 1\farcs64 $\times$ 1\farcs17 (P.A. = 112.6~deg.), and 0\farcs46 $\times$ 0\farcs38 (P.A. = 51.5~deg.), respectively. We also detected CN~(1$_{3/2}$--0$_{1/2}$), CN~(1$_{1/2}$--0$_{1/2}$), CS~(2--1), CH$_3$OH~(2$_k$-1$_k$), and CS~(7--6) line emission for the first time in VV~114. The properties of these molecular lines are summarized in Table~\ref{table_data}. All images which we constructed are corrected for primary beam attenuation. The on-source times of band~3 and band~7 were about 40 minutes and 80 minutes, and the rms noise levels of the channel maps with 30~km~s$^{-1}$ resolution are 1.0~mJy~beam$^{-1}$ and 0.8~mJy~beam$^{-1}$, respectively. Furthermore, we made continuum maps at each observing frequency by adding the line-free channels. The rms level of the continuum images were 0.05~mJy~beam$^{-1}$, 0.11~mJy~beam$^{-1}$, and 0.07~mJy~beam$^{-1}$ for band~3, band~7 in the compact configuration, and band~7 in the extended configuration, respectively. The continuum emission was subtracted in the $uv$-plane before making the line images. Throughout this paper, the pixel scales of the band~3 and the band~7 images are set to 0\farcs3/pixel and 0\farcs08/pixel, respectively, and only the statistical error is considered unless mentioned otherwise. The systematic error on the absolute flux is estimated to be $\sim$ 5\% and $\sim$ 10\% for both sidebands in band 3 and band 7, respectively.
In the following sections, we estimate the missing flux of each molecular line for which the single dish data are available in literature. Although the effect of missing flux becomes critical when we evaluate the global gas properties and the corresponding line ratios, the effect is negligible when we discuss structures that are smaller than the ``maximum recoverable scale" (MRS) of each configuration of ALMA. This is estimated from the minimum baseline lengths of the assigned antenna configurations and the observed frequencies. The MRS of our observations are $\sim$ 8\arcsec~and $\sim$ 7\arcsec~in band 3 and band 7, respectively (Table \ref{table_obs}). Therefore the missing flux effect in this paper is negligible, since we derive physical parameters (e.g., molecular gas mass) only for structures smaller than $\sim$ 2\arcsec.
\begin{figure*}
\begin{center}
\includegraphics[scale=.5]{fig_mom2.eps}
\caption{The same as Figure \ref{fig_mom1} but for (a, b, and c) CH$_3$OH~(2$_k$--1$_k$), (d, e, and f) CN~(1$_{1/2}$--0$_{1/2}$), and (g, h, and i) CN~(1$_{3/2}$--0$_{1/2}$). (a) The contours are 0.02, 0.04, 0.08, 0.16, and 0.32~Jy~km~s$^{-1}$ (d) The contours are 0.04, 0.08, 0.16, 0.32, and 0.50~Jy~km~s$^{-1}$. (g) The contours are 0.04, 0.08, 0.16, 0.32, 0.64, 1.00, and~1.20 Jy~km~s$^{-1}$.
}
\label{fig_mom2}
\end{center}
\end{figure*}
\section{RESULTS} \label{result}
Molecular line and continuum images are shown in Figures~\ref{fig_mom1}, \ref{fig_mom2}, \ref{fig_mom3}, \ref{fig_VV114E}, and \ref{fig_contin}. The channel maps and the spectra of all line emissions are shown in Appendix~\ref{A2} and \ref{A3}.
\subsection{Line Emissions in Band 3}
\subsubsection{$^{12}$CO~(1--0)}
The integrated intensity, velocity field, and velocity dispersion maps of VV~114 are shown in Figures~\ref{fig_mom1}a, \ref{fig_mom1}b, and \ref{fig_mom1}c, respectively. The total $^{12}$CO~(1--0) integrated intensity of VV~114 is 594.6 $\pm$ 1.6~Jy~km~s$^{-1}$, which is 1.3 times larger than that detected using the NRAO 12~m telescope \citep[461~Jy~km~s$^{-1}$;][]{san91}. This is because the pointing center for the NRAO 12~m observation was 25\farcs0 southwest of the CO centroid identified from the ALMA map (NRAO 12~m: 01h07m45.7s, -17d30m36.5s; CO centroid: 01h07m47.2s, -17d30m25.8s). At the adopted distance of VV~114 (86~Mpc), the 1\farcs97 $\times$ 1\farcs35 beam of the $^{12}$CO~(1--0) observation gives us a resolution of 790~pc $\times$ 540~pc. The two crosses shown in all images represent the peaks obtained from the miniTAO/ANIR $K$s-band observation, and we regard them as the progenitor's nuclei.
The integrated $^{12}$CO~(1--0) intensity map of VV~114 (Figure~\ref{fig_mom1}a) shows that the diffuse/cold gas forms two arm-like structures and a filamentary structure located at the center of the image. The global gas distribution is consistent with the previous $^{12}$CO~(1--0) observations \citep{yun94}. The southeastern (SE) arm clearly follows the tidal arm seen in the HST/ACS image \citep[Figure~\ref{fig_HST};][]{eva08}, while the northwestern (NW) arm has no counterpart in any other wavelengths. The region from the center of VV~114 to the eastern nucleus shows a strong concentration of molecular gas ($\simeq$ 5\farcs5 west of the eastern nucleus), and we refer to this region as the ``overlap" region with a molecular ``filament" (see Figure~\ref{fig_mom1}).
\begin{figure*}
\begin{center}
\includegraphics[scale=.5]{fig_mom3.eps}
\caption{The same as Figure \ref{fig_mom1} but for (a, b, and c) $^{12}$CO~(3--2), (d, e, and f) HCN~(4--3), and (g, h, and i) HCO$^+$~(4--3). (a) The contours are 2, 4, 8, 16, 32, 64, 128, and 170~Jy~km~s$^{-1}$ (d) The contours are 0.04, 0.08, 0.16, 0.32, 0.64, 1.28, and 1.80~Jy~km~s$^{-1}$. (g) The contours are 0.04, 0.08, 0.16, 0.32, 0.64, 1.28, and 2.40~Jy~km~s$^{-1}$.
}
\label{fig_mom3}
\end{center}
\end{figure*}
The $^{12}$CO~(1--0) velocity field map of VV~114 (Figure~\ref{fig_mom1}b) shows a significantly broad velocity range across the galaxy disks ($\simeq$ 600~km~s$^{-1}$). The SE arm has a blue-shifted velocity from 5650~km~s$^{-1}$ to 5920~km~s$^{-1}$, while the NW arm has a red-shifted velocity from 5950~km~s$^{-1}$ to 6160~km~s$^{-1}$. One possibility for the larger velocity width in the SE arm may be a highly inclined tidal arm. Two other arm-like features are also detected in the $^{12}$CO~(1--0) observations. One arm is located $\simeq$ 4\farcs0 northeast of the eastern nucleus and shows an arc around the eastern nucleus in the velocity range of 5810~km~s$^{-1}$ to 6180~km~s$^{-1}$. The other arm is located $\simeq$ 10\farcs0 west of the SE arm and has a strong peak ($\simeq$ 262.5 $\pm$ 1.0~Jy~km~s$^{-1}$) in the velocity range of 5610~km~s$^{-1}$ to 5900~km~s$^{-1}$.
The overlap region has the highest velocity dispersion ($\simeq$ 110~km~s$^{-1}$) (Figure~\ref{fig_mom1}c). The NW arm has an average velocity dispersion of $\simeq$ 30~km~s$^{-1}$, while the SE arm has $\simeq$ 40~km~s$^{-1}$. These values are significantly higher than the dispersions seen in Giant Molecular Clouds (GMCs) in the LMC \citep[2 -- 14~km~s$^{-1}$;][]{mnm08, fji14} and slightly higher than that in Giant Molecular Associations (GMAs) in the Antennae galaxy \citep[6 -- 36~km~s$^{-1}$;][]{ued12}. We suggest that the main contribution to the $^{12}$CO~(1--0) velocity dispersion is inter cloud turbulent medium along the tidal arm, and/or shocked region induced by the tidal interaction, rather than the velocity dispersion of the GMCs/GMAs.
\subsubsection{$^{13}$CO~(1--0)}
The integrated intensity, velocity field, and velocity dispersion maps of $^{13}$CO~(1--0) are shown in Figures~\ref{fig_mom1}d, \ref{fig_mom1}e, and \ref{fig_mom1}f, respectively.
The integrated $^{13}$CO~(1--0) intensity map of VV~114 (Figure~\ref{fig_mom1}d) shows a filamentary structure across the galaxy disks, which is consistent with the region where the $^{12}$CO~(1--0) filament is detected. The total $^{13}$CO~(1--0) integrated intensity is 5.9 $\pm 0.4$~Jy~km~s$^{-1}$. The strongest peak is located $\simeq$ 4\farcs2 southwest of the eastern nucleus.
The $^{13}$CO~(1--0) velocity field map of VV~114 (Figure~\ref{fig_mom1}e) shows a narrower velocity range (5670 -- 6000~km~s$^{-1}$) than that of the $^{12}$CO~(1--0) emission (5600 -- 6200~km~s$^{-1}$). This suggests that the $^{13}$CO~(1--0) emission mainly comes from two components, the eastern galaxy and the blue-shifted component of the overlap region.
The $^{13}$CO~(1--0) velocity dispersion map of VV~114 (Figure~\ref{fig_mom1}f) shows a high velocity dispersion component ($\sim$ 100~km~s$^{-1}$) between the eastern nucleus and the overlap region. This significant velocity dispersion may be caused by a superposition of clouds (see the double-peak spectrum at R39 shown in Appendix~\ref{A3}).
\subsubsection{CS~(2--1) and CH$_3$OH~(2$_k$--1$_k$)}
The CS~(2--1) and CH$_3$OH~(2$_k$--1$_k$) lines are only detected at the overlap region (Figures~\ref{fig_mom1}g, \ref{fig_mom1}h, \ref{fig_mom1}i, \ref{fig_mom2}a, \ref{fig_mom2}b, and \ref{fig_mom2}c). This is the first detection of the CH$_3$OH~(2$_k$--1$_k$) emission in a merger-induced overlap region. We observed the blended set of 2$_1$ -- 1$_1$ ($\nu_{\rm{rest}}$ = 96.756~GHz, $E_{\rm{up}}/k$ = 28.0~K), 2$_0$ -- 1$_0$ \textit{E} ($\nu_{\rm{rest}}$ = 96.745~GHz, $E_{\rm{up}}/k$ = 20.1~K), 2$_0$ -- 1$_0$ \textit{A}$^+$ ($\nu_{\rm{rest}}$ = 96.741~GHz, $E_{\rm{up}}/k$ = 7.0~K), and 2$_{-1}$ -- 1$_{-1}$ \textit{E} ($\nu_{\rm{rest}}$ = 96.739~GHz, $E_{\rm{up}}/k$ = 12.5~K), thermal transitions of CH$_3$OH (hereafter designated the 2$_k$ -- 1$_k$ transition). The distribution of these molecular lines is clearly different from the other dense gas tracers detected in the current program. The peaks of CS~(2--1) and CH$_3$OH~(2$_k$--1$_k$) are coincident with one of the peaks of $^{13}$CO~(1--0) to within 0\farcs5. The total CS~(2--1) and CH$_3$OH~(2$_k$--1$_k$) integrated intensities are 0.4 $\pm$ 0.1~Jy~km~s$^{-1}$ and 0.5 $\pm$ 0.1~Jy~km~s$^{-1}$, respectively. The signal to noise is too low to resolve the velocity structure.
\begin{figure}
\begin{center}
\includegraphics[scale=.2]{VV114E.eps}
\caption{(a) 340~GHz continuum flux image of VV~114E. The flux in color scale ranges from 0~mJy~beam$^{-1}$ to 3.4~mJy~beam$^{-1}$. (b) CS~(7--6) integrated intensity image of VV~114E. The flux in color scale ranges from 0~Jy~km~s$^{-1}$ to 0.45~Jy~km~s$^{-1}$. (c) HCN~(4--3) integrated intensity image of VV~114E. The flux in color scale ranges from 0~Jy~km~s$^{-1}$ to 2.0~Jy~km~s$^{-1}$. (d) HCO$^+$~(4--3) integrated intensity image of VV~114E. The flux in color scale ranges from 0~Jy~km~s$^{-1}$ to 2.8~Jy~km~s$^{-1}$. The beam size of each line and continuum is shown in the bottom-left of the images (Table~\ref{table_data}). The red cross shows the position of the eastern nucleus defined by the peak position of the Ks-band observation \citep{tat12}.
}
\label{fig_VV114E}
\end{center}
\end{figure}
\subsubsection{CN~(1$_{3/2}$--0$_{1/2}$) and CN~(1$_{1/2}$--0$_{1/2}$)}
Two radical CN rotational transitions \textit{N} = 1 -- 0 (\textit{J} = 3/2 -- 1/2 and 1/2 -- 1/2) are detected at the eastern nucleus. The $J$ = 3/2 -- 1/2 transition is extended toward the overlap region (Figures~\ref{fig_mom2}d, \ref{fig_mom2}e, \ref{fig_mom2}f, \ref{fig_mom2}g, \ref{fig_mom2}h, and \ref{fig_mom2}i). We can not resolve their multiplet because of the coarse frequency resolution (11.5~MHz $\simeq$ 30~km~s$^{-1}$). Because the critical density of CN is high ($\sim$ 10$^6$~cm$^{-3}$), the CN emission mainly comes from denser gas regions than regions traced by $^{12}$CO~(1--0). The \textit{J} = 3/2 -- 1/2 transition shows a similar distribution to the $^{13}$CO~(1--0) emission, but it is less extended over the overlap region. The total CN~(1$_{1/2}$--0$_{1/2}$) and CN~(1$_{3/2}$--0$_{1/2}$) integrated intensities are 2.0 $\pm$ 0.1~Jy~km~s$^{-1}$ and 5.4 $\pm$ 0.3~Jy~km~s$^{-1}$, respectively. The highest velocity dispersion in the CN~(1$_{3/2}$--0$_{1/2}$) image is also detected between the eastern nucleus and the overlap region, and this is likely caused by a superposition of clouds similar to the case of the $^{13}$CO~(1--0) image (see Appendix ~\ref{A3}).
\subsection{Line Emission in Band 7} \label{co32}
\subsubsection{$^{12}$CO~(3--2)}
The $^{12}$CO~(3--2) emission maps are presented in Figure~\ref{fig_mom3}. The estimated missing flux in our ALMA observation is 21 $\pm$ 1~\% \citep[\textit{James Clerk Maxwell Telescope} (JCMT): 2956 $\pm$ 133~$\rm{Jy~km~s^{-1}}$ and ALMA: 2343.7 $\pm$ 4.7~$\rm{Jy~km~s^{-1}}$;][]{wil08, sai13}. Although our $^{12}$CO~(3--2) observation recovers more flux than the \textit{Submillimeter Array} (SMA) observation \citep[1530 $\pm$ 16~$\rm{Jy~km~s^{-1}}$; the missing flux = 48 $\pm$ 15~\%; ][]{wil08}, there are significant negative sidelobes at the north and south of the image which is likely the cause of missing flux. We made the CLEANed image with a uniform \textit{uv} weighting to minimized the sidelobe level \citep{thomp}.
The $^{12}$CO~(3--2) integrated intensity map of VV~114 (Figure~\ref{fig_mom3}a) shows two arm-like structures and a filamentary structure similar to the $^{12}$CO~(1--0) image, and the strongest peak is at $\simeq$ 5\farcs5 west of the eastern nucleus. The global gas distribution is consistent with the previous $^{12}$CO~(3--2) observations \citep{ion04, wil08}.
The $^{12}$CO~(3--2) velocity field map of VV~114 (Figure~\ref{fig_mom3}b) also shows significant broad velocity range across the galaxy disks ($\simeq$ 600~km~s$^{-1}$), similar to the $^{12}$CO~(1--0) velocity field map. The SE arm has a blue-shifted velocity from 5650~km~s$^{-1}$ to 5920~km~s$^{-1}$, while the NW arm has a red-shifted velocity from 5950~km~s$^{-1}$ to 6160~km~s$^{-1}$. Other two arm-like features are also detected. One located $\simeq$ 4\farcs0 northeast of the eastern nucleus shows an arc around the eastern nucleus and has red-shifted velocities from 5810~km~s$^{-1}$ to 6180~km~s$^{-1}$. This arm coincides with the NE arm detected in the $^{12}$CO~(1--0). The other one located at $\simeq$ 10\farcs0 west of the SE arm has a strong peak ($\simeq$ 262.5 $\pm$ 0.9 ~Jy~km~s$^{-1}$) and blue-shifted velocities from 5610~km~s$^{-1}$ to 5900~km~s$^{-1}$. This arm also coincide with the SW arm detected in the $^{12}$CO~(1--0).
From the $^{12}$CO~(3--2) velocity dispersion map of VV~114 (Figure~\ref{fig_mom3}c), we find that the overlap region has the highest velocity dispersion ($\simeq$ 110~km~s$^{-1}$). The velocity dispersion of the NW arm is $\simeq$ 30~km~s$^{-1}$, while the SE arm is $\simeq$ 60~km~s$^{-1}$.
\begin{figure*}
\begin{center}
\includegraphics[scale=.5]{fig_contin.eps}
\caption{(a) The 110~GHz continuum flux image overlaid on the HST/ACS/F435W image of VV~114. The contours are 0.10, 0.20, 0.40, 0.80, and 1.60~mJy~beam$^{-1}$. The red crosses show the positions of the nuclei defined by the peak positions of the Ks-band observation \citep{tat12}. (b) The low resolution 340~GHz continuum flux image overlaid on the HST/ACS/F435W image of VV~114. The contours are 0.22, 0.44, 0.88, 1.76, 3.52, and 7.04~mJy~beam$^{-1}$. A strong point source at the eastern edge of the image is a distant star-forming galaxy, ALMA J010748.3 -- 173028 \citep[see][]{tam14}. (c) The high resolution 340~GHz continuum flux image overlaid on the HST/ACS/F435W image of VV~114. The contours are 0.14, 0.28, 0.56, 1.12, and 2.24~mJy~beam$^{-1}$. The red cross shows the position of the eastern nucleus defined by the peak position of the Ks-band observation \citep{tat12}. The beam size of each continuum is shown in the bottom-left of the images (Table~\ref{table_data}).
}
\label{fig_contin}
\end{center}
\end{figure*}
\subsubsection{HCN~(4--3) and HCO$^+$~(4--3)}
The HCN~(4--3) and HCO$^+$~(4--3) images are shown in Figures~\ref{fig_mom3}d, \ref{fig_mom3}e, \ref{fig_mom3}f, \ref{fig_mom3}g, \ref{fig_mom3}h, and \ref{fig_mom3}i. While the HCN~(4--3) emission is only seen near the eastern nucleus of VV~114 and is resolved into four peaks, the HCO$^+$~(4--3) emission is more extended and has at least 10 peaks in the integrated intensity map. The total integrated intensities of HCO$^+$~(4--3) and HCN~(4--3) are 15.3 $\pm$ 0.4~Jy~km ~s$^{-1}$ and 4.4 $\pm$ 0.2~Jy~km~s$^{-1}$, respectively. The higher HCO$^+$~(4--3) flux observed with the SMA \citep[17 $\pm$ 2~mJy,][]{wil08} using a 2\farcs8 $\times$ 2\farcs0 beam is likely attributed to missing flux by the ALMA observation. A compact component in the eastern nucleus is unresolved with the current resolution, and the upper limit to the size is 200~pc. The HCN~(4--3) emission is not detected in the overlap region, where both the high $^{12}$CO~(1--0) velocity dispersion and the significant CH$_3$OH~(2$_k$--1$_k$) and HCO$^+$~(4--3) detection suggest the presence of shocked gas \citep{krp08}. We concluded in \citetalias{ion13} from their source size, line widths, and the relative strengths of HCN~(4--3) and HCO$^+$~(4--3) that the unresolved eastern nucleus harbors an obscured AGN, and the dense clumps in the western galaxy are related to extended starbursts.
\subsubsection{CS~(7--6)}
The CS~(7--6) emission has the highest critical density ($n_{\rm{cr}}$ $\simeq$ 10$^7$~cm$^{-3}$) of all of the lines detected in our observations. The CS~(7--6) emission is marginally (S/N $\sim$ 4) detected at the eastern nucleus (Figure~\ref{fig_VV114E}), and the total flux is 0.5 $\pm$ 0.1~Jy~km~s$^{-1}$.
\subsection{Continuum Emission}
The continuum image at 110~GHz shows a filamentary structure similar to the molecular line image (Figure~\ref{fig_contin}a). We construct low resolution (1\farcs33 $\times$ 1\farcs12) and high resolution (0\farcs45 $\times$ 0\farcs38) images of the 340~GHz continuum (Figures~\ref{fig_contin}b, and \ref{fig_contin}c) using the combined data (compact + extended) and the extended configuration data, respectively. We find that the filamentary structure and the unresolved eastern nucleus are both present in dust continuum. The total flux of the 110~GHz and the low resolution 340~GHz continuum emission are 10.3 $\pm$ 0.2~mJy and 38.6 $\pm$ 0.3~mJy, respectively. The estimated missing flux relative to the JCMT 340~GHz observation \citep{wil08} is 75 $\pm$ 4~\% (SMA: 79 $\pm$ 7~\%). The difference in the recovered flux between $^{12}$CO~(3--2) and 340~GHz continuum emission may be caused by the difference in the distribution. The 110~GHz and 340~GHz continuum emission is detected at the eastern nucleus (S/N $\sim$ 50 and 70) and the filamentary structure (S/N $\sim$ 8 and 24) identified in the $^{13}$CO~(1--0) image, both with high significance.
\section{Spatially resolved line ratios} \label{ratio}
We assign 39 ``R" boxes (2\farcs0 $\times$ 2\farcs0; R1 -- R39; see Figure~\ref{fig_ratio}) for the band 3 and $^{12}$CO~(3--2) data and 15 smaller ``S" boxes (1\farcs2 $\times$ 1\farcs2; S0 -- S14; see Figure~\ref{figratio_HCN}) for the rest of the data to estimate the physical parameters, such as the molecular gas mass ($M_{\rm{H_2}}$), dense gas mass ($M_{\rm{dense}}$), dust mass ($M_{\rm{dust}}$), star formation rate (SFR), kinetic temperature ($T_{\rm{kin}}$), gas density ($n_{\rm{H_2}}$), gas column density ($N(\rm{H_2})$), and molecular abundance relative to H$_2$ ([$X$]/[H$_2$]). The positions of the boxes are chosen to cover the CO~(3--2) emission (R1 -- R39) and the HCO$^+$~(4--3) emission (S0 -- S14). The sizes of the boxes are chosen such that they are comparable to the beam size.
Before deriving the parameters and line ratios at each box, we first matched the $uv$ range between our data set and reconstructed the integrated intensity image of each line. The shortest baseline lengths are set to 13.5 k$\lambda$ and 40.0 k$\lambda$ for the molecular lines in the band 3 and the band 7, respectively, and the images are convolved into the same resolution (2\farcs0 $\times$ 1\farcs5 with a P.A. of 83~deg, 1\farcs2 $\times$ 1\farcs0 with a P.A. of 119~deg). For each ratio, the two integrated intensity images were expressed in the units of K~km~s$^{-1}$ before calculating the ratio at locations where both lines are detected above 3~$\sigma$. The derived box-summed spectra are listed in Appendix~\ref{A3}. We carried out a multi Gaussian fit (one - three components) to reproduce the box-summed spectra, and labeled the components as ``a", ``b", and ``c" from the bluest peak (e.g., the bluest peak at R21 is labeled as R21a).
\begin{figure*}
\begin{center}
\includegraphics[scale=.5]{fig_ratio.eps}
\caption{(a) The $R_{3-2/1-0}$ image. The ratio in color scale ranges from 0 to 1. The white crosses show the positions of the nuclei defined by the peak positions of the Ks-band observation \citep{tat12}. (b) The $R_{12/13}$ image. The ratio in color scale ranges from 0 to 40. (c) Locations of 39 boxes (R1 -- R39) that are used to calculate the line ratios and physical parameters. For each ratio, the two integrated intensity images were convolved to the same resolution and expressed in units of K~km~s$^{-1}$ before calculating the ratio at locations where both lines are detected above 3~$\sigma$. The black crosses show the positions of the nuclei defined by the peak positions of the Ks-band observation \citep{tat12}. The beam size of each line ratio is shown in the bottom-left of the images.
}
\label{fig_ratio}
\end{center}
\end{figure*}
\subsection{$^{12}$CO~(3--2)/$^{12}$CO~(1--0), $R_{3-2/1-0}$}
The $^{12}$CO~(3--2)/$^{12}$CO~(1--0) ratio, $R_{3-2/1-0}$, can be used as an indicator of the dense/warm gas content relative to the total molecular gas. The $R_{3-2/1-0}$ of VV~114 varies from 0.2 to 0.8, as shown in Figure~\ref{fig_ratio} (left) and Table~\ref{table_T_R}. This range is larger than the same ratios derived for normal spirals, which is typically 0.15 -- 0.5 when observed with a similar linear resolution \citep{war10}. At the edge and the center of the filament, $R_{3-2/1-0}$ is higher (0.53 -- 0.69) than the highest peaks of each arm ($\sim$ 0.4). This suggests that the CO emitting gas at the filament have higher excitation conditions than normal spirals, while the conditions of each arm of VV~114 are consistent with arms and nuclei of normal spirals. The $R_{3-2/1-0}$ at the eastern nucleus is 0.76 $\pm$ 0.01. It is suggested that the $R_{3-2/1-0}$ is much higher \citep[3.12 $\pm$ 0.03 in NGC 1068;][]{tsa12} for gas surrounding an AGN, and the low $R_{3-2/1-0}$ in VV~114 may be due to the difference in filling factor (160 $\times$ 140~pc beam averaging for NGC~1068, while 800~pc box averaging for VV~114). It is possible, however, that the nuclear excitation conditions are different from source to source.
\subsection{$^{12}$CO~(1--0)/$^{13}$CO~(1--0), $R_{12/13}$}
In general, the $^{12}$CO lines has higher optical depths than the $^{13}$CO~(1--0) line. Therefore, the measured $^{12}$CO~(1--0)/$^{13}$CO~(1--0) line intensity ratio, $R_{12/13}$, gives a lower limit to the CO/$^{13}$CO abundance ratio (hereafter [CO]/[$^{13}$CO]). We present the $R_{12/13}$ image of VV~114 in Figure~\ref{fig_ratio} (center). The $R_{12/13}$ increases from the arms ($<$ 17) to the filament (15 -- 32). Observationally, $R_{12/13}$ increases towards the central region of galaxies \citep{aal95}, where the gas is generally warmer and denser. \citet{aal95} suggest that the moderate optical depth of $^{12}$CO~(1--0) emission and/or the high [CO]/[$^{13}$CO] environment can increase the $R_{12/13}$ in nuclei of U/LIRGs. In order to understand which of the two (optical depths or abundances) is dominant, we calculated and mapped the optical depth of the $^{12}$CO~(1--0) and the $^{13}$CO~(1--0) as shown in Table~\ref{table_LTE} and Figure~\ref{fig_RADEX}. We provide an interpretation of these results in \S\ref{chi}.
\subsection{HCN~(4--3)/HCO$^+$~(4--3), $R_{\rm{HCN/HCO^+}}$}
In \citetalias{ion13}, the HCN~(4--3) and HCO$^+$~(4--3) maps of VV~114 allowed us to investigate the central region at 200 pc resolution for the first time, and we find that both the HCN~(4--3) and HCO$^+$~(4--3) in the eastern nucleus are compact ($<$ 200~pc), and broad [290~km~s$^{-1}$ for HCN~(4--3)]. We present the HCN~(4--3)/HCO$^+$~(4--3), $R_{\rm{HCN/HCO^+}}$, image of VV~114 in Figure~\ref{figratio_HCN}. From the higher $R_{\rm{HCN/HCO^+}}$ along with the past X-ray and NIR observations, we suggest the presence of an obscured AGN in the eastern nucleus. We also detect a 3 -- 4~kpc long filament of dense gas, which is likely to be tracing the active star formation triggered by the ongoing merger, and this is consistent with the results from the numerical model by \citet{tey10} who predict that the fragmentation and turbulent motion of dense gas across the merging disk is responsible for forming dense gas clumps with masses of 10$^6$ -- 10$^8$~M$_{\odot}$.
We present the $R_{\rm{HCN/HCO^+}}$ image in Figure~\ref{figratio_HCN}. The overlap region does not show significant HCN~(4--3) emission, and we provide the 3~$\sigma$ upper limit in Table~\ref{table_T_S}. Three out of the four boxes (i.e., S1 -- S3) in the eastern nucleus have low $R_{\rm{HCN/HCO^+}}$ ($<$ 0.5) whereas S0 has a high $R_{\rm{HCN/HCO^+}}$ (1.34 $\pm$ 0.09). It is suggested that such a high value is only produced around AGN environments \citep[e.g.,][]{khn01, har13, ion13, izm13, ima14}.
\section{Derivation of physical parameters} \label{der}
In this section, we derive the molecular gas mass (\S\ref{X}), and the physical parameters using the radiative transfer code RADEX (\S\ref{chi}) for each box defined in \S\ref{ratio}. The column density is derived using the optically thin $^{13}$CO line under the LTE assumption. We estimate the beam filling factor $\Phi_{\rm{A}}$ and the relative molecular abundance of molecule $X$ (hereafter expressed as [$X$]/[H$_2$]) (\S\ref{LTE}). Finally, we calculate the dust mass using the 340~GHz continuum emission (\S\ref{dust}).
\subsection{Molecular Gas Mass Derivation} \label{X}
The molecular gas mass M$_{\rm{X}}$ is derived by;
\begin{equation}
M_{\rm{X}} = \alpha_{\rm{X}}\:L'_{\rm{X}}\:[\rm{M}_{\odot}],
\end{equation}
where $\alpha_{\rm{X}}$ is the molecular line luminosity-to-H$_2$ mass conversion factor and $L\arcmin_{\rm{X}}$ is the velocity integrated flux \citep{s&v05}. We use the conversion factor known to be appropriate for U/LIRGs \citep[$\alpha_{\rm{CO}}$ = 0.8~M$_{\odot}~\rm{(K~km~s^{-1}~pc^2)^{-1}}$;][]{d&s98}. This is consistent with the value derived by \citet{slw13} in VV~114 ($\alpha_{\rm{CO}}$ = $0.5^{+0.6}_{-0.3}$~M$_{\odot}~\rm{(K~km~s^{-1}~pc^2)^{-1}}$). The molecular gas mass derived at the boxes defined in \S\ref{ratio} ranges between 0.2 $\times$ 10$^8$~$\left(\frac{\alpha_{\rm{CO}}}{0.8}\right)$ and 4.8 $\times$ 10$^8$~$\left(\frac{\alpha_{\rm{CO}}}{0.8}\right)$~M$_{\odot}$ (Table~\ref{table_SF_CO}). We also calculate the dense gas mass $M_{\rm{dense}}$ using $\alpha_{\rm{HCN}}$ = 10~M$_{\odot}~\rm{(K~km~s^{-1}~pc^2)^{-1}}$ \citep{g&s04} and the HCN~(4--3) luminosity which is converted to the HCN~(1--0) luminosity using HCN~(4--3)/HCN~(1--0) = 0.63 \citep[\citetalias{ion13};][]{ima07}. The dense gas mass ranges between 1.8 $\times$ 10$^6$~$\left(\frac{\alpha_{\rm{HCN}}}{10}\right)$ and 3.8 $\times$ 10$^7$~$\left(\frac{\alpha_{\rm{HCN}}}{10}\right)$~M$_{\odot}$ (Table~\ref{table_SF_HCN}).
\begin{figure*}
\begin{center}
\includegraphics[scale=.5]{figratio_HCN.eps}
\caption{(a) The $R_{\rm{HCN/HCO^+}}$ image. The ratio in color scale ranges from 0 to 1. The white cross shows the position of the eastern nucleus defined by the peak position of the Ks-band observation \citep{tat12}. (b) The $R_{\rm{HCN/HCO^+}}$ image near the nucleus of VV~114E. The ratio in color scale also ranges from 0 to 1. (c) Locations of 15 boxes (S0 -- S14) that are used to calculate the line ratios and physical parameters. For each ratio, the two integrated intensity images were convolved to the same resolution and expressed in units of K~km~s$^{-1}$ before calculating the ratio at locations where both lines are detected above 3~$\sigma$. The black cross shows the position of the eastern nucleus defined by the peak position of the Ks-band observation \citep{tat12}. The green open squares are the ``R" boxes shown in Figure \ref{fig_ratio}c.
}
\label{figratio_HCN}
\end{center}
\end{figure*}
We note that the CO luminosity-to-H$_2$ mass conversion factor, $\alpha_{\rm{CO}}$, is very uncertain, and varies significantly from source to source \citep[0.4 -- 0.8 for LIRGs;][]{d&s98, yao03, pap12, bol13}. It may be possible that $\alpha_{\rm{CO}}$ varies from region to region within a galaxy. While one would ideally adopt a spatially varying $\alpha_{\rm{CO}}$ for a better quantification of the H$_2$ mass, such a study is beyond the scope of this present paper. For simplicity, here we adopt a constant $\alpha_{\rm{CO}}$ across all regions in VV~114, bearing in mind that the uncertainties could be as large as a factor of two. The same applies to $\alpha_{\rm{HCN}}$ \citep{g&s04}.
\subsection{Radiative Transfer Analysis using $\rm{RADEX}$} \label{chi}
We used the non-LTE radiative transfer code RADEX \citep{vdt07} and varied the parameters until the residuals between the observed line fluxes and the modeled line fluxes are minimized in a $\chi^2$ sense. We assumed a uniform spherical geometry ($dv$ = 1.0~$\rm{km~s^{-1}}$), and derived the physical conditions of molecular gas ($T_{\rm{kin}}$, $n_{\rm{H_2}}$, and $N(\rm{H}_2)$). RADEX uses an escape probability approximation to solve the non-LTE excitation assuming that all lines are from the same region. Since the molecular lines in the band~7 have significantly higher critical densities than that in the~band 3, we used two sets of molecular lines; (case 1) 2\farcs0 box-summed $^{12}$CO~(1--0), $^{13}$CO~(1--0), and $^{12}$CO~(3--2), and (case 2) 1\farcs2 box-summed HCN~(4--3), HCO$^+$~(4--3), $^{12}$CO~(3--2), and $^{12}$CO~(1--0), to solve for the degeneracy of the physical parameters. In case 2, we made the \textit{uv} and beam-matched HCN~(4--3), HCO$^+$~(4--3), and $^{12}$CO~(3--2) images (1\farcs2 $\times$ 1\farcs0 resolution with the P.A. = 119~deg.), and we defined three HCO$^+$~(4--3) peaks as E0, E1, and E2 (Figure~\ref{fig_spec_RADEX}). We also use the $uv$ and beam-matched $^{12}$CO~(1--0) data to constrain the $N(\rm{H}_2)$, allowing us to vary the [HCN]/[HCO$^+$] in case 2. All line parameters, such as the upper state energies and the Einstein coefficients, were taken from the \textit{Leiden Atomic and Molecular Database} \citep[LAMDA;][]{sco05}. In order to find the set of physical parameters that can reproduce the observed line intensities, we run RADEX by varying $T_{\rm{kin}}$, $n_{\rm{H_2}}$, and $N(\rm{H_2})$ for case 1, and $T_{\rm{kin}}$, $n_{\rm{H_2}}$, and [HCN]/[HCO$^+$] for case 2. The adopted $N(\rm{H_2})$ are 10$^{21.2}$, 10$^{21.6}$, and 10$^{21.5}$~cm$^{-2}$, at E0, E1, and E2, respectively.
We varied the gas kinetic temperature within a range of $T_{\rm{kin}}$ = 5 -- 300~K using steps of d$T_{\rm{kin}}$ = 5~K, and a gas density of $n_{\rm{H_2}}$ = $10^2 $ -- $ 10^5~\rm{cm^{-3}}$ using steps of d$n_{\rm{H_2}}$ = $10^{0.1}~\rm{cm^{-3}}$. For case 1, we fixed [$^{13}$CO]/[H$_2$] = 1.4 $\times$ 10$^{-6}$ \citep{dav13} and [CO]/[$^{13}\rm{CO}$] = 70, which are the Galactic values \citep{w&r94}. In case 2, we changed the parameters, $T_{\rm{kin}}$ = 5 -- 400~K using steps of d$T_{\rm{kin}}$ = 5~K, $n_{\rm{H_2}}$ = $10^3$ -- $10^7~\rm{cm^{-3}}$ using steps of d$n_{\rm{H_2}}$ = $10^{0.1}~\rm{cm^{-3}}$, and fixed [CO]/[H$_2$] = 1.0 $\times$ 10$^{-4}$ and [HCO$^+$]/[H$_2$] = 1.0 $\times$ 10$^{-9}$, which are the standard values observed in Galactic molecular clouds \citep{blk87}. We varied [HCN]/[HCO$^+$] from 1 -- 10, in steps of one. The parameters we used are summarized in Table~\ref{table_radex_parm}. We list the results that are within the 95~\% confidence level with 3-degree of freedom ($\chi^2 < 7.81$) (Tables~\ref{table_RADEX} and \ref{table_RADEX_HCN}). Finally, we created velocity-averaged channel maps of $n_{\rm{H_2}}$ and the optical depth of the transitions (Figure~\ref{fig_RADEX}).
We note that the uncertainty of the $N(\rm{H_2})$ for case 2 did not strongly affect the results, while that of the [CO]/[$^{13}\rm{CO}$] for case 1 changed. The effect of varying the [CO]/[$^{13}\rm{CO}$] will be discussed in \S\ref{1}. Future multi-transition HCN/HCO$^+$/CO/$^{13}$CO imaging will help us to derive these parameters directly.
\begin{figure*}
\begin{center}
\includegraphics[scale=.5]{fig_spec_RADEX.eps}
\caption{(top) The $uv$- and beam-matched (grey color) $^{12}$CO~(3--2), (red contour) HCN~(4--3), and (blue contour) HCO$^+$~(4--3) images. The integrated intensity of the $^{12}$CO~(3--2) in color scale ranges from 0~Jy~km~s$^{-1}$ to 100~Jy~km~s$^{-1}$. The contours are 5, 10, 20, 40, and 50~Jy~km~s$^{-1}$ for HCN~(4--3), and 12, 24, 48, and 96~Jy~km~s$^{-1}$ for HCO$^+$~(4--3). The white cross shows the position of the eastern nucleus defined by the peak position of the Ks-band observation \citep{tat12}. (bottom) 1\farcs2 box-summed spectra of (black line) $^{12}$CO~(3--2), (red dashed line) HCN~(4--3) $\times$ 10, and (blue dashed line) HCO$^+$~(4--3) $\times$ 10 at the each box, labeled E0 -- E2. The spectra are taken from the ALMA data cubes after correcting the cubes for the primary beam attenuation and convolving them to 1\farcs2 $\times$ 1\farcs0 resolution (P.A. = 119 deg.).}
\label{fig_spec_RADEX}
\end{center}
\end{figure*}
\subsubsection{Case 1} \label{1}
The (box-averaged) kinetic temperature near the eastern nucleus (R21a) is constrained to within 25 -- 90~K (the best fit is 50~K), as shown in Table~\ref{table_RADEX}. The $T_{\rm{ex}}$ (58.8 $\pm$ 2.9~K) obtained from the LTE assumption at R21a (see \S\ref{LTE}) is higher than the best-fitted $T_{\rm{kin}}$. In fact, we also find five regions (R10b, R11b, R14, R16, and R25a) that show similarly high excitation temperatures. Four out of five regions are in the central filament. In general, spontaneous emission dominates over collisional excitation in sub-thermally excited conditions, and hence $T_{\rm{ex}}$ should be lower than $T_{\rm{kin}}$. One reason for this discrepancy could be attributed to the incorrect assumption of [CO]/[$^{13}$CO]. By varying this abundance ratio, we find that the temperature reversal (i.e. $T_{\rm{kin}} > T_{\rm{ex}}$) occurs only when [CO]/[$^{13}$CO] $>$ 150. This is consistent with the results obtained by \citet{slw13} who used RADEX along with their multi CO and $^{13}$CO line data to find evidence of a cold/dense molecular gas component with extremely high [CO]/[$^{13}$CO] of 229, which is 3 times higher than that of the Galactic value \citep{w&r94}.
The derived $T_{\rm{kin}}$ at the other regions are generally higher than 100~K. The derived $T_{\rm{ex}}$ in each region are typically 10 -- 40~K, which may suggest sub-thermal conditions. The kinetic temperatures derived at the SE and NW arms are estimated to be $<$ 90 K, with higher temperature at the NW arm. The NW arm is also associated with relatively strong Pa$\alpha$ emission and $K$s-band emission, which is consistent with the higher relative temperature due to star-forming activities \citep{mnm08}. However, this is inconsistent with the general understanding that strong tidal shear in tidal arms prevents active star formation to occur \citep{aal10}.
The derived $n_{\rm{H_2}}$ in most of the boxes are less than 10$^{3.0}$~cm$^{-3}$, which is consistent with the critical densities of the low-$J$ CO lines observed here. The highest density of 10$^{3.4}$ -- 10$^{5.0}$~cm$^{-3}$ is estimated at R21a, and this is consistent with the location of the eastern nucleus. Since we also observed the strongest HCN~(4--3) and HCO$^+$~(4--3) emission at R21a at the same line-of-sight velocity \citep[see also Appendix~\ref{A3}]{ion13}, it is possible that the main contribution to the CO emission at R21a arises from dense gas (10$^{3.4}$ -- 10$^{5.0}$~cm$^{-3}$) near the eastern nucleus, with a minor contribution from the diffuse gas clouds along the same line of sight observed within the same beam. In contrast to the eastern nucleus, the boxes that cover the western galaxy (R1 -- R11 and R26 -- R29) show moderately dense condition of 10$^{2.0}$ -- 10$^{4.0}$~cm$^{-3}$. This extended and moderately dense gas is associated with the disk-like structure seen in optical images \citep{eva08}, and the star formation traced in Pa$\alpha$ emission and UV/X-ray emission \citep{grm06, tat12}. We note that the strongest off-nuclear Pa$\alpha$ peak (R27 in Table~\ref{table_SF_CO}; SFR = 3.15 $\pm$ 0.05~M$_{\odot}$~yr$^{-1}$) coincides with relatively low gas density ($\sim$ 10$^{3.0}$~cm$^{-3}$). The density of the surrounding region labeled R25a is similar (10$^{3.5}$ -- 10$^{5.0}$~cm$^{-3}$) and this is comparable to the nucleus of the eastern galaxy. The secondary Pa$\alpha$ peak (R29; SFR = 0.92 $\pm$ 0.05~M$_{\odot}$~yr$^{-1}$) is not associated with any molecular line emission.
It is usually believed that the $^{12}$CO~(1--0) emission is optically thick ($\tau_{\rm{CO}} \gg 1$), while the $^{13}$CO~(1--0) emission is optically thin ($\tau_{\rm{^{13}CO}} \ll 1$) even in luminous mergers \citep{dav13}. In most regions, we find that the optical depth of the $^{12}$CO~(1--0) line is $\gg$ 1 (Figure~\ref{fig_RADEX}). In contrast, the $^{12}$CO~(1--0) opacity at the eastern nucleus and the filament is moderately optically thick ($\tau_{\rm{CO}}$ $\sim$ 1). However, the elevated $R_{12/13}$ at the eastern nucleus (see \S\ref{ratio}) cannot be explained by the relatively low $^{12}$CO~(1--0) opacity alone (the opacity has to be $\tau_{\rm{CO}}$ $\ll$ 0.1; see also \citet{TORA}). Finally, we find that indeed the $^{13}$CO~(1--0) emission is optically thin ($\tau_{\rm{^{13}CO}} \ll 1$) averaged over the whole galaxy, except for the southern dust lane ($\tau_{\rm{^{13}CO}}$ = 0.3 -- 1.5).
\begin{figure*}
\begin{center}
\includegraphics[scale=.5]{fig_RADEX.eps}
\caption{The channel maps of the box-averaged RADEX modeling based on the $^{12}$CO~(1--0), $^{13}$CO~(1--0), and $^{12}$CO~(3--2). (top) The best fitted values of logarithmic gas density of the $^{12}$CO~(1--0) emission. The value in color scale ranges from 2.0 to 4.0~cm$^{-3}$. (bottom) The best fitted values of optical depth of the $^{12}$CO~(1--0) emission. The value in color scale ranges from 0 to 9. The black crosses show the positions of the nuclei defined by the peak positions of the Ks-band observation \citep{tat12}. The open squares are regions which we cannot solve the RADEX calculations because of non-detection of the $^{12}$CO~(1--0), $^{13}$CO~(1--0), or $^{12}$CO~(3--2) emission.}
\label{fig_RADEX}
\end{center}
\end{figure*}
From these results, we suggest that the peak of the molecular gas in the central 800~pc of the eastern galaxy is cold ($T_{\rm{kin}}$ = 25 -- 90~K), dense ($n_{\rm{H_2}}$ = 10$^{3.4}$ -- 10$^{5.0}$~cm$^{-3}$), and moderately optically thick ($\tau_{\rm{CO(1-0)}}$ $\sim$ 3), while peaks in the overlap region are warm ($T_{\rm{kin}}$ $>$ 50~K, best-fitted $T_{\rm{kin}}$ is 95 and 175~K at R39a and R39b, respectively), moderately dense ($n_{\rm{H_2}}$ = 10$^{2.3}$ -- 10$^{4.1}$~cm$^{-3}$), and moderately optically thick ($\tau_{\rm{CO(1-0)}}$ $\sim$ 1). The derived density of the eastern galaxy is slightly higher than the range of values found in U/LIRGs using low-$J$ CO emission with$\sim$ kpc resolution \citep[$n_{\rm{H_2}}$ = 10$^{2.3}$ -- 10$^{4.3}$~cm$^{-3}$;][]{d&s98}. In addition, the low opacities predicted from these analyses are consistent with earlier results that investigate the opacities in M82 \citep[$\tau$ = 0.5 -- 4.5;][]{mao00} and U/LIRGs \citep[$\tau$ = 3 -- 10;][]{d&s98}, and the central region of NGC 6240 \citep[$\tau$ = 0.2 -- 2;][]{ion07}. However, the derived temperature of the eastern galaxy is inconsistent with the high values found in nearby starburst galaxies M82, NGC 253, and NGC 6240 \citep{wld92, jac95, s&f00, ion07}. The disagreement is possibly due to the uncertainties in the [CO]/[$^{13}$CO], or the difference in the observed molecular gas tracers.
\subsubsection{Case 2} \label{2}
The values for $T_{\rm{kin}}$, $n_{\rm{H_2}}$, and the optical depth of HCN~(4--3) and HCO$^+$~(4--3) are shown in Table~\ref{table_E}. The derived parameters for the unresolved component E0, are $T_{\rm{kin}}$ $>$ 100~K, $n_{\rm{H_2}}$ = 10$^{5.0}$ -- 10$^{5.4}$~cm$^{-3}$, and [HCN]/[HCO$^+$] $>$ 5. The lower limit to the kinetic temperature is higher than those of E1 and E2, mainly due to the unusually high $R_{\rm{HCN/HCO^+}}$ and $R_{\rm{HCN/CO}}$. In contrast to E0, the derived parameters near E1 show high H$_2$ densities ($n_{\rm{H_2}}$ = 10$^{5.6}$ -- 10$^{5.9}$~cm$^{-3}$). The overlap region (E2), where the star-formation rate (1.70 $\pm$ 0.05~M$_{\odot}$~yr$^{-1}$) is lower than the eastern nucleus, has densities in the range of $n_{\rm{H_2}}$ = 10$^{5.0}$ -- 10$^{5.6}$~cm$^{-3}$. Finally, the optical depths for the HCO$^+$~(4--3) and HCN~(4--3) lines are calculated for each gas clump, yielding $\tau_{\rm{HCN}}$ $\simeq$ 0.7 and $\tau_{\rm{HCO^+}}$ $\simeq$ 0.2 for E0, $\tau_{\rm{HCN}}$ $\simeq$ 0.2 and $\tau_{\rm{HCO^+}}$ $\simeq$ 0.6 for E1, and $\tau_{\rm{HCN}}$ $\simeq$ 0.4 and $\tau_{\rm{HCO^+}}$ $\simeq$ 0.4 for E2.
The higher linear resolution observations of HCN~(4--3) and HCO$^+$~(4--3) toward NGC~1097 \citep{izm13} revealed that the gas in the central region of NGC 1097 has $T_{\rm{kin}}$ = 70 -- 550~K and $n_{\rm{H_2}}$ = 10$^{4.5}$ -- 10$^{6.0}$~cm$^{-3}$. Moreover, by comparing to LVG models, \citet{krp08} found that HCN and HCO$^+$ emission in AGN-dominated sources appears to emerge from regions with lower H$_2$ densities, higher temperatures, and higher HCN abundance relative to starburst-dominated (SB-dominated) galaxies. Our results obtained toward VV~114 are consistent with these previous results.
\subsection{Filling factor and Column Density under LTE} \label{LTE}
In order to determine the bulk properties of the CO emitting gas, we used an excitation temperature analysis \citep{dav13}. The excitation temperature at each box can be calculated from
\begin{equation}
T_{\rm{ex}} = T_0\:\left(\ln\left[\left(\frac{T_{\rm{b, CO(1-0)}}}{\Phi_{\rm{A}}T_0(1 - e^{-\tau_{\rm{CO}}})} + \frac{1}{e^{T_0/T_{\rm{bg}}} - 1} \right)^{-1} + 1\right] \right)^{-1}
\end{equation}
where $T_0 = h\nu/k$ [= 5.53~K for $^{12}$CO~(1--0) emission], $\nu$ is the frequency of the transition, $h$ is the Planck's constant, $k$ is the Boltzmann's constant, $T_{\rm{b, CO(1-0)}}$ is the brightness temperature of $^{12}$CO~(1--0) emission in Kelvin, $\tau_{\rm{CO}}$ is the optical depth of the $^{12}$CO~(1--0) emission, and $T_{\rm{bg}}$ is the cosmic microwave background temperature (2.73~K). Using $T_{\rm{kin}}$ estimated from the RADEX calculation (\S\ref{chi}), we estimate the beam filling factor $\Phi_{\rm{A}}$,
\begin{equation}
\Phi_{\rm{A}} = \frac{T_{\rm{b, CO(1-0)}}}{T_{\rm{kin}}}
\end{equation}
The optical depth of the $^{12}$CO~(1--0) emission is also estimated from the RADEX calculation in \S\ref{chi}. Assuming that the $^{13}$CO and CO arise from the same molecular cloud, and that the $^{12}$CO~(1--0) is optically thick, we estimate the optical depth of a given molecule using,
\begin{equation}
\tau_{\rm{X}} \simeq \ln\left[\left(1 - \frac{T_{\rm{b}, X}}{T_{\rm{b, CO(1-0)}}} \right)^{-1}\right]
\end{equation}
where $\tau_{\rm{X}}$ is the optical depth of a given transition, and $T_{\rm{b, X}}$ is the observed brightness temperature for transition X. Using $T_{\rm{ex}}$ and $\tau_{\rm{X}}$, we estimate the column density for a given molecule from,
\begin{displaymath}
N_{\rm{X}} = \frac{3k}{8\pi^3\mu^2B(J + 1)}\frac{\exp\left(\frac{2hJ(J + 1)}{kT_{\rm{ex}}} \right)}{\left(1 - \exp\left(-\frac{h\nu}{kT_{\rm{ex}}} \right) \right)}
\end{displaymath}
\begin{equation}
\times\:\frac{\tau_{\rm{X}}}{1 - e^{-\tau_{\rm{X}}}}\frac{1}{J(T_{\rm{ex}}) - J(T_{\rm{bg}})}\int T_{\rm{R}}^{*}dV
\end{equation}
\begin{equation}
J(T) = \frac{h\nu}{k}\frac{1}{\exp(h\nu/kT) - 1}
\end{equation}
where $\mu$ is the dipole moment, $B$ is the rotational constant, $J$ is the lower energy level, and $\int T_{\rm{R}}^{*}dV$ is the integrated intensity \citep{TORA}. The derived column densities are listed in Tables~\ref{table_LTE} and \ref{table_chem}.
\subsection{Dust Mass and ISM Mass Derivation from 340~GHz continuum} \label{dust}
We calculated the dust mass from the 340~GHz (880~$\mu$m) continuum emission (Table~\ref{table_contin}) using \citep{wil08},
\begin{equation}
M_{\rm{dust}} = 74220\:S_{340}\:D_\mathrm{L}^2\frac{e^{\frac{17}{T_d}} - 1}{\kappa_{340}}\:{\rm{M}_{\odot}}
\end{equation}
where $S_{340}$ is the 340~GHz flux in Jy and $D_{\rm{L}}$ is the luminosity distance in Mpc. We assumed a dust emissivity, $\kappa_{340} = 0.9~\rm{cm^2~g^{-1}}$, and the dust temperature $T_{\rm{d}}$ of 39.4~K \citep{wil08}. The box-summed dust masses ranges between 2.0 $\times$ 10$^4$~$\left(\frac{0.9}{\kappa_{340}}\right)$ and 2.8 $\times$ 10$^6$~$\left(\frac{0.9}{\kappa_{340}}\right)$~M$_{\odot}$ (Table~\ref{table_SF_CO}). We note that we used the \citet{d&l84} dust model for $\kappa_{340}$, because the $\kappa_{340}$ derived from observations has a large error \citep{hnn95}.
\citet{scv14} suggested that the submillimeter continuum emission traces the total ISM mass ($M_{\rm{ISM}}$), since the long wavelength Rayleigh-Jeans (RJ) tail of thermal dust emission is often optically thin. In order to compare the $M_{\rm{ISM}}$ with the $M_{\rm{H_2}}$ (see \$\ref{X}) using spatially-resolved data, we calculated the total ISM mass from the 340~GHz continuum emission \citep{scv14}. For $\nu_{\rm{rest}}$ $\lesssim$ 1199~GHz,
\begin{displaymath}
S_{\rm{\nu_{obs}}} = 0.83\frac{M_{\rm{ISM}}}{10^{10}\:\rm{M}_{\odot}}(1 + z)^{4.8}\left(\frac{\nu_{\rm{obs}}}{353~\rm{GHz}}\right)^{3.8}
\end{displaymath}
\begin{equation}
\times\frac{\Gamma_{\rm{RJ}}}{\Gamma_0}\left(\frac{\rm{Gpc}}{D_{\rm{L}}}\right) \rm{mJy}
\end{equation}
where $S_{\rm{\nu_{obs}}}$ is the observed flux, $M_{\rm{ISM}}$ is the ISM mass, $\nu_{\rm{obs}}$ is the observed frequency, and $\Gamma_{\rm{RJ}}$ and $\Gamma_0$ are given by
\begin{equation}
\Gamma_{\rm{RJ}}(T_d, \nu_{\rm{obs}}, z) = \frac{h\nu_{\rm{obs}}(1 + z)/kT_{d}}{e^{h\nu_{\rm{obs}}(1 + z)/kT_d} - 1}
\end{equation}
\begin{equation}
\Gamma_0 = \Gamma_{\rm{RJ}}(T_d, 353~\rm{GHz}, 0).
\end{equation}
The derived box-summed ISM masses of VV~114 range between 5.2 $\times$ 10$^7$ and 7.2 $\times$ 10$^8$ M$_{\odot}$ (Table~\ref{table_SF_CO}). This is comparable to the box-summed H$_2$ masses ($M_{\rm{H_2}}$ = (0.2 -- 4.7) $\times$ 10$^8$~$\left(\frac{\alpha_{\rm{CO}}}{0.8}\right)$ M$_{\odot}$). We find that the $M_{\rm{ISM}}$/$M_{\rm{H_2}}$ ratio is close to unity (0.5 -- 2.0, the average $M_{\rm{ISM}}$/$M_{\rm{H_2}}$ = 0.9 $\pm$ 0.1), while the total $M_{\rm{ISM}}$/$M_{\rm{H_2}}$ ratio is 0.6 $\pm$ 0.1. This means that the spatially-resolved $M_{\rm{ISM}}$ is a good tracer of the ``resolved" H$_2$ mass. However, the total $M_{\rm{ISM}}$ underestimates the H$_2$ mass (even using the $\alpha_{\rm{CO}}$ for ULIRGs to derive the $M_{\rm{H_2}}$) because the global distribution of the 340~GHz continuum emission is significantly different from that of the CO~(1--0) emission (Figures~\ref{fig_mom1} and \ref{fig_contin}). This difference between the 340~GHz continuum and the CO~(1--0) is also seen in recent observations of nearby LIRGs \citep[e.g.,][]{sak14}.
\section{Discussion} \label{dis}
\subsection{Conditions of ``Dense" Gas near the Eastern Nucleus} \label{dense}
Our RADEX modeling yields lower molecular gas density near the AGN ($n_{\rm{H_2}}$ = 10$^{5.0}$ -- 10$^{5.4}$~cm$^{-3}$) compared to the surrounding clumps (10$^{5.6}$ -- 10$^{5.9}$~cm$^{-3}$). Similarly high values are obtained near AGNs in other galaxies \citep{aln02, wil03, krp08}. \citet{krp08} suggest that the gas densities in AGN host galaxies ($<$ 10$^{4.5}$~cm$^{-3}$) are lower than starburst host galaxies (10$^{5.0}$ -- 10$^{6.5}$~cm$^{-3}$), and a common interpretation relies on a clumpy ISM near star-forming regions (which reduces the filling factor) and a continuous ISM near the AGN. Since our current observations ($\sim$ 200 pc resolution) cover a significantly large area and the beam filling factor may be small ($\Phi_{\rm{A}}$ at E0, E1, and E2 are $\lesssim$ 0.03, 0.04 -- 0.06 and 0.01 -- 0.04, respectively), higher resolution observations ($<$ 0\farcs5) are required to confirm this scenario.
\begin{figure*}
\begin{center}
\includegraphics[scale=.7]{KS_law_all.eps}
\caption{(a) The Kennicutt-Scimidt law of VV~114 overlaid on other galaxies. Filled squares show regions of VV~114, while crosses show galaxies in the sample of \citet{ken98}. The dashed line and the dotted line indicate the ``starburst" sequence and ``normal disk" sequence, respectively \citep{dad10}. (b) Distribution of the box-averaged gas depletion time ($\tau_{\rm{gas}}$ = $\Sigma_{\rm{H_2}}/\Sigma_{\rm{SFR}}$). The gas depletion time in color scale ranges from 0 - 1 Gyr. The black crosses show the positions of the nuclei defined by the peak positions of the Ks-band observation \citep{tat12}. The open square is a region which we cannot estimate the gas depletion time because of non-detection of the Pa$\alpha$ emission \citep{tat12}. (c) The Keniccutt-Schmidt law with the $R_{3-2/1-0}$. The ratio in color scale ranges from 0 to 0.8. (d) The Keniccutt-Schmidt law with the $R_{12/13}$. The ratio in color scale ranges from 0 to 40.
}
\label{fig_KS}
\end{center}
\end{figure*}
In addition, our modeling shows higher [HCN]/[HCO$^+$] near the eastern nucleus ($>$ 5) than that in the surrounding clumps ($<$ 4) and the overlap region (1 - 9). The elevated [HCN]/[HCO$^+$] is explained by two mechanisms \citep{krp08}. One is far-UV radiation from OB stars in young starbursts \citep{s&d95}, and the other is strong X-ray radiation from an AGN \citep{mal96}. Because of different penetrating lengths between far-UV and X-ray emission, photon dominated regions (PDRs) are created at the surface of gas clouds and X-rays penetrate deeply into the circumnuclear disk (CND), forming large X-ray dominated regions. As a consequence of this volume versus surface effect, the X-ray radiation from an AGN may produce higher HCN abundances than the UV radiation of starburst activities \citep{krp08}. To some degree, ionization effects from cosmic rays \citep{wld92} such as supernovae or strong shocks are suspected to significantly increase the HCO$^+$ abundance while potentially decreasing the HCN abundance, thus yielding lower $R_{\rm{HCN/HCO^+}}$ in evolved starbursts than in AGNs. The high [HCN]/[HCO$^+$] near the eastern nucleus and low [HCN]/[HCO$^+$] and strong/extended 8~GHz continuum detection at the surrounding clumps \citep{cnd91} are all consistent with a presence of an AGN in the eastern nucleus, surrounded by star-forming dense clumps.
\subsection{Spatially Resolved Kennicutt-Schmidt Law} \label{SFR}
Observational studies of galaxies at global scales have shown that the surface density of SFR and that of cold gas traced in CO~(1--0) obey a power law relation \citep[KS law;][]{scm59, ken98}. ULIRGs are systemically shifted from the normal galaxy population in the $\Sigma_{\rm{SFR}}$ -- $\Sigma_{\rm{H_2}}$ phase \citep{kmg05, dad10, gen10, ler13}. It is suggested that systems lower in IR luminosity (e.g., LIRGs) occupy the region between the ``starburst" sequence and the ``normal disk" sequence in the KS law. Galaxies in the ``starburst" sequence have shorter gas depletion time ($\tau_{\rm{gas}}$ = $\Sigma_{\rm{H_2}}$/$\Sigma_{\rm{SFR}}$ $\sim$ 0.1~Gyr) relative to galaxies in the ``normal disk" sequence \citep[$\tau_{\rm{gas}}$ $\sim$ 1 Gyr;][]{dad10, bou11}. The spatially resolved surface densities of the SFR and the molecular gas mass of VV~114 are shown in Table~\ref{table_SF_CO} and Figure~\ref{fig_KS}. The star-forming regions of VV~114 fill the gap between the ``normal disk" and ``starburst" sequences (Figure~\ref{fig_KS}a). We also show the spatial distribution of $\tau_{\rm{gas}}$ in Figure~\ref{fig_KS}b. The data points close to the ``starburst" sequence are located along the eastern nucleus ($<$ 0.2~$\left(\frac{\alpha_{\rm{CO}}}{0.8}\right)$~Gyr) and the overlap region (= 0.2 -- 0.4~$\left(\frac{\alpha_{\rm{CO}}}{0.8}\right)$~Gyr), while those near the ``normal disk" sequence are located in the NW and SE arms ($>$ 0.8~$\left(\frac{\alpha_{\rm{CO}}}{0.8}\right)$~Gyr). The spatial distribution of $\Sigma_{\rm{SFR}}$ and $\Sigma_{\rm{H_2}}$ are consistent with the distributions of previous optical, UV, and X-ray studies \citep{aln02, lfl02, grm06}. Regions with higher $\Sigma_{\rm{SFR}}$ and $\Sigma_{\rm{H_2}}$ clearly show higher $R_{3-2/1-0}$ and $R_{12/13}$ (Figures~\ref{fig_KS}c and \ref{fig_KS}d).
In summary, transition from the ``normal disk" to ``starburst" sequence may occur when the molecular clouds become excited and dense at the nuclei and the overlap region. Moreover, gas clouds with high $R_{3-2/1-0}$ have high $\Sigma_{\rm{SFR}}$ -- $\Sigma_{\rm{H_2}}$, and this is consistent with past studies which suggest that the $R_{3-2/1-0}$ correlates with the local H$\alpha$ flux \citep{mnm08, fji14}. The $R_{12/13}$ also shows a similar trend, and this is also consistent with the past studies \citep[$>$ 20 in central kpc regions of U/LIRGs, 10 -- 15 in normal starburst galaxies, and $\sim$ 5 in Galactic GMCs;][]{aal97}: The reason for the elevated $R_{12/13}$ in starburst regions of VV~114 will be discussed in detail in \$\ref{isotope}.
\subsection{CO Isotope Ratio Enhancement in the Molecular ``Filament"} \label{isotope}
We suggest from our RADEX modelings that the eastern nucleus and the overlap region have extremely high [CO]/[$^{13}$CO] ($>$ 200), which is at least two times higher than the Galactic value \citep[$\simeq$ 70;][]{w&r94}. The Pa$\alpha$ peaks roughly coincide with the regions where high [CO]/[$^{13}$CO] are expected, suggesting that the increased [CO]/[$^{13}$CO] is related to the star formation activity. Similarly high values are seen in the overlap region of NGC4038/9 \citep{wil03} and the Taffy \citep{zhu07}. \citet{zhu07} suggested that the extreme [CO]/[$^{13}$CO] value in the bridge is explained by three scenarios, 1) selective isotope photodissociation in the diffuse clouds and shocked region, 2) CO enrichment around starburst activities, and/or 3) the destruction and recombination of molecules after shock. We briefly explain each scenario below, but our current data is insufficient for us to identify the exact cause of the high [CO]/[$^{13}$CO] in VV~114.
The first possibility of [CO]/[$^{13}$CO] enhancement is the deficiency in $^{13}$CO. \citet{she92} suggest that selective isotope photodissociation can reduce the $^{13}$CO abundance in diffuse clouds, because CO is self-shielded to a greater extent. Thus, the ISM surrounding young starbursts and/or shocked regions show elevated [CO]/[$^{13}$CO] \citep{zhu07}. The ISM in the nuclei and the overlap region of VV~114 show extremely high [CO]/[$^{13}$CO], presumably due to intense starburst activities and/or large-scale shocks.
The second possibility is that massive stars end their life as supernovae and expel a large amount of $^{12}$C in the interstellar medium. While the elemental abundances (e.g. C and S) are not directly related to the molecular abundances \citep[e.g., CS;][]{cas92}, once the synthesized elements are dispersed in the interstellar medium, molecules \citep[e.g., CO, CS, and CN;][]{hen14} form as soon as the temperature and density conditions are favorable. This occurs with a timescale of a few 10$^5$ yr \citep{l&g89}.
For the overlap region, the destruction and recombination of molecules after shocks (see \S\ref{chemical}) are possible mechanisms to enhance the [CO]/[$^{13}$CO] (the third possibility). The recombination timescale of H$_2$ and CO molecules after shock destruction are shorter than that of $^{13}$CO, since ionized photons from shocked regions lead to selective isotope photodissociation \citep{zhu03}. Shielded regions from the radiation field are needed to form rare $^{13}$CO (Abundant CO can form self-shielded regions). Moreover, the rare isotope molecules generally need a longer time to form, because collisions between molecules and dust grains are less frequent \citep{zhu07}.
\begin{figure}
\begin{center}
\includegraphics[scale=.3]{gas-to-dust.eps}
\caption{Distribution of the box-averaged gas-to-dust ratio map. The black crosses show the positions of the nuclei defined by the peak positions of the Ks-band observation \citep{tat12}. The open squares are regions which we cannot estimate the gas-to-dust ratio because of non-detection of the 340~GHz continuum emission.}
\label{plot_G/D}
\end{center}
\end{figure}
\subsection{Gas-to-Dust Ratio, $M_{\rm{H_2}}/M_{\rm{dust}}$} \label{G/D}
The gas-to-dust ratio, $M_{\rm{H_2}}$/$M_{\rm{dust}}$, provides an important measure of the relative abundance between gas and metallicity. The average $M_{\rm{H_2}}$/$M_{\rm{dust}}$ over the entire galaxy is often derived in single-dish work, and typical $M_{\rm{H_2}}$/$M_{\rm{dust}}$ is 200 -- 300 for local U/LIRGs \citep{c&c03, yao03, sea04}, and 15 -- 231 in high-z sources \citep{s&v05}. \citet{wil08} found $M_{\rm{H_2}}$/$M_{\rm{dust}}$ = 357 $\pm$ 95 from a sample of 13 U/LIRGs, including VV~114, observed at kpc resolution.
We use the gas and dust masses derived in \S\ref{dust} to investigate the distribution of $M_{\rm{H_2}}$/$M_{\rm{dust}}$ (Figure~\ref{plot_G/D}). The smallest value of (128 $\pm$ 16)~$\left(\frac{\alpha_{\rm{CO}}}{0.8}\right)\left(\frac{0.9}{\kappa_{340}}\right)$ occurs in the eastern nucleus, which is similar to the Galactic value \citep[100;][]{hil83}, while higher values of (371 $\pm$ 118)~$\left(\frac{\alpha_{\rm{CO}}}{0.8}\right)\left(\frac{0.9}{\kappa_{340}}\right)$ and (339 $\pm$ 60)~$\left(\frac{\alpha_{\rm{CO}}}{0.8}\right)\left(\frac{0.9}{\kappa_{340}}\right)$ occur in the western nucleus and the overlap region, respectively. The clear differences between the two nuclei may suggest a local gradient in the metallicity. For the overlap region, cold dust associated with diffuse gas clouds cannot avoid the collision. This tends to increase the $M_{\rm{H_2}}$/$M_{\rm{dust}}$, because shocks destruct dust particles preferentially \citep{zhu07}. On the other hand, the low $M_{\rm{H_2}}$/$M_{\rm{dust}}$ in the eastern nucleus may be due to intense starbursts producing dust-rich environments.
\subsection{Fractional Abundances of CS, CH$_3$OH, and CN} \label{chemical}
Table~\ref{table_chem} shows the properties of the detected molecular lines which are not used in the RADEX calculations. Either the dense gas component of VV~114 has extreme variations in excitation among the molecular clumps in the filament (see \S\ref{dense}), or there is widespread chemical differentiation across the filament. The fractional abundances [$N_{\rm{X}}/N_{\rm{H_2}}$] of the different astrochemical species provide evidence of varying chemical influences due to star formation, physical conditions, and dynamics across the galaxy disks. We use the H$_2$ column densities, derived from the RADEX calculations, which are 10$^{20.8}$, 10$^{21.1}$, and 10$^{21.1}$~cm$^{-2}$ at R18 (AGN), R21a (starburst), and R39a (overlap region), respectively. Column densities of each molecules are determined by equation (7) assuming an optically thin emission under LTE. The $T_{\rm{ex}}$ values determined from equation (3) are 38.7 $\pm$ 1.9~K, 58.8 $\pm$ 2.9~K, and 52.6 $\pm$ 2.6~K at R18, R21a, and R39a, respectively. The derived [$N_{\rm{X}}/N_{\rm{H_2}}$] are listed in Table~\ref{table_chem}.
\begin{figure}
\begin{center}
\includegraphics[scale=.3]{mol_abu2.eps}
\caption{Logarithmic fractional abundances relative to H$_2$ ([$X$]/[H$_2$]) of selected extragalactic sources compared to those of specific regions of VV~114, as presented in Table~\ref{table_chem}. The red, green, and blue columns show R18, R21a, and R39a, respectively. The pink, light blue, and yellow columns show NGC~253, M~82, and IC~342, respectively. Arrows represent upper limits.}
\label{plot_molabu}
\end{center}
\end{figure}
In Figure~\ref{plot_molabu}, we show the fractional abundances for CS, CH$_3$OH, and CN in VV~114, and the same ratios for a sample of nearly galaxies, NGC 253, M82, and IC 342, taken from line surveys available in the literature \citep{hen88, mau89, hut97, mar06}. M82 has a relatively old starburst at its core, with an average stellar population age of $\simeq$ 10~Myr \citep{kon09}. This creates strong UV fields, therefore the PDR dominates its chemistry \citep{ala11}. Figure~\ref{plot_molabu} shows that R39a has higher CH$_3$OH abundance than M82, and small CS and CN abundances. A pure PDR similar to M82 may explain the molecular abundances we observe in R18.
The molecular abundances for the overlap region and NGC 253 share similar characteristics. NGC 253 is thought to be in an early stage of starburst evolution, and has young stellar populations in its nucleus \citep[$\simeq$ 6~Myr;][]{fer09}. The chemistry in the nucleus of NGC 253 is dominated by large-scale shocks \citep{ala11}, and we suggest that the overlap region of VV~114 is also dominated by shocks. The low $R_{\rm{HCN/HCO^+}}$ at the overlap region are further evidences for a shock dominated region \citep{krp08}.
\subsection{Merger-driven Tidal Dwarf Galaxy Formation} \label{TDG}
Tidal dwarf galaxies (TDGs) are gas-rich irregular galaxies made out of stellar and gaseous material pulled out by tidal forces from the disks of the colliding parent galaxies into the intergalactic medium. They are found at the ends of long tails and host active star-forming regions \citep{bra00}. \citet{hib01} and \citet{gao01} found the HI gas mass of 4.1 $\times$ 10$^8$ M$_{\odot}$ and the molecular gas mass of 4 $\times$ 10$^6$ M$_{\odot}$ at the edge of the southern tail of NGC 4038/9.
We found an elevated $R_{3-2/1-0}$ (0.36 $\pm$ 0.01), SFR (0.10 $\pm$ 0.05~M$_{\odot}$~yr$^{-1}$), and $M_{\rm{H_2}}$ ($\sim$ 3.8 $\times$ 10$^7$~$\left(\frac{\alpha_{\rm{CO}}}{0.8}\right)$ M$_{\odot}$) at the edge of the southern tidal arm (R38). The derived SFR and $M_{\rm{H_2}}$ of R38 are comparable to those of TDG candidates in other galaxies \citep{bra01}. The gas depletion time of (0.40 $\pm$ 0.22)~$\left(\frac{\alpha_{\rm{CO}}}{0.8}\right)$~Gyr is shorter than the rest of the gas in the tidal arm ($>$ 0.5~$\left(\frac{\alpha_{\rm{CO}}}{0.8}\right)$~Gyr). According to the RADEX modeling, while the ranges of $T_{\rm{kin}}$ and $n_{\rm{H_2}}$ are not confined well, the best fitting values (35~K, 10$^{2.5}$~cm$^{-3}$) are slightly higher than those in the middle of the tidal arm, R36a and R37a (25 -- 30~K, 10$^{2.0}$ -- 10$^{2.2}$~cm$^{-3}$). We suggest that R38 is a forming tidal dwarf galaxy at the edge of the tidal arm of VV~114. Future high sensitivity optical and high resolution HI observations will allow us to constrain the star formation and the atomic gas properties of R38.
\section{CONCLUSION}\label{conclusion}
We investigate the physical conditions of the molecular gas in the mid-stage merger VV~114. We present high-resolution observations of molecular gas and dust continuum emission in this galaxy using ALMA band 3 and band 7. This study includes the first detection of extranuclear CH$_3$OH~(2--1) emission in interacting galaxies. The results can be summarized as follows:
\begin{enumerate}
\item We find that the CO~(1--0) and CO~(3--2) lines show significantly extended structures (i.e., the northern and southern tidal arms), the central filament across the galaxy disks, and double-peaks in the overlap region, while the $^{13}$CO~(1--0) line is only detected at the central filament. The filament is also identified by the strong CN~(1$_{3/2}$ -- 0$_{1/2}$), HCO$^+$~(4--3), 110~GHz, and 340~GHz continuum emission.
\item Higher $R_{\rm{3-2/1-0}}$ (0.5 -- 0.8) and $R_{\rm{12/13}}$ (20 -- 50) are detected at the central filament. These higher ratios indicate that the central filament has highly excited (but not thermalized) molecular ISM, and the eastern nucleus is nearly thermalized when it is observed with a 800~pc beam.
\item The unresolved eastern nucleus has the highest $R_{\rm{HCN/HCO^+}}$ (1.34 $\pm$ 0.09), while the dense gas clumps near the eastern nucleus have significantly lower values ($\sim$ 0.5). The broad HCN~(4--3) and HCO$^+$~(4--3) ($\sim$ 290~km~s$^{-1}$) emission lines seen in the unresolved eastern nucleus suggests an obscured AGN (see also \citetalias{ion13}).
\item Radiative transfer analysis of the CO~(1--0), CO~(3--2), and $^{13}$CO~(1--0) emission enables us to map physical parameters of the ``diffuse" gas of an interacting LIRG with 800~pc scale for the first time. The analysis suggests that ``diffuse" gas clouds in the filament have warmer/denser conditions than those in the galaxy disks. This is consistent with predictions from merger simulations. Our analysis also suggest that the [CO]/[$^{13}$CO] is enhanced in the central filament. The extremely high [CO]/[$^{13}$CO] values are more important than the moderately optically thick $^{12}$CO~(1--0) emission to explain the high $R_{12/13}$ in VV~114.
\item Radiative transfer analysis of the HCN~(4--3), HCO$^+$~(4--3), and $^{12}$CO~(3--2) allow us to compare the dense gas clouds around AGN, starburst activities, and the overlap region. These results show that dense gas clouds around AGN have $n_{\rm{H_2}}$ = 10$^{5.0}$ -- 10$^{5.4}$~cm$^{-3}$ and $T_{\rm{kin}}$ $>$ 100~K with [HCN]/[HCO$^+$] $>$ 5, while gas clumps around starburst activities show $n_{\rm{H_2}}$ = 10$^{5.6}$ -- 10$^{5.9}$~cm$^{-3}$ and $T_{\rm{kin}}$ = 40 --100~K with [HCN]/[HCO$^+$] $<$ 4. In addition, the analysis shows that the overlap region has $n_{\rm{H_2}}$ = 10$^{5.0}$ -- 10$^{5.6}$~cm$^{-3}$ and $T_{\rm{kin}}$ = 5 -- 90~K with [HCN]/[HCO$^+$] = 1 -- 9.
\item The spatially resolved Kennicutt-Schmidt law in VV~114 clearly connects the ``starburst" sequence with the ``normal disk" sequence. Most of the data points near the ``starburst" sequence are found in the nuclei and the overlap region, whereas the data points near the ``normal disk" sequence are found in the tidal arms. We also find the $R_{\rm{3-2/1-0}}$ and $R_{\rm{12/13}}$ are well correlated with the $\Sigma_{\rm{SFR}}$.
\item The $M_{\rm{H_2}}$/$M_{\rm{dust}}$ of (128 $\pm$ 16)~$\left(\frac{\alpha_{\rm{CO}}}{0.8}\right)\left(\frac{0.9}{\kappa_{340}}\right)$ in the eastern nucleus of VV~114 is comparable to the Galactic value, but it is a factor of two higher than that in the overlap region of (339 $\pm$ 60)~$\left(\frac{\alpha_{\rm{CO}}}{0.8}\right)\left(\frac{0.9}{\kappa_{340}}\right)$. Since the 340~GHz emission is spatially correlated with dense gas tracers, the cold dust in VV~114 appears to be closely related to the dense molecular component in the filament. The lowest $M_{\rm{H_2}}$/$M_{\rm{dust}}$ in the eastern nucleus may be due to the dusty starburst.
\item Comparing the CS, CN, and CH$_3$OH emission with other galaxies, we suggest that the overlap region is dominated by large-scale shocks similar to the nucleus of NGC~253. From the abundance analysis and distribution of the line ratios, we postulate that the HCN-rich AGN, the HCO$^+$-rich starbursts, and the CH$_3$OH-rich overlap region are important drivers of the molecular chemistry of VV~114.
\item We find a region with relatively high excitation ($\simeq$ 35~K, $\simeq$ 10$^{2.5}$~cm$^{-3}$) and star formation (SFR = 0.10 $\pm$ 0.05~M$_{\odot}$~yr$^{-1}$) at the edge of the southern tail. This region has a shorter $\tau_{\rm{gas}}$ of (0.40 $\pm$ 0.22)~$\left(\frac{\alpha_{\rm{CO}}}{0.8}\right)$~Gyr than the rest of the southern tail ($>$ 1.35~$\left(\frac{\alpha_{\rm{CO}}}{0.8}\right)$~Gyr), and we suggest that it is a forming tidal dwarf galaxy.
\end{enumerate}
\acknowledgements
The authors thanks the anonymous referee for comments that improved the contents of this paper. TS thanks for Yoichi Tamura, Takuma Izumi, and Akio Taniguchi's help on the RADEX calculation. We used a script developed by Y. Tamura for this calculation (http://www.ioa.s.u-tokyo.ac.jp/\verb|~|ytamura/Wiki/?Science\%2FUsingRADEX). TS and other authors thank ALMA staff for their kind support. TS, J. Ueda, and K. Tateuchi are financially supported by a Research Fellowship from the Japan Society for the Promotion of Science for Young Scientists. D. Iono was supported by the ALMA Japan Research Grant of NAOJ Chile Observaory, NAOJ-ALMA-0011 and JSPS KAKENHI Grant Number 2580016. This paper makes use of the following ALMA data: ADS/JAO.ALMA\#2011.0.00467.S. ALMA is a partnership of ESO (representing its member states), NSF (USA) and NINS (Japan), together with NRC (Canada) and NCS and ASIAA (Taiwan), in cooperation with the Republic of Chile. The Joint ALMA Observatory is operated by ESO, AUI/NRAO, and NAOJ.
\bibliographystyle{yahapj}
|
\section{Introduction}
In a seminal paper \cite{j}, V. Jones observed formulae that relate the HOMFLY polynomial to the Alexander polynomial and the algebraic linking number (exponent sum) for closed 3- and 4-braids \cite[(8.4) and (8.10)] {j}. This leads him to write {\it ``Formulae (8.4) and (8.10) lend some weight to the possibility that the exponent sum in a minimal braid representation is a knot invariant''.}
This question, whether the algebraic linking number yields a topological knot invariant when a knot is represented as a closed braid of the minimal braid index, is later called \emph{Jones' conjecture}. In \cite{k1} K. Kawamuro proposed a generalization of Jones' conjecture which we call the \emph{Jones-Kawamuro conjecture}:
if two closed braids $\widehat{\alpha}$ and $\widehat{\beta}$ represent the same oriented link $L$, the inequality
\begin{equation}
\label{eqn:JKconj}
|w(\widehat{\alpha})-w(\widehat{\beta})| \leq n(\widehat{\alpha}) + n(\widehat{\beta}) -2 b(L)
\end{equation}
holds. Here $w$ and $n$ denotes the algebraic linking number and the braid index of a closed braid, and $b(L)$ is the minimal braid index of $L$, the minimum number of strands needed to represent $L$ as a closed braid. Recently, the Jones-Kawamuro conjecture (\ref{eqn:JKconj}) was solved affirmatively by Dynnikov-Prasolov \cite{dp} and LaFountain-Menasco \cite{lm}, by different but related methods.
By Bennequin's formula $sl(\widehat{\alpha})= w(\widehat{\alpha})-n(\widehat{\alpha})$ of the self-linking number of a closed braid \cite{be}, the inequality (\ref{eqn:JKconj}) implies
\begin{equation}
\label{eqn:JKconj2}
|sl(\widehat{\alpha})-sl(\widehat{\beta})| \leq 2 (\max \{n(\widehat{\alpha}),n(\widehat{\beta})\} -b(L)).
\end{equation}
Thus, in a point of view of contact geometry, the Jones-Kawamuro conjecture can be understood as an interaction between the self-linking number and the braid indices. In particular, Jones' conjecture states a surprising phenomenon that the self-linking number, the most fundamental \emph{transverse} knot invariant, yields a \emph{topological} knot invariant when it attains the minimal braid index.
In this paper we prove a generalization of the Jones-Kawamuro conjecture for planar open books, under some additional assumptions and conditions. Our main theorem includes the original Jones-Kawamuro conjecture as its special case, and provides an optimal generalization of the Jones-Kawamuro conjecture for general open books and closed braids, in some sense.
To state our main theorem, we first set up notations.
Let $(S,\phi)$ be an open book decomposition of a contact 3-manifold $(M,\xi)=(M_{(S,\phi)},\xi_{(S,\phi)})$ with respect to the Giroux correspondence \cite{gi}, and let $B$ be the binding.
An oriented link $L$ in $M-B$ is a \emph{closed braid} (with respect to $(S,\phi)$), if $L$ is positively transverse to each page. The number of the intersections with $L$ and a page $S$ is denoted by $n(L)$ and called the \emph{braid index} of $L$.
By cutting $M$ along the page $S_{0}$, $L$ gives rise to an element $\alpha$ of $B_{n(L)}(S)$, the $n(L)$-strand braid group of the surface $S$. We say that $L$ is a \emph{closure} of $\alpha$, and denote by $L=\widehat{\alpha}$. Throughout the paper, we will fix a page $S_0$ and always see a closed braid as the closure of a braid.
A closed braid is regarded as a transverse link in the contact 3-manifold $(M,\xi)$. For a null-homologous transverse link $L$ with Seifert surface $\Sigma$, we denote the self-linking number of $L$ with respect to $[\Sigma] \in H_{2}(M,L)$ by $sl(L,[\Sigma])$. To make notation simpler, we will omit to write $[\Sigma]$.
Apparently, the Jones-Kawamuro conjecture, even for original Jones' conjecture, fails for general open books and closed braids. Here is the simplest counter example.
\begin{example}
\label{exam:counter}
Let $(A,T_{A}^{-1})$ be an annulus open book with negative twist monodromy.
As we have seen in \cite[Example 2.20]{ik1-1}, there is a closed 1-braid $\widehat{\alpha}$ which is a transverse push-off of the boundary of an overtwisted disc (which we call a \emph{tranverse overtwisted disc}), so $sl(\widehat{\alpha})=1$.
On the other hand, the meridian of a connected component of the binding is a closed 1-braid $\widehat{\beta}$ with $sl(\widehat{\beta}) = -1$.
(See Example \ref{example:a} for further discussion).
\end{example}
Since this counter example comes from an overtwisted disc,
one may first hope that an open book supporting a \emph{tight} contact structure satisfies the inequality (\ref{eqn:JKconj2}).
However, as the next example due to Baykur, Etnyre, Van Horn-Morris and Kawamuro, shows this is not true, even for an open book decomposition of the standard contact $S^{3}$.
\begin{example}
\label{exam:counterBEHK}
Let $(A,T_{A})$ be an annulus open book with positive twist monodromy, and $\rho \in B_{1}(A) \cong \pi_{1}(A) \cong \mathbb{Z}$ be a generator of the 1-strand group of an annulus that turns $A$ once in counter clockwise direction. The closed 1-braid $\widehat{\rho^{2}}$ is an unknot with $sl(\widehat{\rho^{2}}) = -3$. (See Example \ref{example:movie} for how to see this).
\end{example}
In fact, as we will discuss in Section \ref{sec:counterexamples}, almost all open books have closed braids violating the inequality (\ref{eqn:JKconj2}). Thus, to get a reasonable generalization of the Jones-Kawamuro conjecture, we need to add some assumptions and modify the statement.
The first assumption and modification we adopt is a topological one concerning closed braids. We concentrate our attention for the case that a knot can across only one particular component of the binding. Let us fix a connected component $C$ of the binding $B$, which we call the \emph{distinguished binding component}.
We say two links $L_1$ and $L_2$ in $M_{(S,\phi)}-B$ are \emph {$C$-topologically isotopic} if they are topologically isotopic in $M-(B-C) =(M-B)\cup C$.
We define the \emph{minimal $C$-braid index} of $L$ by
\[ b_{C}(L)= \min\{n(\widehat{\beta})\: | \: \widehat{\beta} \textrm{ is } C\textrm{-topologically isotopic to }L \}.\]
As we will see in Corollary \ref{cor:Markov}, two closed braids are $C$-topologically isotopic if and only if two closed braids are moved to the other by applying a sequence of braid isotopy and (de)stabilizations along the distinguished binding component $C$.
The second and the third assumptions we add concern the property of an open book. We consider the conditions
\begin{description}
\item[Planar] The page $S$ is planar.
\item[FDTC] The fractional Dehn twist coefficient (FDTC) along the distinguished binding $C$ satisfies $|c(\phi,C)| > 1$.
\end{description}
Here it is interesting to compare these two conditions with \cite[Corollary 1.2]{ik4} that states a planar open book $(S,\phi)$ with $c(\phi,C)>1$ for all $C \subset \partial S$ supports a tight contact structure.
Now our generalization of the Jones-Kawamuro conjecture is stated as follows:
\begin{theorem}[Generalization of the Jones-Kawamuro conjecture]
\label{theorem:main}
Let $(S,\phi)$ be an open book satisfying {\bf [Planar]} and {\bf [FDTC]} and $L \subset M_{(S,\phi)} -B$ be a null-homologous oriented link. If two closed braids $\widehat{\alpha}$ and $\widehat{\beta}$ are $C$-topologically isotopic to $L$, then the inequality
\begin{equation}
\label{eqn:themain}
|sl(\widehat{\alpha})-sl(\widehat{\beta})| \leq 2( \max\{ n(\widehat{\alpha}), n(\widehat{\beta}) \} - b_{C}(L))
\end{equation}
holds.
\end{theorem}
\begin{remark}
For the case of the open book $(D^{2},\mathsf{Id})$,
according to a convention $c(\mathsf{Id}_{D^{2}},\partial D^{2})=\infty$ explained in \cite{ik2} we may regard the open book $(D^{2},\mathsf{Id})$ satisfies {\bf [FDTC]}. In this case being $C$($=\partial D^{2}$)-topologically isotopic is equivalent to being topologically isotopic, so Theorem \ref{theorem:main} contains the Jones-Kawamuro conjecture (\ref{eqn:JKconj2}) as its special case.
\end{remark}
Although the assumptions we add seem too restrictive at first glance, as we will see in Section \ref{sec:counterexamples}, Theorem \ref{theorem:main} is optimal in the sense that we cannot drop any assumptions from Theorem \ref{theorem:main}.
We will present examples of closed braids $\widehat{\alpha}$ and $\widehat{\beta}$ in an open book $(S,\phi)$ violating the inequality (\ref{eqn:themain}), satisfying:
\begin{enumerate}
\item[(a)] $S$ is planar, $\widehat{\alpha}$ and $\widehat{\beta}$ are $C$-topologically isotopic, but $|c(\phi,C)|=1$ (Example \ref{example:a}).
\item[(b)] $S$ is planar, $|c(\phi,C)|>1$, and $\widehat{\alpha}$ and $\widehat{\beta}$ are topologically isotopic but are not $C$-topologically isotopic (Example \ref{example:b}).
\item[(c)] $\widehat{\alpha}$ and $\widehat{\beta}$ are $C$-topologically isotopic, and $|c(\phi,C)| >1$, but $S$ is not planar (Example \ref{example:c}).
\end{enumerate}
Our proof is inspired by LaFountain-Menasco's proof of the Jones-Kawamuro conjecture \cite{lm}, based on the braid foliation machinery developed by Birman and Menasco (see \cite{bf} for a basics of braid foliation). Among other things, foliation change and exchange move introduced in \cite{bm4,bm5}, and various observations and techniques developed in proving Markov Theorem Without Stabilization (MTWS) \cite{BM1,BM2} and usual Markov theorem \cite{BM0} play crucial roles. In our proof, we use an open book foliation machinery developed in \cite{ik1-1,ik2,ik3,ik4} which is a generalization of the braid foliation.
In Section \ref{sec:obf}, we review the open book foliation machinery, for pairwise disjoint annuli cobounded by two closed braids. We also summarize various operations on open book foliation which will be used in later.
In Section \ref{sec:top}, we prove that after suitable stabilizations of particular signs, topologically isotopic closed braids always cobound pairwise disjoint, embedded annuli. It should be emphasized that results in Section \ref{sec:top} hold for all open books and closed braids. As a corollary, we prove a slightly stronger version of the Markov theorem for closed braids in general open books in Corollary \ref{cor:Markov}, which is interesting in its own right.
In Section \ref{sec:proof} we prove Theorem \ref{theorem:main}. This is the point where we need to use assumptions {\bf [Planar]} and {\bf [FDTC]}, and the notion of $C$-topologically isotopic plays crucial roles.
In Section \ref{sec:c-circles}, we prove two Lemmas concerning the property of cobounding annuli with c-circles, which are used in the proof of Theorem \ref{theorem:main}. Existence of such cobounding annuli is a new feature of general open book foliation, which did not appear in braid foliation settings.
In Section \ref{sec:counterexamples} we give various counter examples of the Jones-Kawamuro conjecture (\ref{eqn:themain}) for general open books to explain how our result is best-possible in a certain sense. In particular, in Proposition \ref{prop:counter}, we show that counter examples for a naive generalization of the inequality (\ref{eqn:JKconj2}) are quite ubiquitous. This justifies our modification (\ref{eqn:themain}), a notion of $C$-topologically isotopic and the minimal $C$-braid index.
\section{Open book foliation machinery}
\label{sec:obf}
In this section we review open book foliation machinery which will be used in the proof of Theorem \ref{theorem:main}. For details, see \cite{ik1-1,ik2,ik3}.
\subsection{Open book foliation for cobounding annuli}
Let $\widehat{\alpha}$ and $\widehat{\beta}$ be closed braids in $M_{(S,\phi)}$. Let $A$ be pairwise disjoint embedded annuli such that $\partial A = \widehat{\alpha} \cup (-\widehat{\beta})$. We call such $A$ \emph{cobounding annuli between $\widehat{\alpha}$ and $\widehat{\beta}$}, and write $\widehat{\alpha} \sim_{A} \widehat{\beta}$.
In this section we review open book foliation machinery for cobounding annuli. Note that connected components $-\widehat{\beta}$ of $\partial A$ is \emph{negatively} transverse to pages. This gives rise to some new features in open book foliation, which we will briefly discuss.
Let us consider the the singular foliation $\mathcal F_{ob} (A)$ on $A$ which is induced by intersections with pages
\[ \mathcal F_{ob} (A)=\left\{ A \cap S_t \ | \ t \in [0, 1] \right\}. \]
We say that $A$ admits an open book foliation if $\mathcal F_{ob} (A)$ satisfies the following conditions.
\begin{description}
\item[($\mathcal{F}$ i)]
The binding $B$ pierces $A$ transversely in finitely many points.
Moreover, for each $p \in B \cap A$ there exists a disc neighborhood $N_{p} \subset \textrm{Int}(A)$ of $p$ on which the foliation $\mathcal F_{ob} (N_p)$ is radial with the node $p$, see Figure~\ref{fig:sign}-(i). We call $p$ an {\em elliptic} point.
\item[($\mathcal{F}$ ii)]
The leaves of $\mathcal F_{ob} (A)$ are transverse to $\partial A$.
\item[($\mathcal{F}$ iii)]
All but finitely many pages $S_{t}$ intersect $A$ transversely.
Each exceptional page is tangent to $A$ at a single point.
In particular, $\mathcal F_{ob} (A)$ has no saddle-saddle connections.
\item[($\mathcal{F}$ iv)]
All the tangencies of $A$ and fibers are of saddle type, see Figure~\ref{fig:sign}-(ii).
We call them {\em hyperbolic} points.
\end{description}
By isotopy fixing $\partial A$, $A$ can be put so that it admits an open book foliation (see \cite[Theorem 2.5]{ik1-1}).
A leaf of $\mathcal F_{ob} (A)$, a connected component of $A \cap S_t$ is {\it regular} if it does not contain a tangency point and is {\it singular} otherwise. We will often say that a hyperbolic point $h$ \emph{is around an elliptic point $v$}, if $v$ is an end point of the singular leaf that contains $h$.
The regular leaves are classified into the following four types:
\begin{enumerate}
\item[a-arc]: An arc where one of its endpoints lies on $B$ and the other lies on $\partial A$.
\item[b-arc]: An arc whose endpoints both lie on $B$.
\item[s-arc]: An arc whose endpoints both lie on $\partial A$.
\item[c-circle]: A simple closed curve.
\end{enumerate}
By orientation reasons, an a-arc connects a positive elliptic point and a point of $\widehat{\alpha}$, or a negative elliptic point and a point of $\widehat{\beta}$.
Similarly, an s-arc connects a point of $\widehat{\alpha}$ and a point of $\widehat{\beta}$. A $b$-arc may connect different components of the binding.
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb=120 514 507 712,width=120mm]{zu3.pdf}
\caption{(i) Regular leaves of open book foliation. (ii) essential and strongly essential b-arcs}
\label{fig:leaves}
\end{center}
\end{figure}
An elliptic point $p$ is {\em positive} (resp. {\em negative}) if the binding $B$ is positively (resp. negatively) transverse to $A$ at $p$.
The hyperbolic point $q$ is {\em positive} (resp. {\em negative}) if the positive normal direction $\vec{n}_{A}$ of $A$ at $q$ agrees (resp. disagrees) with the direction of the fibration. We denote the sign of a singular point $v$ by ${\tt sgn}(v)$.
See Figure \ref{fig:sign}.
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb= 163 551 442 715]{zu4.pdf}
\caption{Singular points and its signs for open book foliation: If the positive normal direction (illusrated by dotted arrow) of $A$ is opposite, we have a singular point with minus sign.}
\label{fig:sign}
\end{center}
\end{figure}
According to the types of nearby regular leaves, hyperbolic points are classified into nine types: Type $aa$, $ab$, $bb$, $ac$, $bc$, $cc$, $as$, $abs$, and $cs$. In the case of annuli, $ss$-singularity does not occur.
Each hyperbolic point has a canonical neighborhood as depicted in Figure ~\ref{fig:region}, which we call a {\em region}. We denote by ${\tt sgn}(R)$ the sign of the hyperbolic point contained in the region $R$.
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb=166 502 439 708]{zu5.pdf}
\caption{Nine types of regions.}
\label{fig:region}
\end{center}
\end{figure}
If $\mathcal F_{ob} (A)$ contains at least one hyperbolic point, then we can decompose $A$ as a union of regions whose interiors are disjoint \cite[Proposition 3.11]{ik1-1}. We call such a decomposition a \emph{region decomposition}.
In the region decomposition, some boundaries of a region $R$ can be identified. In such case, we say that $R$ is \emph{degenerated} (see Figure \ref{fig:degenerated}). Some degenerated region cannot exist, because around an elliptic point, all leaves must sit on distinct pages by {\bf ($\mathcal{F}$ i)}.
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb=149 577 451 735,width=100mm]{zu16.pdf}
\caption{Degenerated regions: (iii) illustrates a forbidden degenenerated region. To see this is impossible, look at the leaf illustrated in bold-line. }
\label{fig:degenerated}
\end{center}
\end{figure}
A topological property of b-arc plays an important role. We say a b-arc $b \subset S_{t}$ is
\begin{itemize}
\item {\em essential} if $b$ is not boundary-parallel in $S_{t} \setminus(S_{t} \cap \partial A)$.
\item {\em strongly essential} if $b$ is not boundary-parallel in $S_t$.
\item {\em separating} if $b$ separates the page $S_t$ into two components.
\end{itemize}
See Figure \ref{fig:leaves} (ii).
The conditions `{\em boundary parallel in $S_t$}' and `{\em non}-strongly essential' are equivalent. In this paper we prefer to use the former. Also note that a non-separating b-arc is always strongly essential.
Finally we say that an elliptic point $v$ is {\em strongly essential} if every $b$-arc that ends at $v$ is strongly essential.
For an element $\phi$ of the mapping class group of surface with boundary $S$ and a connected component $C$ of $\partial S$, a rational number $c(\phi,C)$, called the \emph{fractional Dehn twist coefficients} (\emph{FDTC}, in short), is defined \cite{hkm1}. This number measures to what extent $\phi$ twists the boundary $C$, and plays an important role in contact geometry.
A key property of strongly essential elliptic point is that one can estimate the FDTC of the monodromy from such an elliptic point.
\begin{lemma}\cite[Lemma 5.1]{ik2}
\label{lemma:FDTC}
Let $v$ be an elliptic point of $\mathcal F_{ob} (A)$ lying on a binding component $C \subset \partial S$. Assume that $v$ is strongly essential and there are no a-arc or s-arc around $v$. Let $p$ (resp. $n$) be the number of positive (resp. negative) hyperbolic points that lies around $v$. Then
\begin{enumerate}
\item If ${\tt sgn}(v)= +1$ then $-n \leq c(\phi,C) \leq p.$
\item If ${\tt sgn}(v)= -1$ then $-p \leq c(\phi,C) \leq n.$
\end{enumerate}
\end{lemma}
We recall the following observation.
\begin{proposition}\cite[Proposition 2.6]{ik1-1}
\label{prop:withoutc}
If cobounding annuli $A$ admit an open book foliation, then by ambient isotopy fixing $\partial A$ we can put $A$ so that $\mathcal F_{ob} (A)$ have no c-circles. Moreover, if the original cobounding annuli $A$ do not intersect with a component $C'$ of the binding $B$, then $\mathcal F_{ob} (A)$ can be chosen so that no elliptic point of $\mathcal F_{ob} (A)$ lie on $C'$.
\end{proposition}
We remark that when we put $A$ so that $\mathcal F_{ob} (A)$ has no c-circles, in exchange, $\mathcal F_{ob} (A)$ may have a lot of boundary-parallel b-arcs.
Finally, we remind the relation between the open book foliation and the self-linking number.
\begin{proposition}\cite[Proposition 3.2]{ik1-1}
\label{prop:slform}
Let $\Sigma$ be a Seifert surface of a closed braid $\widehat{\alpha}$, admitting an open book foliation. Then the self-linking number is given by
\[ sl(\widehat{\alpha},[\Sigma]) = -(e_{+}-e_{-})+(h_{+}-h_{-}),\]
where $e_{\pm}$ and $h_{\pm}$ the number of positive/negative hyperbolic points of the open book foliation of $\Sigma$.
\end{proposition}
\subsection{Movie presentation}
A movie presentation is a method to visualize an open book foliation of a surface $F$. See \cite[Section 2.1.5]{ik1-1} for details.
Let $F$ be an oriented surface embedded in $M_{(S,\phi)}$ so that it admits an open book foliation $\mathcal F_{ob} (F)$.
We identify $\overline{ M_{(S,\phi)}-S_{0} }$ with $S\times[0,1]/\sim_{\partial}$, where $\sim_{\partial}$ is an equivalence relation given by $(x,t) \sim_{\partial} (x,s)$ for $x \in \partial S$ and $s,t \in [0,1]$.
Let $\mathcal{P}: \overline{M_{(S,\phi)}-S_{0}} \cong S\times[0,1]/\sim_{\partial} \rightarrow S$ be the projection given by $\mathcal{P}(x,t)=x$. We use $\mathcal{P}$ to fix the way of identification of the page $S_t$ with abstract surface $S$. In particular, when we draw the slice $(S_{t},S_{t}\cap F)$, we will actually draw $\mathcal{P}(S_{t},S_{t}\cap F)$.
First we review a notion of describing arc for a hyperbolic point. By definition, a hyperbolic point $h$ is a saddle tangency of a singular page $S_{t^{*}}$ and $F$.
Let $N(h) \subset F$ be a saddle-shaped neighborhood of $h$. We put $F$ so that in the interval
$[t^{*}-\varepsilon,t^{*}+\varepsilon]$ for a small $\varepsilon>0$, $F-N(h)$ is just a product. That is, the complement $F-N(h)$ is identified with $(S_{t^{*}} \cap (F-N(h)) \times [t^{*}-\varepsilon,t^{*}+\varepsilon]$.
The embedding of $N(h)$ is understood as follows: For $t \in [t^{*}-\varepsilon,t^{*})$, as $t$ increases two leaves $l_1(t)$ and $l_2(t)$ in $S_t$ approach along a properly embedded arc $\gamma \subset S_t$ joining $l_{1}$ and $l_{2}$, and at $t=t^{*}$ these two leaves collide to form a hyperbolic point. For $t \in (t^{*},t^{*}+\varepsilon]$, the configuration of leaves are changed (See Figure \ref{fig:desarc}). Thus, the saddle $h$ is determined, up to isotopy, by an arc $\gamma \in S_{t^{*}-\varepsilon}$, which illustrates how two leaves $l_1(t)$ and $l_{2}(t)$ collide. We call $\gamma$ the \emph{describing arc} of the hyperbolic point $h$.
The describing arc also determines the sign of $h$: ${\tt sgn}(h)$ is positive (resp. negative) if and only if the positive normals $\vec n_F$ of $F$ pointing out of (resp. into) its describing arc.
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb=176 582 424 729,width=80mm]{zu20.pdf}
\caption{Describing arc of hyperbolic point (for the case ${\tt sgn} = +$). We indicate the positive normal $\vec n_F$ by dotted gray arrows. We will illustrate the describing arc by dotted line.}
\label{fig:desarc}
\end{center}
\end{figure}
Take $0=s_{0}<s_{1}<\cdots < s_{k} =1$ so that $S_{s_{i}}$ is a regular page and that in each interval $(s_{i},s_{i+1})$ there exists exactly one hyperbolic point $h_i$.
The sequence of slices $( S_{s_{i}}, S_{s_{i}} \cap F )$ with a describing arc of the hyperbolic point $h_i$ is called a \emph{movie presentation} of $F$. A movie presentation completely determines how the surface $F$ is embedded in $M_{(S,\phi)}$ and its open book foliation.
For convenience, to make it easier to chase how the surface and the braid move, we often add redundant slices $(S_t,S_t \cap F)$ in the movie presentation.
\begin{example}[Movie for Example \ref{exam:counterBEHK}]
\label{example:movie}
Here we give a movie presentation of the disc $D$ bounding the unknot $\widehat{\rho^{2}}$, in the open book $(A,T_{A})$ in Example \ref{exam:counterBEHK}.
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb= 142 449 452 735,width=100mm]{zu25.pdf}
\caption{Movie presentation of the disc $D$ bounding an unknot $\widehat{\rho^{2}}$ in the open book $(A,T_{A})$ (Example \ref{exam:counterBEHK})}
\label{fig:movieBEHK}
\end{center}
\end{figure}
\begin{enumerate}
\item[(i)] At $t=0$, we have one a-arc and two b-arcs. T positive normal $\vec n_{D}$ of $D$ is indicated by the gray, dotted arrow. As $t$ increases, the a-arc from $v_0$ and the b-arc connecting $v_1$ and $w_1$ forms a hyperbolic point $h_1$, whose describing arc is indicated by the dotted line. By the positive normal $\vec n_{D}$, the sign of $h_{1}$ is negative.
\item[(ii)] After passing the hyperbolic point $h_1$, we get an a-arc from $v_1$ and b-arc connecting $v_0$ and $w_1$. As $t$ increases, we then have a negative hyperbolic point $h_2$, indicated by the dotted line.
\item[(iii)] After passing the hyperbolic point $h_2$, we get an a-arc from $v_2$ and b-arc connecting $v_1$ and $w_2$. As $t$ increases, we then have a negative hyperbolic point $h_3$, indicated by the dotted line.
\item[(iv)] After passing the hyperbolic point $h_3$, we get an a-arc from $v_1$ and b-arc connecting $v_2$ and $w_2$. As $t$ increases, we then have a positive hyperbolic point $h_4$, indicated by the dotted line.
\item[(v)] After passing the hyperbolic point $h_4$, and at $t=1$ we have one a-arc and two b-arcs. During the passage (i) -- (v), the boundary of a-arc winds twice in the annulus $A$. Finally, the slice at $t=1$ is mapped to the first slice (i) by the monodromy $T_{A}$ to give an embedded disc in $M_{(A,T_{A})}$.
\end{enumerate}
See Figure \ref{fig:movieBEHK}. From this movie presentation, we conclude $\mathcal F_{ob} (D)$ is depicted as Figure \ref{fig:movieBEHK} so by Proposition \ref{prop:slform} we confirm that $sl(\widehat{\rho^{2}})=-3$, as asserted in Example \ref{exam:counterBEHK}.
\end{example}
\subsection{Review of operations on open book foliation}
\label{sec:operation}
In \cite{ik3}, we developed operations that modify the open book foliation. Such operations allow us to simplify the open book foliations and to put surfaces and closed braids in better positions.
These operations are realized by certain ambient isotopy which will often change the braid isotopy class and a position of surface dramatically, but when we just look at the open book foliation, they are local in the following sense:
For each operation there is a certain subset $U$ of $A$ such that the operation changes $\mathcal F_{ob} (A)$ and the pattern of a region decomposition inside $U$, but it preserves $\mathcal F_{ob} (A)$ outside of $U$.
Before describing operations on open book foliation, first we make it clear the meaning of stabilizations of closed braids. Let $C$ be a connected component of the binding $B$, and let $\mu_{C}$ be the meridian of $C$. We say a closed braid $\widehat{\alpha}$ is a positive (resp. negative) \emph{stabilization} of a closed braid $\widehat{\beta}$ along $C$, if $\widehat{\alpha}$ is obtained by connecting $\mu_C$ and $\widehat{\beta}$ along a positively (resp. negatively) twisted band.
Here a positively (resp. negatively) twisted band is a rectangle whose open book foliation has unique hyperbolic point with positive (resp. negative) sign. See Figure \ref{fig:stabilization}.
A positive stabilization preserves the transverse link types whereas a negative stabilization does not. If $\widehat{\alpha}$ is a negative stabilization of a closed braid $\widehat{\beta}$, $sl(\widehat{\alpha}) = sl(\widehat{\beta})-2$.
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb=163 632 434 733,width=90mm]{zu18.pdf}
\caption{Stabilization of closed braid}
\label{fig:stabilization}
\end{center}
\end{figure}
Now we summarize operations on open book foliation in somewhat casual way.
In Figure \ref{fig:operation} we illustrate five operations on open book foliations. The reader can understand these figures as a rule of changing the open book foliation, preserving the topological link types (or, the braid isotopy classes, or the transverse knot types) of $\partial A$. For detailed discussions and more precise statements, see \cite{ik3}.\\
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb=114 380 479 739,width=120mm]{zu13.pdf}
\caption{Operations on open book foliations:
(a): b-arc foliation change, (b): interior exchange move, (c):boundary-shrinking exchange move, (d): destabilization (of sign $\varepsilon$), and (e): stabilization (of sign $\varepsilon$)}
\label{fig:operation}
\end{center}
\end{figure}
{\bf (a): b-arc foliation change}\\
The b-arc foliation change is an operation which changes the pattern of a region decomposition, designed to reduce the number of hyperbolic points around certain elliptic points. This operation preserves the braid isotopy class of $\partial A$.\\
Here is a precise setting. Assume that two ab- or bb- tiles $R_{1}$ and $R_{2}$ of the \emph{same} sign are adjacent at exactly one \emph{separating} b-arc $b$. Let $v_{\pm}$ be the positive and negative elliptic points which are the endpoints of $b$.
Then by ambient isotopy preserving the binding, one can change $R_{1} \cup R_{2}$ as a union of two new regions $R'_{1} \cup R'_{2}$ so that the number of hyperbolic points around $v_{\pm}$ decreases by one.\\
{\bf(b): Interior exchange move}\\
An interior exchange move, which was simply called an \emph{exchange move} in \cite{ik3}, is an operation that removes four singular points. This operation may change the braid isotopy class of $\partial A$, but preserves the transverse link types. \\
Assume that there exists an elliptic point $v$ contained in exactly two ab- or bb-tiles $R_{1}$ and $R_{2}$ of the \emph{opposite} signs, and that at least one of the common b-arc boundary $b$ of $R_1$ and $R_{2}$ is \emph{boundary parallel}. Then by ambient isotopy preserving the \emph{transverse link type} of $\partial A$ one can remove two hyperbolic points in $R_{1} \cup R_{2}$ and elliptic points which are the endpoints of $b$.\\
{\bf (c): Boundary-shrinking exchange move}\\
A boundary-shrinking exchange move is similar to the interior exchange move. Like interior exchange move, this operation may change the braid isotopy class of $\partial A$, but it preserves the transverse link type. A critical difference is that for a boundary-shrinking exchange move we do not require the common b-arc to be boundary-parallel. (This is the reason why we distinguish two exchange moves in a context of open book foliation.)\\
Assume that there exists an elliptic point $v$ contained in exactly two ab-tiles $R_{1}$ and $R_{2}$ of the \emph{opposite} signs. Then by ambient isotopy preserving the \emph{transverse link type} of $\partial A$, one can remove two regions $R_{1} \cup R_{2}$.\\
{\bf (d): Destabilization along a degenerated aa- or as- tile}\\
Let $R$ be a degenerated aa- or as-tile of sign $\varepsilon$, and $v$ is the positive elliptic point in $R$ which lies on a component $C$ of the binding $B$.
Then one can apply a destabilization of sign $\varepsilon$ along $C$ to remove the region $R$. In particular, the transverse link type of $\partial A$ is preserved if $\varepsilon = +$. \\
{\bf (e): Stabilization along an ab- or abs- tile }\\
Let $R$ be an ab- or abs- tile $R$ of sign $-\varepsilon$, and $v$ is the negative elliptic point in $R$ which lies on a component $C$ of the binding $B$. Then by applying stabilization of sign $\varepsilon$ along $C$, we can remove the region $R$. In particular, the transverse link type of $\partial A$ is preserved if $\varepsilon = -$. \\
Since a boundary shrinking exchange move is not discussed in \cite{ik3}, we give a concise explanation. The boundary shrinking exchange move is as a composite of the stabilization along an ab-tile {\bf (e)} and the destabilization along a degenerated aa-tile {\bf (d)}, as shown in Figure \ref{fig:bexmove}.
The condition ${\tt sgn} (R_{1}) \neq {\tt sgn}(R_{2})$ guarantees that we are able to choose the signs of stabilizations and destabilizations are positive so the boundary shrinking exchange move preserves the transverse link type.
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb=125 645 471 729, width=120mm]{zu15.pdf}
\caption{Boundary-shrinking exchange move is realized by positive stabilization (i) and positive destabilization (ii).}
\label{fig:bexmove}
\end{center}
\end{figure}
In a 3-dimensional picture, boundary shrinking exchange move can be understood as a move sliding the braid along a part of surface $R_1 \cup R_2$ which forms a ``pocket''. See Figure \ref{fig:bex}.
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb=107 609 495 732, width=130mm]{zu14.pdf}
\caption{Isotopy realizing a boundary shrinking exchange move}
\label{fig:bex}
\end{center}
\end{figure}
In the rest of the paper, we will often simply call an \emph{exchange move} to mean both an interior exchange move and a boundary-shrinking exchange move, if their differences are not important. Also, we say an exchange move is \emph{along $C$} if two elliptic points which will be removed by the move lie on $C$. In such case, the original closed braid and the resulting closed braid after exchange move are $C$-topologically isotopic.
\section{Topologically isotopic closed braids stably cobound annuli}
\label{sec:top}
In this section, we prove a generalization of \cite[Proposition 1.1]{lm}, which asserts that every topologically isotopic closed braids $\widehat{\alpha}$ and $\widehat{\beta}$ in $S^{3}$ cobound embedded annuli, after positively stabilizing $\widehat{\alpha}$ and negatively stabilizing $\widehat{\beta}$.
This results can be generalized to arbitrary open books and closed braids, without any additional assumptions.
\begin{theorem}
\label{theorem:stablecobound}
If two closed braids $\widehat{\alpha}$ and $\widehat{\beta}$ in an open book $(S,\phi)$ are topologically isotopic, then there exist closed braids $\widehat{\alpha_{+}}$ and $\widehat{\beta_{-}}$ which are positive stabilizations of $\widehat{\alpha}$ and negative stabilizations of $\widehat{\beta}$ respectively, such that $\widehat{\alpha_{+}}$ and $\widehat{\beta_{-}}$ cobound pairwise disjoint embedded annuli $A$.
Moreover, if $\widehat{\alpha}$ and $\widehat{\beta}$ are $C$-topologically isotopic for a distinguished binding component $C$, then
all stabilizations are stabilizations along $C$, and the cobounding annuli $A$ can be chosen so that they do not intersect with the rest of the binding components, $B-C$.
\end{theorem}
\begin{proof}
First we take a sequence of closed braids and cobounding annuli
\begin{equation}
\label{eqn:seqproof}
\widehat{\alpha} \sim_{A} \widehat{\alpha_1} \sim_{A_1} \widehat{\alpha_2} \sim_{A_2} \cdots \sim_{A_{k-1}} \widehat{\alpha_k} \sim_{A_k} \widehat{\beta}
\end{equation}
so that the property
\begin{enumerate}
\item[($\ast$)] $\widehat{\alpha}$ intersects the $i$-th cobounding annuli $A_i$ with at most one point for each $i\geq 1$
\end{enumerate}
holds.
Such a sequence of cobounding annuli and closed braids are obtained as follows:
Since $\widehat{\alpha}$ and $\widehat{\beta}$ are topologically isotopic, there exists a sequence of links which may not be closed braids,
\begin{equation}
\label{eqn:1stseq}
\widehat{\alpha} = L_0 \rightarrow L_1 \rightarrow \cdots \rightarrow L_{k-1} \rightarrow L_{k} = \widehat{\beta}
\end{equation}
such that $L_i \cup (-L_{i+1})$ cobound pairwise disjoint embedded annuli $A'_i$ in $M_{(S,\phi)}$.
By subdividing the sequence (\ref{eqn:1stseq}), we may assume that each $A'_{i}$ intersects $\widehat{\alpha}=L_{0}$ with at most one point.
We inductively modify the sequence (\ref{eqn:1stseq}) to produce the desired sequence of closed braids and cobounding annuli. First we put $\widehat{\alpha_0}= \widehat{\alpha}$ and $A''_0 =A'_0$.
Assume that we have obtained a sequence of links, closed braids and cobounding annuli
\[ \widehat{\alpha} = \widehat{\alpha_0} \sim_{A} \widehat{\alpha_1} \sim_{A_1} \cdots \sim_{A_{i-1}} \widehat{\alpha_{i-1}} \rightarrow L_{i} \rightarrow L_{i+1} \rightarrow \cdots \rightarrow L_k\]
so that the property ($\ast$) holds and that $\widehat{\alpha_{i-1}} \cup (-L_{i})$ cobound annuli $A''_{i}$ that intersect $\widehat{\alpha}$ with at most one point.
We apply Alexander's trick to $L_{i}$ to get a closed braid $\widehat{\alpha_{i}}$ as follows. With no loss of generality, we may assume that $L_{i}$ is transverse to pages except finitely many points. Assume that some portion $\gamma$ of $L_{i}$ is negatively transverse to pages.
Then we take a disc $\Delta$ with properties
\begin{enumerate}
\item The boundary $\partial \Delta$ is a closed 1-braid that is decomposed as union of two arcs, $\partial \Delta = (-\gamma) \cup \gamma'$ and $\Delta \cap L_i = \gamma$.
\item The disc $\Delta$ is positively transverse to the binding at one point. Moreover, the intersection $\Delta \cap B$ lies on the distinguished binding component $C$, if necessary.
\item The disc $\Delta$ is disjoint from $\widehat{\alpha} \cup \widehat{\alpha_{i-1}} \cup L_{i+1}$.
\item An interior of the regular neighborhood $N(\gamma)$ of $\gamma$ in the disc $\Delta$ is disjoint from both $A''_{i}$ and $A'_{i+1}$.
\end{enumerate}
Then we replace the link $L_i$ with a new link $(L_i -\gamma)\cup \gamma'$. This removes the negatively transverse portion $\gamma$ of $L_{i}$.
Moreover, by property (3) and (4) above, by attaching $\Delta$ to $A''_i$ or $A'_{i+1}$, we extend the cobounding annuli. Here, if $\Delta$ intersects with the cobounding annuli $A'= A''_i$ or $A'_{i+1}$, we push $A'$ along $\Delta$ to make them disjoint from $\Delta$, as we illustrate in Figure \ref{fig:Atrick} (By (3), we may assume that other types of intersections does not appear). Since $\Delta$ is chosen to be disjoint from $\widehat{\alpha}$, this modification does not produce new intersections with $\widehat{\alpha}$. In particular, the resulting cobounding annuli preserves the property that they intersect $\widehat{\alpha}$ with at most one point.
After applying this operation (which we call Alexander's trick) finitely many times, we modify $L_i$ so that it is a closed braid $\widehat{\alpha_i}$, and obtain the cobounding annuli $A_i$ between $\widehat{\alpha_{i-1}}$ and $\widehat{\alpha_i}$, and the cobounding annuli $A''_{i+1}$ between $\widehat{\alpha_i}$ and $L_{i+1}$ as desired.
Note that if $\widehat{\alpha}$ and $\widehat{\beta}$ are $C$-topologically isotopic, then one can choose the cobounding annuli $A'_i$ so that they are disjoint from $B-C$. Hence we can take all cobounding annuli $A_i$ so that they are disjoint from $B-C$.
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb=153 569 452 708,width=90mm]{zu11.pdf}
\caption{Alexander's trick:}
\label{fig:Atrick}
\end{center}
\end{figure}
Now we use a sequence (\ref{eqn:seqproof}) to prove the theorem.
We show that by shrinking the first cobounding annuli $A_1$ appropriately, we obtain a new sequence of closed braids and cobounding annuli with shorter length
\begin{equation}
\label{eqn:newseq}
\widehat{\alpha_+} \sim_{A^*} \widehat{\alpha_{2-}} \sim_{A^*_2}\widehat{\alpha_3} \sim_{A_3} \cdots \sim_{A_{k-1}} \widehat{\alpha_k} \sim_{A_k} \widehat{\beta}
\end{equation}
where $\widehat{\alpha_+}$ and $\widehat{\alpha_{2-}}$ are the positive and negative stabilizations of $\widehat{\alpha}$ and $\widehat{\alpha_{2}}$ respectively, and the new cobounding annuli $A^{*}, A_2^{*},A_{3},\ldots$ satisfy the property corresponding to ($\ast$),
\begin{enumerate}
\item[($\ast$)] $\widehat{\alpha_{+}}$ intersects the cobounding annuli $A^{*}$, $A_{2}^{*},A_{3},\ldots$ with at most one point.
\end{enumerate}
Once this is done, an induction on the length of the sequence (\ref{eqn:seqproof}) of cobounding annuli proves the theorem.
In the rest of the proof, we give a construction of shorter sequence (\ref{eqn:newseq}). By Proposition \ref{prop:withoutc}, we can put $A_1$ so that it admits an open book foliation without c-circles. We modify and shrink the annuli $A_1$ in the following five steps. See Figure \ref{fig:summary} for an overview of our construction.\\
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb= 128 484 527 713,width=140mm]{zu8.pdf}
\caption{Summary: how to get new sequence of cobounding annuli and closed braids $\widehat{\alpha_+}$ and $\widehat{\alpha_{2-}}$.}
\label{fig:summary}
\end{center}
\end{figure}
{\bf (i): Removing negative elliptic points (stabilizations for $\widehat{\alpha_1}$) }\\
In the first two steps, we do not care about the sign of hyperbolic points.
We stabilize $\widehat{\alpha_1}$ along ab- or abs- tiles of $A_1$ (see Section \ref{sec:operation} {\bf (d)}) to remove all negative elliptic points from $\mathcal F_{ob} (A_1)$. We denote the resulting closed braid by $\widehat{\alpha_1}'$.
The cobounding annuli $A$ yield the cobounding annuli $A'$ between $\widehat{\alpha}$ and $\widehat{\alpha_1}'$.
\\
{\bf (ii) Removing degenerated aa-tiles (destabilizations for $\widehat{\alpha_1}'$)}\\
After the step (i), the region decomposition of $\mathcal F_{ob} (A_1)$ consists of only aa- and as-tiles and $A_1$ is a union of discs $D_1,\ldots,D_m$ and strips foliated by s-arcs. We may assume that the intersection point $A_1 \cap \widehat{\alpha}$ lies on $D_1$.
For each $D_i$, let us consider the tree $G_i$ whose vertices are elliptic points and a point $w$ on $\widehat{\alpha_2}$ which is the end point of a singular leaf, an whose edges are singular leaves connecting vertices.
Except $w$, a vertex of valence one in the tree $G_i$ is nothing but an elliptic point contained in a degenerated aa- or as-tile $R$.
If $R$ does not intersect with $\widehat{\alpha}$, by destabilizing $\widehat{\alpha_1}'$ along $R$ (see Section \ref{sec:operation} {\bf (e)}), we remove $R$ to simplify the disc $D_i$, without affecting $\widehat{\alpha}$.
Since $D_i$ does not intersect with $\widehat{\alpha}$ for $i>1$, by destabilizations we eventually remove $D_i$.
For the disc $D_1$, we destabilize $\widehat{\alpha_1}$ until the unique intersection point $D_1 \cap \widehat{\alpha}$ obstructs.
Let us write the resulting closed braid by $\widehat{\alpha_1}''$. Again, the cobouding annuli $A'$ give the cobounding annuli $A''$ between $\widehat{\alpha}$ and $\widehat{\alpha_1}''$.
\\
{\bf (iii): Re-ordering the sign of hyperbolic points (exchange moves for $\widehat{\alpha_1}''$)}\\
From now on, we carefully look at the sign of hyperbolic points.
After the step (ii), $\mathcal F_{ob} (A_1)$ is a union of a strip foliated by s-arcs and the disc $D_1$. The graph $G_1$ is a linear graph and $D_1$ is a linear string of as- and aa-tiles. We re-order the sign of hyperbolic points in $D_1$ as follows.
Let us consider the situation that there is a positive elliptic point $v$ such that $v$ is contained in two aa-tiles with opposite signs (see Figure \ref{fig:swap}). Let $C$ be the connected component of $B$ on which $v$ lies, and let $N(v) \cong D^{2} \times [-1,1] \subset D^{2}\times C$ be the regular neighborhood of $v$ in $M$. By suitable ambient isotopy, we put two aa-tiles so that their hyperbolic points are contained in $N(v)$. In a ball $N(v)$, we apply the classical exchange move to exchange the over and under strands (this notion make sense, by considering the projection $N(v) \cong D^{2} \times [-1,1] \to D^{2}$). As a consequence, the sign of the two hyperbolic points are swapped.
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb= 182 468 429 705,width=80mm]{zu9.pdf}
\caption{Exchange move to swap the sign of adjacent two aa-tiles}
\label{fig:swap}
\end{center}
\end{figure}
Thus, by applying exchange moves for $\widehat{\alpha}''$, we can arrange the sign of hyperbolic points in $D_1$ so that the negative hyperbolic points are compiled to the end $w$ and the positive hyperbolic points are compiled to the other end. We denote the resulting closed braid by $\widehat{\alpha_1}'''$.
The cobounding annuli $A''$ produce the cobounding annuli $A'''$ between $\widehat{\alpha}$ and $\widehat{\alpha_1}'''$.
\\
{\bf (iv) Removing negative hyperbolic points (negative stabilizations for $\widehat{\alpha_2}$)}\\
After the step (iii), the sign of the as-tile is negative unless the sign of all hyperbolic points in $D_1$ are positive. We negatively stabilize $\widehat{\alpha_2}$ along a negative as-tiles, until the sign of as-tile become positive (see Section \ref{sec:operation} {\bf (d)}). As a consequence, we remove all negative hyperbolic points from $\mathcal F_{ob} (A)$ and we get a new closed braid $\widehat{\alpha_{2-}}$, a negative stabilizations of $\widehat{\alpha_2}$.
The cobounding annuli $A_2$ between $\widehat{\alpha_2}$ and $\widehat{\alpha_3}$ produce the cobouning annuli $A^{*}_2$ between $\widehat{\alpha_{2-}}$ and $\widehat{\alpha_3}$.
In the construction of $A^{*}_2$, new intersection with $\widehat{\alpha}$ is never created, so $\widehat{\alpha}$ intersects the new cobounding annuli $A^{*}_2$ with at most one point.\\
{\bf (v) Removing positive hyperbolic points (positive stabilizations for $\widehat{\alpha}$)}\\
After the step (iv), all hyperbolic points in $D_1$ are positive.
In the last step, we positively destabilize $\widehat{\alpha_1}$ to shrink the rest of the cobounding annuli $A_1$. Since the degenerated aa-tile in $D_1$ intersects $\widehat{\alpha}$, the destabilization along the degenerated aa-tile causes a change of the closed braid $\widehat{\alpha}$. The change of $\widehat{\alpha}$ induced by the destabilization is understood as follows.
Let $v$ be the positive elliptic point in the degenerated aa-tile $R$ and $C$ be the connected point on which $v$ lies. As in the step (iv), let $N(v) \cong D^{2} \times [-1,1] \subset D^{2}\times C$ be the regular neighborhood of $v$ in $M_{(S,\phi)}$. We may assume that $R$ is contained in $N(v)$, and that the cobounding annuli $A_3,\ldots$ does not intersect with $N(v)$, to guarantee that the change of $\widehat{\alpha}$ does not create new intersection points.
As is noted in \cite{lm}, for a link $\widehat{\alpha} \cup \widehat{\alpha_1}'''$, positive destabilization $\widehat{\alpha_1}'''$ induces a move which is called the microflype, the simplest flype move in braid foliation theory. (see \cite[Section 2.3, Section 5.3]{BM1}).
A positive destabilization $\widehat{\alpha_1}'''$ leads to a positive stabilization of $\widehat{\alpha}$. This isotopy of link $\widehat{\alpha} \cup \widehat{\alpha_1}'''$ is supported in $N(v)$.
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb= 179 593 440 707,width=90mm]{zu10.pdf}
\caption{Microflype: a positive destabilization of $\widehat{\alpha_1}'''$ induces a positive stabilization of $\widehat{\alpha}$.}
\label{fig:microflype}
\end{center}
\end{figure}
Therefore after applying microflypes which induces positive destabilizations for $ \widehat{\alpha_1}'''$ and positive stabilizations for $\widehat{\alpha}$, we eventually remove all singular points, so $A_1$ is now foliated by s-arcs.
Let us call the resulting closed braids $\widehat{\alpha_1}''''$ and $\widehat{\alpha_+}'$, respectively. The cobounding annuli $A'''$ give the cobounding annuli $A''''$ between $\widehat{\alpha_+}'$ and $\widehat{\alpha_1}''''$.
Let $N(A_1)$ be a regular neighborhood of $A_1$ in $M_{(S,\phi)}$.
We put each cobounding annuli $A^{*}_2$, $A_{3},\ldots$ so that the intersection with $\widehat{\alpha_+}'$ does not lie in $N(A_1)$.
Since $A_1$ is foliated by s-arcs, there is an ambient isotopy $\Phi_t: M_{(S,\phi)} \rightarrow M_{(S,\phi)}$ such that:
\begin{itemize}
\item $\Phi_t$ preserves each page of open book.
\item $\Phi_0= \textsf{id}$ and $\Phi_1(\widehat{\alpha''''}) = \widehat{\alpha_{2-}}$.
\item $\Phi_t = \textsf{id}$ outside $N(A_1)$.
\end{itemize}
Let $\widehat{\alpha_{+}} = \Phi_1(\widehat{\alpha_{+}}')$ and $A^{*} = \Phi_{1}(A'''')$. Then $A^{*}$ is a cobounding annuli between $\widehat{\alpha_{+}}$ and $\widehat{\alpha_{2-}}$. Moreover, $\widehat{\alpha_{+}}$ can intersect with each cobounding annuli $A^{*}_2$, $A_{3},\ldots$ at most one point.
This completes the construction of new sequence (\ref{eqn:newseq}).
\end{proof}
We noticed that Theorem \ref{theorem:stablecobound} shows the Markov theorem for general open book, in a slightly stronger form than stated in \cite{sk}:
\begin{corollary}[Markov Theorem for general open book]
\label{cor:Markov}
If two closed braids $\widehat{\alpha}$ and $\widehat{\beta}$ are topologically isotopic, then they admit a common stabilization:
namely, there exists a sequence of closed braids
\[ \widehat{\alpha} = \widehat{\alpha_0}\rightarrow \widehat{\alpha_1} \rightarrow \cdots \rightarrow \widehat{\alpha_{k}} \cong \widehat{\beta_l} \leftarrow \cdots \leftarrow \widehat{\beta_1} \leftarrow \widehat{\beta_0}=\widehat{\beta} \]
such that $\widehat{\alpha_{i+1}}$ (resp. $\widehat{\beta_{j+1}}$) is obtained from $\widehat{\alpha_i}$ (resp. $\widehat{\beta_j}$) by a stabilization or a braid isotopy.
Moreover, if $\widehat{\alpha}$ and $\widehat{\beta}$ are $C$-topologically isotopic for some component of the binding $C$, then all stabilizations are chosen to be a stabilization along $C$.
\end{corollary}
\begin{proof}
By Theorem \ref{theorem:stablecobound}, after stabilizations, $\widehat{\alpha}$ and $\widehat{\beta}$ cobound annuli $A$. As we have seen in Step (i) of the proof of Theorem \ref{theorem:stablecobound}, by stabilizing $\widehat{\alpha}$, we may eliminate all negative elliptic points. Dually, by stabilizing $\widehat{\beta}$ we eliminate all positive elliptic points, hence eventually $\mathcal F_{ob} (A)$ consists of s-arcs, so two boundaries of $A$ are braid isotopic.
\end{proof}
We point out how to read the difference of self-linking number from cobounding annuli (Compare with Proposition \ref{prop:slform}).
\begin{proposition}
\label{prop:sldiff}
Assume that two closed braids $\widehat{\alpha}$ and $\widehat{\beta}$ cobound annuli $A$ admitting an open book foliation. Then
\[
sl(\widehat{\alpha}) -sl(\widehat{\beta}) = -(e_{+}-e_{-})+(h_{+}-h_{-}).
\]
Here $e_{\pm},h_{\pm}$ denote the number of positive/negative ellipic and hyperbolic points in open book foliation $\mathcal F_{ob} (A)$.
\end{proposition}
\begin{proof}
By Corollary \ref{cor:Markov}, stabilizations of $\widehat{\alpha}$ and $\widehat{\beta}$ remove all singular points on $A$, and gives rise to a common stabilization, say $\widehat{\gamma}$.
Let $a_{\pm}$ (resp. $b_{\pm}$) be the number of positive and negative stabilizations to get $\widehat{\gamma}$ from $\widehat{\alpha}$ (resp. $\widehat{\beta}$). Since the positive stabilization preserves the self-linking number whereas the negative stabilization decreases the self-linking number by two, $sl(\widehat{\gamma}) = sl(\widehat{\alpha}) - 2 a_{-} = sl(\widehat{\beta})-2b_{-}$ hence $ sl(\widehat{\alpha}) -sl(\widehat{\beta}) = 2(a_{-}-b_{-})$.
On the other hand, one positive (resp. negative) stabilization of $\widehat{\alpha}$ removes one negative elliptic point and one negative (resp. positive) hyperbolic point. Similarly, one positive (resp. negative) stabilization of $\widehat{\beta}$ removes one positive elliptic point and one positive (resp. negative) hyperbolic point.
This implies
\[ e_{-}=a_{+} + a_{-}, \ e_{+}=b_{+} + b_{-} ,\ h_{+}=a_{-} + b_{+}, \ h_{-}=a_{+} + b_{-}, \]
which show
\[ a_{-}-b_{-} = -e_{+}+h_{+} = e_{-}-h_{-}.\]
\end{proof}
\section{Proof of generalization of Jones-Kawamuro conjecture }
\label{sec:proof}
In this section we prove a generalization of Jones-Kawamuro conjecture.
We prove the following theorem.
\begin{theorem}
\label{theorem:commondestab}
Let $\widehat{\alpha}$ and $\widehat{\beta}$ be closed braids in an open book $(S,\phi)$ that cobound pairwise disjoint embedded annuli $A$. Assume that the cobounding annuli $A$ and the open book $(S,\phi)$ satisfies the following three conditions.
\begin{description}
\item[C-Top] All intersections between $A$ and the binding $B$ lie on the distinguished binding component $C$.
\item[Planar] The page $S$ is planar.
\item[FDTC] $|c(\phi,C)| >1$.
\end{description}
Then there exists closed braids $\widehat{\alpha_{0}}$ and $\widehat{\beta_{0}}$ such that
\begin{enumerate}
\item $\widehat{\alpha_0}$ is obtained from $\widehat{\alpha}$ by braid isotopy, exchange moves and destabilizations along $C$.
\item $\widehat{\beta_0}$ is obtained from $\widehat{\beta}$ by braid isotopy, exchange moves and destabilizations along $C$.
\item $n(\widehat{\alpha_0}) = n(\widehat{\beta_0})$ and $sl(\widehat{\alpha_0})= sl(\widehat{\beta_0})$.
\end{enumerate}
\end{theorem}
The assumptions {\bf [C-Top]} and {\bf [Planar]} leads to the following properties of open book foliations of cobounding annuli, which allow us to perform b-arc foliation change freely.
\begin{lemma}
\label{lemma:key}
Under the assumptions of {\bf [C-Top]} and {\bf [Planar]},
\begin{enumerate}
\item All b-arcs of $A$ are separating.
\item If $v$ is an elliptic point such that all leaves that end at $v$ are b-arcs, then around $v$ there must be both positive and negative hyperbolic points \cite[Lemma 7.6]{ik3}.
\end{enumerate}
\end{lemma}
Note that the statement (1) is nothing but a simple fact that if two endpoints of a properly embedded arc $b$ in a planar surface lie on the same component, then $b$ is separating. Also, the assertion (2) essentially follows from (1).
In the proof of Theorem \ref{theorem:commondestab}, we need to treat cobounding annuli with c-circles (see Remark \ref{rem:c-circle}), and we use the following two results, which will be proven in Section \ref{sec:c-circles}.
First, we observe c-circles should be essential in the cobounding annuli $A$.
\begin{lemma}
\label{lemma:essentialc}
Under the assumptions of Theorem \ref{theorem:commondestab}, the cobounding annuli $A$ does not contain a c-circle which is null homotopic in $A$.
\end{lemma}
Second, we observe that for a planar open book, if cobounding annuli with c-circles is the simplest, then Theorem \ref{theorem:commondestab} is true.
\begin{lemma}
\label{lemma:degac}
Let $(S,\phi)$ be a planar open book.
Assume that closed braids $\widehat{\alpha}$ and $\widehat{\beta}$ representing a knot cobound an annulus $A$ consisting of two degenerated ac-annuli (see Figure \ref{fig:twodegac}). Then $n(\widehat{\alpha})=n(\widehat{\beta})=1$ and $sl(\widehat{\alpha})=sl(\widehat{\beta})$.
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb= 196 598 401 708,width=55mm]{zu6.pdf}
\caption{Special case: cobounding annulus consisting of two degenerated ac-annuli}
\label{fig:twodegac}
\end{center}
\end{figure}
\end{lemma}
Using these results, we now prove Theorem \ref{theorem:commondestab}.
\begin{proof}[Proof of Theorem \ref{theorem:commondestab}]
In the following proof, we prove the theorem for the case $\widehat{\alpha}$ and $\widehat{\beta}$ are knots, namely, $A$ is an annulus, since the general link case follows from the component-wise argument.
Let us put $A$ so that it admits an open book foliation. We prove theorem by induction on the number of singular points in $A$. If $A$ contains no singular points, then $\widehat{\alpha}$ and $\widehat{\beta}$ are braid isotopic so the result is trivial.
First assume that $A$ contains c-circles. By Lemma \ref{lemma:essentialc}, c-circles are homotopic to the core of $A$. In particular, $A$ has no cc-pants or cs-annuli, and an ac- or bc- annulus always appears in pairs sharing their c-circle boundaries.
Then, in a neighborhood of a c-circle, there is a sub-annulus $A' \subset A$ which consists of two degenerated ac-annuli (see Figure \ref{fig:thinannuli}).
Let $\partial A' = \widehat{\alpha}' \cup (-\widehat{\beta}')$. Since c-circle is homotopic to the core of $A$, the sub-annulus $A'$ splits the annulus $A$ into three cobounding annuli $A= A_{\alpha} \cup A' \cup A_{\beta}$, with $\partial A_{\alpha} = \widehat{\alpha} \cup (-\widehat{\alpha}')$, $\partial A_{\beta} = \widehat{\beta}' \cup (-\widehat{\beta})$. By Lemma \ref{lemma:degac}, $n(\widehat{\alpha}')= n(\widehat{\beta}')=1$ and $sl(\widehat{\alpha}')= sl(\widehat{\beta}')$. In particular, $\widehat{\alpha}'$ and $\widehat{\beta}'$ never admits destabilizations.
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb= 196 639 384 728,width=60mm]{zu17.pdf}
\caption{If $\mathcal F_{ob} (A)$ contains c-circles, we may find a sub-annulus $A'$ consisting of two degenerated ac-tiles.}
\label{fig:thinannuli}
\end{center}
\end{figure}
Since $A_{\alpha}$ and $A_{\beta}$ are sub-annuli of $A$, the number of singular points of $\mathcal F_{ob} (A_{\alpha})$ and $\mathcal F_{ob} (A_{\beta})$ is strictly smaller than that of $\mathcal F_{ob} (A)$.
Therefore by induction, there exists a closed braid $\widehat{\alpha_{0}}$ (resp. $\widehat{\beta_{0}}$) which is obtained from $\widehat{\alpha}$ (resp. $\widehat{\beta}$) by braid isotopy, destabilizations and exchange moves along $C$, such that
\[ n(\widehat{\alpha_{0}}) = n(\widehat{\alpha}') = 1 = n(\widehat{\beta}') = n(\widehat{\beta_{0}}), \ \ sl(\widehat{\alpha_{0}}) = sl(\widehat{\alpha}')= sl (\widehat{\beta}') = sl(\widehat{\beta_{0}}). \]
This completes the proof for the case $A$ contains c-circles.
Therefore we will always assume that $\mathcal F_{ob} (A)$ has no c-circles. Then the region decomposition of $A$ only consists of regions which are homeomorphic to 2-cells. By collapsing the boundaries $\widehat{\alpha}$ and $\widehat{\beta}$ to a point $v_{\alpha}$ and $v_{\beta}$ respectively, we get a sphere $\mathcal{S}$, and the region decomposition of $A$ induces a cellular decomposition of $\mathcal{S}$: the 0-cell (vertices) are elliptic points and $v_{\alpha}$ and $v_{\beta}$, and the 1-cells are either a-arc, b-arc or s-arc that are the boundaries of regions, and the 2-cells are aa-, ab-, as-, abs-, or bb-tiles.
Let $V,E$ and $R$ be the number of 0-, 1- and 2-cells. We say an elliptic point $v$ is of \emph{type $(a,b)$} if, in the cellular decomposition, $v$ is the boundary of $a$ 1-cells which are a-arcs and $b$ 1-cells which are b-arcs. Let $V(a,b)$ be the number of elliptic points of $\mathcal F_{ob} (A)$ which are of type $(a,b)$. Then
\begin{equation}
\label{eqn:eul0}
V= 2 + \sum_{v=1}^{\infty}\sum_{a=0}^{v} V(a,v-a).
\end{equation}
where the first $2$ comes from 0-cells $v_{\alpha}$ and $v_{\beta}$.
In the cellular decomposition, every 2-cell has four 1-cells as its boundary, and every 1-cell is adjacent to exactly two 2-cells, so $2R=E$ holds. Thus, by combining the equality $V-E+R = \chi(\mathcal{S}) = 2$ we get
\begin{equation}
\label{eqn:eul1}
4 = 2V - E.
\end{equation}
Next let $E_a$, $E_b$ and $E_s$ be the number of 1-cells which are a-arcs, b-arcs, and s-arcs, respectively. Each a-arc has exactly one elliptic point as its boundary, and each b-arc has exactly two elliptic points as its boundary, so
\begin{equation}
\label{eqn:eul2}
E_a = \sum_{v=1}^{\infty}\sum_{a=0}^{v}a V(a,v-a), \ 2E_b = \sum_{v=1}^{\infty}\sum_{a=0}^{v}(v-a)V(a,v-a).
\end{equation}
Combining (\ref{eqn:eul0}), (\ref{eqn:eul1}), (\ref{eqn:eul2}) altogether, we get
\[ 0= 2 E_{s} + \sum_{v=1}^{\infty}\sum_{a=0}^{v}(v+a-4)V(a,v-a)\]
By rewriting this equality, we get the \emph{Euler characteristic equality}
\begin{equation}
\label{eqn:euler}
2V(1,0) + V(1,1) + 2V(0,2) + V(0,3) = 2E_s + V(2,1) + 2V(3,0) + \sum_{v=4}^{\infty} \sum_{a=0}^{v} (v+a-4)V(a,v-a).
\end{equation}
Assume that the right-hand side of (\ref{eqn:euler}) is non-zero, so one of $V(1,0)$, $V(1,1)$, $V(0,2)$ and $V(0,3)$ is non-zero.\\
{\bf Case (i): $V(1,0) \neq 0$.}\\
An elliptic point of type $(1,0)$ is contained in a degenerated aa-tile. Such an elliptic point is removed by destabilization.\\
{\bf Case (ii): $V(1,1) \neq 0 $.}\\
Let $v$ be an elliptic point of type $(1,1)$, and $\varepsilon$ and $\delta$ be the signs of the hyperbolic points around $v$. If $\varepsilon \neq \delta$, then we can remove $v$ by applying the boundary-shrinking exchange move. If $\varepsilon = \delta$, then we can apply b-arc foliation change to reduce the number of hyperbolic points around $v$. As a result, we get an elliptic point of type $(1,0)$ which can be removed by destabilization, as discussed in {\bf Case (i)}.\\
{\bf Case (iii): $V(0,2) \neq 0 $.}\\
Let $v$ be an elliptic point of type $(0,2)$, and $\varepsilon$ and $\delta$ be the signs of the hyperbolic points around $v$. By Lemma \ref{lemma:key}(2), $\varepsilon \neq \delta$.
Moreover, $v$ cannot be strongly essential, because otherwise by Lemma \ref{lemma:FDTC} we get $-1 \leq c(\phi,C) \leq 1$, which contradicts {\bf [FDTC]}. Hence we can remove such an elliptic point by an interior exchange move.\\
{\bf Case (iv): $V(0,3) \neq 0$. }\\
Let $v$ be an elliptic point of type $(0,3)$. By Lemma \ref{lemma:key}, around $v$ there must be both positive and negative hyperbolic points. Around $v$ there are three hyperbolic points, so we can find two hyperbolic points of the same sign which are adjacent, so the b-arc foliation change can be applied. After the b-arc foliation change, we get an elliptic point of type $(0,2)$ which can be removed as discussed in {\bf Case (iii)}.\\
By {\bf Case (i)--(iv)} above, if the right-hand side of (\ref{eqn:euler}) is non-zero, then we can reduce the number of singular points of $\mathcal F_{ob} (A)$ hence by induction, we find the desired closed braids.
Therefore we now assume that the right hand side of (\ref{eqn:euler}) is zero.
Thus, $V(1,0)=V(1,1)= V(0,2) =V(0,3) = E_s=0$ and all elliptic points are either of type $V(1,2)$ or $V(0,4)$.
Assume that around some elliptic point $v$ of type $(0,4)$, the sign of hyperbolic points are not alternate. This means that there is a situation where the b-arc foliation change can be applied, and applying the b-arc foliation change, $v$ is changed to be of type $(0,3)$, which can be removed by {\bf Case (iv)}.
Next we look at an elliptic point $v$ of type $(1,2)$. Such $v$ lies in one bb-tile $R_{bb}$ and two ab-tiles $R_{ab}^{1}$, $R_{ab}^{2}$.
Assume that ${\tt sgn}(R_{ab}^{1}) \neq {\tt sgn}(R_{ab}^{2})$, or ${\tt sgn}(R_{ab}^{1})= {\tt sgn}(R_{ab}^{2}) = {\tt sgn}(R_{bb})$. Then by b-arc foliation change we get an elliptic point of type $(1,1)$, which can be removed by {\bf Case (ii)}.
Therefore unless the open book foliation $\mathcal F_{ob} (A)$ is in a particular form, a tiling with alternate signs (see Figure \ref{fig:standannuli}(i) -- around each elliptic point of type $(0,4)$ the sign of hyperbolic points are alternate and around each elliptic point of type $(1,2)$, ${\tt sgn}(R_{ab}^{1})= {\tt sgn}(R_{ab}^{2}) \neq {\tt sgn}(R_{bb})$), we can reduce the number of singular point of $\mathcal F_{ob} (A)$.
Thus, we eventually reduced the proof for the case that the cobounding annulus $A$ is tiled with alternate signs.
Let $\varepsilon$ be the sign of ab-tile containing $\widehat{\alpha}$. If $\varepsilon = +$ (resp. $-$), then by negatively (resp. positively) stabilizing $\widehat{\alpha_0}$ we eliminate all negative elliptic points and by negatively (positively) stabilizing $\widehat{\beta}$, we eliminate all positive elliptic points (see Figure \ref{fig:standannuli} (ii)) to get isotopic closed braids.
The observation that $A$ is tiled with alternate signs implies that the number of necessary stabilizations are the same, so $sl(\widehat{\alpha})= sl(\widehat{\beta})$ and $n(\widehat{\beta}) = n(\widehat{\alpha})$.
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb= 85 566 514 740,width=140mm]{zu12.pdf}
\caption{(i) Cobounding annuli $A$ with special open book foliation, a tiling with alternate signs. (ii) The boundaries $\widehat{\alpha}$ and $\widehat{\beta}$ become braid isotopic by performing the stabilization of sign $-\varepsilon$ same times.}
\label{fig:standannuli}
\end{center}
\end{figure}
\end{proof}
\begin{remark}
\label{rem:c-circle}
By Proposition \ref{prop:withoutc}, we may always assume that the first cobounding annulus $A$ has no c-circles. However, when we apply the interior exchange move ({\bf Case (iii)}) one may encounter cobounding annulus with c-circles. Such c-circles cannot be eliminated without increasing the number of singular points. This is a reason why we need to treat open book foliation with c-circles. In the braid foliation case, this problem does not occur since one can always remove c-circles without increasing the number of singular points.
\end{remark}
The Jones-Kawamuro conjecture is a direct consequence of Theorem \ref{theorem:stablecobound} and Theorem \ref{theorem:commondestab}.
\begin{proof}[Proof of Theorem \ref{theorem:main}]
Assume to the contrary that there exist closed braids $\widehat{\alpha}$ and $\widehat{\beta}$ which are $C$-topologically isotopic to the same link $L$ violating the inequality (\ref{eqn:themain}):
\[
|sl(\widehat{\alpha})-sl(\widehat{\beta})| > 2( \max\{ n(\widehat{\alpha}), n(\widehat{\beta})\} - b_{C}(L)). \]
With no loss of generality, we assume that $sl(\widehat{\alpha}) \geq sl(\widehat{\beta})$.
By Theorem \ref{theorem:stablecobound}, there exists closed braids $\widehat{\alpha_{+}}$ and $\widehat{\beta_{-}}$ that cobound annuli $A$, where $\widehat{\alpha_{+}}$ is a positive stabilizations of $\widehat{\alpha}$, and $\widehat{\beta_{-}}$ is a negative stabilizations of $\widehat{\beta}$ along the distinguished binding component $C$. By taking further negative stabilizations of $\widehat{\beta}$ if necessary, we may assume that $n(\widehat{\beta_{-}}) \geq n(\widehat{\alpha_{+}})$.
Since a positive stabilization preserves the self-linking number whereas one negative stabilization decreases the self-linking number by two, we have
\[ sl(\widehat{\alpha_{+}})-sl(\widehat{\beta_{-}}) = sl(\widehat{\alpha}) - sl(\widehat{\beta}) + 2(n(\widehat{\beta_{-}}) - n(\widehat{\beta})). \]
This shows that $\widehat{\alpha_{+}}$ and $\widehat{\beta_{-}}$ also violate the inequality (\ref{eqn:themain}), namely,
\begin{equation}
\label{eqn:contradict}
|sl(\widehat{\alpha_+})-sl(\widehat{\beta_-})| = sl(\widehat{\alpha_+})-sl(\widehat{\beta_-}) > 2( \max\{n(\widehat{\alpha_+}),n(\widehat{\beta_{-}})\}) - b_{C}(K)=2n(\widehat{\beta_-}) -2b_{C}(L).
\end{equation}
Since $\widehat{\alpha}$ and $\widehat{\beta}$ are $C$-topologically isotopic, the cobounding annuli $A$ between $\widehat{\alpha_{+}}$ and $\widehat{\beta_{-}}$ can be chosen so that the assumption {\bf [C-Top]} in Theorem \ref{theorem:commondestab} is satisfied. Hence by Theorem \ref{theorem:commondestab}, there are closed braids $\widehat{\alpha_{0}}$ and $\widehat{\beta_{0}}$ with $n(\widehat{\alpha_0})=n(\widehat{\beta_0})$ and $sl(\widehat{\alpha_0})=sl(\widehat{\beta_0})$, obtained from $\widehat{\alpha_+}$ and $\widehat{\beta_-}$ by destabilizations and exchange moves along $C$.
Since exchange move preserves the self-linking number we have
\[
-2( n(\widehat{\alpha_+}) - n(\widehat{\alpha_0}) ) \leq sl(\widehat{\alpha_+}) -sl(\widehat{\alpha_0}) \leq 0, \ -2( n(\widehat{\beta_+}) - n(\widehat{\beta_0}) ) \leq sl(\widehat{\beta_+}) - sl(\widehat{\beta_0}) \leq 0
\]
Hence
\[ -2( n(\widehat{\alpha_+}) - n(\widehat{\alpha_0}) ) \leq sl(\widehat{\alpha_+}) -sl(\widehat{\beta_-}) \leq 2( n(\widehat{\beta_+}) - n(\widehat{\beta_0})).
\]
This contradicts with (\ref{eqn:contradict}).
\end{proof}
\section{Cobounding annuli with c-circles}
\label{sec:c-circles}
In this section we prove results on cobounding annuli with c-circles used in the previous section.
\begin{proof}[Proof of Lemma \ref{lemma:essentialc}]
The proof is essentially the same as an argument already appeared in \cite[pp. 3016 Case II]{ik3}, the proof of split closed braid theorem for the case that a splitting sphere contains c-circles.
Assume that the cobounding annuli $A$ contain a c-circle which is null-homotopic.
Take an innermost bc-annulus $R$. Here by `innermost' we mean that the c-circle boundary of $R$ bounds a disc $D \subset A$ with $R \subset D$ so that $D-R$ contains no c-circles. Then either $R$ is degenerate bc-annulus (see Figure \ref{fig:degenerated}), or, the region decomposition of $D-R$ consists only of bb-tiles.
We prove the lemma by induction of the number of bb-tiles in $D-R$.
First assume that $D-R$ contains no bb-tiles, namely, $R$ is degenerated. Take a binding component $C$ so that one of the elliptic points in $R$ lies on $C$. Then by \cite[Lemma 7.7]{ik3}, $|c(\phi,C)| \leq 1$ which is a contradiction.
Assume that $D-R$ contains $k>0$ elliptic points, and let $v_{\pm}$ be the elliptic points which lie on $R$. Let us consider the 2-sphere $\mathcal{S}$ obtained by gluing two b-arc boundaries of $D-R$. Then the region decomposition of $D-R$ induces a cellular decomposition of $\mathcal{S}$.
For $i>0$, let $V(i)$ be the number of 0-cells of valence $i$ in the cellular decomposition of $\mathcal{S}$. Then by a similar argument as the equation (\ref{eqn:euler}) in the proof of Theorem \ref{theorem:commondestab}, we have the Euler characteristic equality
\[ 2V(2) + V(3) = 8 + \sum_{i\geq 4}(i-4)V(i). \]
This shows that $\mathcal{S}$ has a 0-cell $v$ (elliptic point) of valence $\leq 3$ which is not $v_{\pm}$.
By applying b-arc foliation change if necessary (thanks to Lemma \ref{lemma:key}, this is always possible) we may assume that $v$ is of valence two (cf. {\bf Case (iv)} in the proof of Theorem \ref{theorem:commondestab}). Then as we have discussed, by interior exchange move we can remove elliptic point $v$. Hence we can reduce the number of elliptic points in $D-R$, so by induction we conclude that a null-homotopic c-circle never exists.
\end{proof}
Next we prove Lemma \ref{lemma:degac}. As we will see in Lemma \ref{lemma:cexam0} and Example \ref{example:c}, Lemma \ref{lemma:degac} does not hold for non-planar open books.
\begin{proof}[Proof of Lemma \ref{lemma:degac}]
Let $A=R_{\alpha} \cup R_{\beta}$, where $R_{\alpha}$ and $R_{\beta}$ are degenerated ac-annuli containing $\widehat{\alpha}$ and $\widehat{\beta}$, respectively, and let $v$ and $w$ be the positive and negative elliptic points in $\mathcal F_{ob} (A)$. (See Figure \ref{fig:twodegac} again).
Since there are no b- and s-arcs in $\mathcal F_{ob} (A)$, $n(\widehat{\alpha})$ and $n(\widehat{\beta})$ are equal to the number of positive and negative elliptic points so $n(\widehat{\alpha})=n(\widehat{\beta})=1$. We look at the movie presentation of $A$ to determine the closed braids $\widehat{\alpha}$ and $\widehat{\beta}$.
We denote the a-arcs in a page $S_t$ whose endpoints are $v$ and $w$ by $a_v=a_v(t)$ and $a_w =a_{w}(t)$, respectively.
Take $S_{0}$ so that the number of c-circles in $S_{0}$ is minimal among all $S_{t}$ $(t\in [0,1])$. Then $S_{0} \cap A$ consists of two a-arcs $a_{w}(0), a_v(0)$ and c-circles $c_1,\ldots,c_k$. We denote the c-circle in $S_t$ that corresponds to $c_i$ by $c_i(t)$, or simply by $c_i$.
Take $t_{\alpha},t_{\beta} \in [0,1]$ so that $S_{t_{\alpha}}$ and $S_{t_{\beta}}$ are the singular pages that contain the hyperbolic point $h_{\alpha}$ in $R_{\alpha}$ and $h_{\beta}$ in $R_{\beta}$, respectively. We treat the case $0<t_{\alpha} < t_{\beta} <1$. The case $0< t_{\beta} < t_{\alpha} < 1$ is similar. With no loss of generality, we may assume that $0 < t_{\alpha} < \frac{1}{2} < t_{\beta} <1$.
Let us look at what will happen as $t$ moves from $0$ to $1$.
Since we have assumed that the number of c-circles in $S_0$ is minimum, the first ac-singular point $h_{\alpha}$ splits the a-arc $a_{v}$ into an a-arc and a new c-circle, say $c_{k+1}$. Similarly, the second ac-singular point in $h_{\beta}$ merges the a-arc $a_w$ and one of c-circles, say $c_{i}$.
Finally, $S_1\cap A$ is identified with $S_0 \cap A$ by the monodromy $\phi:S_1 \rightarrow S_0$.
Recall that every simple closed curve in a planar surface is separating. Take $j \in \{1,\ldots,k\}$ so that $j \neq i$. Since the monodromy $\phi$ preserves $\partial S$, $c_{j}(1)$ is separating implies that $\phi(c_j(1)) = c_{j}(0)$, unless $c_j(1)$ is null-homotopic in $S_1$. However, $\phi(c_j(1)) = c_{j}(0)$ means that a family of curves $c_{j}(t)$ $(t \in [0,1])$ yields an embedded torus, which is absurd. Thus, we conclude we have either
\begin{enumerate}
\item $k=0$, that is, $S_0 \cap A$ consists of two a-arcs $a_{v}$ and $a_{w}$.
\item All the c-circles $c \in S_t$ are null-homotopic in $S_t$.
\end{enumerate}
In the case (1), $A\cap S_{\frac{1}{2}}$ consists of two a-arcs $a_v, a_w$ and the unique c-circle $\mathcal{C}$. The movie presentation of $A$ is described as follows (see Figure \ref{fig:movieacac}).
\begin{enumerate}
\item[(i)] At $t=0$, we have two a-arcs $a_v$ and $a_w$. Here we write the future position of c-circle $\mathcal{C}$ by gray, dotted line.
\item[(ii)] As $t$ approaches to $t_{\alpha}$, the arc $a_v(t)$ deforms to enclose the position of c-circle $\mathcal{C}$, and at $t=t_\alpha$, $a_{v}$ forms a hyperbolic point $h_{\alpha}$. At $t=t_{\alpha} + \varepsilon$ for small $\varepsilon>0$, we have an a-arc $a_v$ and a new c-circle $\mathcal{C}$.
\item[(iii),(iv)] As $t$ approaches to $\frac{1}{2}$, the point $\widehat{\alpha}\cap S_t$ moves along $a_{v}(t)$ to go back to the position at $t=0$. As a consequence, the 1-braid $\alpha$ turns around $\mathcal{C}$ once.
\item[(v)] At $t=\frac{1}{2}$, we have two a-arcs $a_v$ and $a_w$, and a c-circle $\mathcal{C}$, which is a separating simple closed curve in $S_{\frac{1}{2}}$.
\item[(vi)] As $t$ approaches to $t_{\beta}$, the arc $a_w(t)$ deforms to approaches the c-circle $\mathcal{C}$, and at $t=t_\beta$, $a_{w}$ and $\mathcal{C}$ forms a hyperbolic point $h_{\beta}$. At $t=t_{\alpha} + \varepsilon$ for small $\varepsilon>0$, c-circle $\mathcal{C}$ disappears.
\item[(vii,viii)] As $t$ approaches to $1$, the point $\widehat{\beta}\cap S_t$ moves along $a_{w}(t)$ to go back to the position at $t=0$. As a consequence, the 1-braid $\beta$ turns around $\mathcal{C}$ once. Finally, two pages $S_1$ and $S_0$ are identified by the monodromy $\phi$.
\end{enumerate}
Thus in particular, $\mathcal{C}$ is separating implies
\begin{equation}
\label{eqn:critical}
{\tt sgn}(R_{\alpha}) \neq {\tt sgn}(R_{\beta}).
\end{equation}
By Proposition \ref{prop:sldiff} we conclude $sl(\widehat{\alpha}) = sl(\widehat{\beta})$.
In the case (2), a similar argument shows that both $\widehat{\alpha}$ and $\widehat{\beta}$ are closure of the trivial 1-braid so $sl(\widehat{\alpha}) = sl(\widehat{\beta})=-1$.
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb=126 459 493 710,width=130mm]{zu7.pdf}
\caption{Movie presentation of $\mathcal F_{ob} (A)$}
\label{fig:movieacac}
\end{center}
\end{figure}
\end{proof}
As we will see in Example \ref{example:c} in the next section, (\ref{eqn:critical}) does not necessarily hold if a page $S$ is not planar.
\section{Counter examples of Jones-Kawamuro conjecture}
\label{sec:counterexamples}
We close the paper by giving several counter examples that illustrate the necessity of assumptions in Theorem \ref{theorem:main}.
First of all, the following example, coming from our first counterexample Example \ref{exam:counter}, shows that the FDTC assumption {\bf [FDTC]} is necessary and the inequality $>1$ is best-possible: One can not replace the condition $>1$ with $\geq 1$.
\begin{example}[Example \ref{exam:counter}, revisited]
\label{example:a}
Let $A$ be an annulus with boundary $C_1$ and $C_2$, and $T_{A}$ be the right-handed Dehn twist along the core of $A$.
Let us recall the counter example in Example \ref{exam:counter}: A closed 1-braid $\widehat{\alpha}$, the boundary of transverse overtwisted disc, and a closed 1-braid $\widehat{\beta}$, the meridian of $C_1$ in an annulus open book $(A,T_{A}^{-1})$. The binding $\partial A = C_{1} \cup C_2$ forms a negative Hopf link in $S^{3}$. Also note that $c(\phi,C_1) = c(\phi,C_2)=-1$.
As a link in $S^{3}$, $\widehat{\alpha}$ and $\widehat{\beta}$ are depicted in Figure \ref{fig:counterexam1} (i).
Let $\widehat{\alpha}'$ (resp. $\widehat{\beta}'$) be the positive (resp. negative) stabilization of $\widehat{\alpha}$ (resp. $\widehat{\beta}$) along $C_2$. Then $\widehat{\alpha}'$ and $\widehat{\beta}'$ are $C_1$-topologically isotopic (see Figure \ref{fig:counterexam1} (ii)). On the other hand,
\[ |sl(\widehat{\alpha}') - sl(\widehat{\beta}')| = |1-(-3)| = 4 > 2 = 2( \max \{n(\widehat{\alpha}'), n(\widehat{\beta}') \} -b_{C}(K) ) \]
hence they violate the inequality (\ref{eqn:themain}).
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb=168 572 425 736,width=80mm]{zu19.pdf}
\caption{Counter example for Jones-Kawamuro conjecture for planar open book with FDTC=1. (i) Closed braids $\widehat{\alpha}$ and $\widehat{\beta}$ in $S^{3}$, (ii) A negative stabilization of $\widehat{\alpha}$ along $C_2$ is $C_1$-topologically isotopic to $\widehat{\beta}$.}
\label{fig:counterexam1}
\end{center}
\end{figure}
\end{example}
The next example shows that the notion of $C$-topologically isotopic is also necessary.
\begin{example}
\label{example:b}
Let us consider the open book $(A,T_{A}^{2})$, which is an open book decomposition of the unique tight (indeed, Stein fillable) contact structure of $\mathbb{R} P^{3}=L(2,1)$. The FDTCs are $c(T_{A}^{2},C_1) = c(T_{A}^{2},C_2)=2$, so the open book $(A,T_{A}^{2})$ satisfies two assumptions {\bf [Planar]} and {\bf [FDTC]} in Theorem \ref{theorem:main} for both $C_1$ and $C_2$.
Let $\widehat{\alpha} = \partial D$ be a closed braid which is a boundary of a disc $D$, given by the movie presentation in Figure \ref{fig:countermovie}.
From the movie presentation we read that $sl(\widehat{\alpha}) = -5$ and $n(\widehat{\alpha})=2$. On the other hand, let $\widehat{\beta}$ be a closed braid which is a meridian of $C_1$, so $sl(\widehat{\beta})= -1$ and $n(\widehat{\beta})=1$.
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb=128 540 467 734,width=120mm]{zu23.pdf}
\caption{A movie presentation of a disc $D$: The last slice (iv) at $t=1$ is identified with the first slice (i) at $t=0$ by the monodromy $T_{A}^{2}$.}
\label{fig:countermovie}
\end{center}
\end{figure}
Both $\widehat{\alpha}$ and $\widehat{\beta}$ are unknots hence they are topologically isotopic. However,
\[ 4 = |sl(\widehat{\alpha})-sl(\widehat{\beta})| > 2 ( \max\{n(\widehat{\alpha}) ,n(\widehat{\beta})\} -1) = 2 \]
so they violate the inequality (\ref{eqn:themain}).
Note that if $\widehat{\alpha}$ and $\widehat{\beta}$ are $C_{1}$-topologically isotopic, then the links $\widehat{\alpha} \cup C_{2}$ and $\widehat{\beta} \cup C_2$ must be isotopic in $M_{(A,T_{A}^{2})}= \mathbb{R} P^{3}$, hence their linking number must be the same.
However,
\[ 3= lk(\widehat{\alpha},C_1) \neq lk(\widehat{\alpha},C_1) = 1, \ \ -1= lk(\widehat{\alpha},C_2) \neq lk(\widehat{\beta},C_2) = 0.\]
hence $\widehat{\alpha}$ and $\widehat{\beta}$ are neither $C_1$-topologically isotopic nor $C_{2}$-topologically isotopic,
\end{example}
Actually as the next proposition shows, similar counter examples are quite ubiquitous. This shows that in Theorem \ref{theorem:main} the minimal $C$-braid index $b_{C}(K)$ cannot be replaced with the usual minimal braid index $b(K)$, the minimum number of strands needed to represent $K$ as a closed braid in $M_{(S,\phi)}$.
\begin{proposition}
\label{prop:counter}
Let $S$ be a (not necessarily planar) surface with more than one boundary components. For arbitrary open book $(S,\phi)$ with $\phi \neq \textsf{Id}$, there are two closed braids $\widehat{\alpha}$ and $\widehat{\beta}$ in $M_{(S,\phi)}$ which represents the unknot (hence they are topologically isotopic) but they violate the inequality (\ref{eqn:themain}).
\end{proposition}
\begin{proof}
Take two different boundary components $C_1$ and $C_2$ of $S$.
By applying a construction in \cite[Theorem 2.4]{ik1-2}, if $\phi \neq \textsf{Id}$, one gets an embedded disc $D$ admitting open book foliation with the following properties (see Figure \ref{fig:cexamD}).
\begin{enumerate}
\item $\mathcal F_{ob} (D)$ has unique negative elliptic point $v$ which lies on $C_1$ and $n(>1)$ positive elliptic points $w_1,\ldots,w_n$ $(n\geq 2)$ which lie on $C_2$.
\item The region decomposition of $\mathcal F_{ob} (D)$ consists of $n$ ab-tiles of the same sign $\varepsilon$. $\varepsilon = +$ (resp. $\varepsilon = -$) if $\phi$ is not right-veering (resp. right-veering) at $C_1$.
\end{enumerate}
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb=216 628 380 727,width=50mm]{zu24.pdf}
\caption{The disc $D$ and its open book foliation. In the case $\varepsilon = +$, $D$ is a transverse overtwisted disc.}
\label{fig:cexamD}
\end{center}
\end{figure}
Let $\widehat{\alpha}= \partial D$. Then $n(\widehat{\alpha})=n-1$ and $sl(\widehat{\alpha}) = -(n-1) + \varepsilon n$.
In the case $\varepsilon = +$, let $\widehat{\beta}$ be a closed $(n-1)$-braid which is obtained from the meridian of $C_1$ by negatively stabilizing $(n-2)$ times along $C_2$.
Then $sl(\widehat{\beta})= 3-2n$ and $n(\widehat{\beta})= n-1$, hence they violate the inequality (\ref{eqn:themain}),
\[ 2n-2= |sl(\widehat{\alpha})-sl(\widehat{\beta})| > 2 ( \max\{n(\widehat{\alpha}) ,n(\widehat{\beta})\} -1) = 2n-4. \]
In the case $\varepsilon = -$, let $\widehat{\beta}$ be a closed $1$-braid which is a meridian of $C_1$. Then $sl(\widehat{\beta})= -1$ and $n(\widehat{\beta})= 1$, hence they violate the inequality (\ref{eqn:themain}),
\[ 2n-2= |sl(\widehat{\alpha})-sl(\widehat{\beta})| > 2 ( \max\{n(\widehat{\alpha}) ,n(\widehat{\beta})\} -1) = 2n-4. \]
As in Example \ref{example:b}, one can check that $\widehat{\alpha}$ and $\widehat{\beta}$ are not $C$-isotopic for any every boundary component $C$ of $S$, by looking at the linking number with $C_1$ and $C_2$.
\end{proof}
To illustrate the necessity of planarity, we give a counter example of the property (\ref{eqn:critical}) appeared in the proof Lemma \ref{lemma:degac}.
\begin{lemma}
\label{lemma:cexam0}
Let $S$ be non-planar surface. Then for arbitrary monodromy $\phi$, there exist closed 1-braids $\widehat{\alpha}$ and $\widehat{\beta}$ and a cobounding annulus $A$ between them in $M_{(S,\phi)}$, such that
\begin{enumerate}
\item The region decomposition of $A$ consists of two degenerated ac-annuli $R_{\alpha}$ and $R_{\beta}$ (see Figure \ref{fig:twodegac} again).
\item ${\tt sgn}(R_{\alpha}) = {\tt sgn}(R_{\beta})$.
\end{enumerate}
\end{lemma}
\begin{proof}
We give such a cobounding annulus $A$ by movie presentation. Here we give an example ${\tt sgn}(R_{\alpha}) = {\tt sgn}(R_{\beta})=+$. An example of ${\tt sgn}(R_{\alpha}) = {\tt sgn}(R_{\beta})=-$ is obtained similarly. See Figure \ref{fig:movieacac2}. Note that the movie is quite similar to Figure \ref{fig:movieacac}, and the main difference is the slice (v), where the description arc of the hyperbolic point shows $R_{\beta} = +$.
\begin{figure}[htbp]
\begin{center}
\includegraphics*[bb= 116 481 486 730,width=130mm]{zu21.pdf}
\caption{Movie presentation of a cobounding annulus $A$ consisting of two degenerated ac-annuli with the \emph{same} sign (compare with Figure \ref{fig:movieacac}) }
\label{fig:movieacac2}
\end{center}
\end{figure}
\end{proof}
A cobounding annulus $A$ in Lemma \ref{lemma:cexam0} gives a counter example of the Jones-Kawamuro conjecture for non-planar open books.
\begin{example}
\label{example:c}
Let $S$ be the once-hold surface of genus $>0$ and take a monodromy $\phi$ so that $M_{(S,\phi)}$ is an integral homology sphere, with $|c(\phi,\partial S)| >1$.
Let $\widehat{\alpha}$ and $\widehat{\beta}$ be closed 1-braidsgiven by the movie presentation Figure \ref{fig:movieacac2} in Lemma \ref{lemma:cexam0}.
Then $\widehat{\alpha}$ and $\widehat{\beta}$ are $\partial S$-topologically isotopic and null-homologous in $M_{(S,\phi)}$, but by Proposition \ref{prop:sldiff}
\[ sl(\widehat{\alpha})-sl(\widehat{\beta}) = 2 > 2(\max \{n(\widehat{\alpha},n(\widehat{\beta})\} -1) = 0. \]
\end{example}
\section*{Acknowledgements}
The author gratefully thanks Keiko Kawamuro. Without a series of collaborations with her, he cannot write the present paper, and this work is philosophically a continuation of a joint work with her. He also thanks Inanc Baykur, John Etnyre, Jeremy Van Horn-Morris for sharing Example \ref{exam:counterBEHK}.
The author was partially supported by JSPS KAKENHI
Grant Numbers 25887030 and 15K17540.
|
\section{Introduction}
The all-sky survey of the {\it Planck} satellite \citep{Tauber2010} has enabled a new
approach to studying the earliest stages of star formation. The sub-millimetre
measurements, with high sensitivity and an angular resolution down to $\sim4.5\arcmin$,
have enabled detecting and classifying of a large number of cold and compact
Galactic sources. These probably represent different phases in the evolution of
dense interstellar clouds that leads to the formation of stars. Careful analysis of the
{\it Planck} data has led to a list of more than 10000 objects that form the Cold Clump
Catalogue of Planck Objects \citep[C3PO, see][]{planck2011-7.7b}. At {\it Planck}
resolution, it is not possible to resolve gravitationally bound cores even in the
nearest molecular clouds. The low colour temperature of most of the sources ($T<14$\,K)
strongly suggests that the {\it Planck} clumps must have high column densities,
possibly at scales not resolved by {\it Planck}, and they probably contain even
dense cores. A significant fraction of the clumps may be transient structures
produced by turbulent flows, however.
Within the {\it Herschel} Open Time Key Programme {\em Galactic Cold Cores}, we have
carried out dust continuum emission observations of 116 fields that were selected based
on {\it Planck} detections listed in C3PO. The fields, which are typically
$\sim$40$\arcmin$ in size, were mapped with {\it Herschel} PACS and SPIRE instruments
\citep{Pilbratt2010, Poglitsch2010, Griffin2010} at wavelengths 100--500\,$\mu$m. The
higher angular resolution of {\it Herschel} \citep{Poglitsch2010, Griffin2010} enables studying the internal structure of the {\it Planck} clumps, detecting
individual cores, and, in conjunction with mid-infrared data, studying
protostellar sources \citep{Montillaud2014}. The inclusion of far-infrared wavelengths
helps to determine the physical characteristics of the regions and, in particular, to
study the properties of dust emission. First results have been presented in
\citet{planck2011-7.7b, planck2011-7.7a} and in \citet{Juvela2010, Juvela2011, GCC-III}
(papers I, II, and III, respectively). \citet{Montillaud2014} presented the analysis of all
clumps and cores found in our {\it Herschel} fields, including a comparison with the
population of young stellar objects (YSOs). Further studies concentrated especially on
high latitude clouds (\citet{Malinen2014}, Rivera-Ingraham et al. and Ristorcelli et
al., in preparation).
In this paper we concentrate on dust properties and especially on the submillimetre
dust opacity.
Variations of dust emission properties have been investigated with far-infrared (FIR)
and submillimetre observations of diffuse and molecular clouds, using data from IRAS,
COBE, ISOPHOT, the PRONAOS balloon-borne experiment, and ground-based telescopes. It
was shown that the low temperatures found in a sample of molecular clouds
\citep{Laureijs1991, Abergel1994, Abergel1996} and the translucent Polaris Flare cirrus
cloud \citep{Bernard1999} cannot be explained by the extinction of the radiation
field. An increase of the dust emissivity by a factor of 3 compared to the standard
diffuse value was needed to reproduce the cold temperatures observed in the Taurus
filament L1506 \citep{Stepnik2003}. In dense regions, several studies have shown an
opacity increase by a factor between 2 to 4 \citep{Cambresy2001, Kramer2003,
Bianchi2003, delBurgo2003, Kiss2006b, Ridderstad2006, Lehtinen2007}. More recently,
similar results have been obtained with {\it Herschel} and {\it Planck} in molecular
clouds and cold cores \citep{Juvela2011, planck2011-7.13, Martin2012, Roy2013}. In
their detailed modelling of {\it Herschel} observations of the L1506 filament,
\citet{Ysard2013} characterised the dust evolution toward the dense part of the
filament. The dust emissivity in the outer layers of the filament was found to be
consistent with standard grains from the diffuse medium, whereas the emissivity
increases by a factor of $\sim$2 above gas densities of a few times $10^{3} {\rm
cm}^{-3}$. This change has been attributed to the formation of fluffy aggregates in the
dense medium, resulting from grain coagulation, as suggested by previous studies
\citep{Cambresy2001, Stepnik2003, Bernard1999, delBurgo2003, Kiss2006b,
Ridderstad2006}. The average size of aggregates required to fit the FIR,
submillimetre, and extinction profiles in the L1506 filament is about 0.4 $\mu$m
\citep{Ysard2013}. This value is close to the smallest grain size needed to scatter
light efficiently in the mid-IR, which produces the 'coreshine' observed toward a
number of dense cores, which has also been interpreted as a result of grain growth
\citep{Pagani2010, Steinacker2010}.
On the theoretical side, various dust optical property calculations have predicted
a significant increase in the emissivity of aggregates at long wavelengths compared to
compact spherical grains \citep{Wright1987, Bazell1990, Ossenkopf1993, Ossenkopf1994,
Stognienko1995, Kohler2011, Kohler2012}. This variation is shown to be mainly due to
the increase in the porosity fraction with aggregate growth, but the shape, structure,
material composition, and accretion of mantles can also all contribute.
Moreover, \citet{Malinen2011, JuvelaYsard2012, Ysard2012} have investigated the impact
of radiative transfer on the results derived from observations under the assumption of
a single average colour temperature. They showed that the mixing of different
temperatures along the line of sight produces a tendency that is opposite to the
observed one. They concluded that in the absence of internal heating sources, the
observed emissivity increase toward dense clouds cannot be explained by radiative
transfer effects. It must originate in intrinsic variations of the optical
properties of the grains.
It is, however, important to note that the dust emissivity increase is not
systematically observed in the interstellar medium (ISM, see \citet{Nutter2008, Juvela2009, Paradis2009}) .
These intriguing results call for a broader investigation, making use of the large
observations statistics provided by {\it Herschel} and {\it Planck}, probing different
Galactic environments. The key questions are still open today: when, where, and how dust
evolves between diffuse and dense regions, what the physical conditions enhancing
(or preventing) the efficiency of the coagulation process are, what the time scales are,
and whether the process is directly related to specific stages in the cloud or core evolution.
Understanding these questions is critical since knowing the dust
opacity has a direct impact on many key parameters derived from dust emission, such as
the column densities, masses, and volume densities of the clouds. For this reason, it
is also necessary to investigate the possible systematic effects on the emission
and extinction measurements that could cause errors in the opacity estimates.
The structure of the paper is as follows: The observations are described in
Sect.~\ref{sect:obs}. The main results are presented in Sect.~\ref{sect:results},
including the estimates of submillimetre and near-infrared (NIR) optical depths, the
correlations between these variables, and the correlations with environmental factors.
The results are discussed in Sect.~\ref{sect:discussion} before we list the
final conclusions in Sect.~\ref{sect:conclusions}.
\section{Observations and basic data analysis} \label{sect:obs}
\subsection{Target selection}
The selection of the {\it Herschel} fields is described in \citet{GCC-III} and an
overview of all the maps is given in \citet{Montillaud2014}. We only repeat the
main points here.
{\it Planck} sub-millimetre observations, together with IRAS 100\,$\mu$m data, enabled
the detection of over 10000 compact sources in which the dust is significantly colder than
in the surrounding regions \citep{planck2011-7.7b}. The detection procedure is based on
this colour temperature difference and, furthermore, limits the size of the detected
clumps to values below $\sim$12$\arcmin$ \citep{Montier2010}. The sources are believed
to be Galactic cold clumps or, at larger distance, entire clouds
\citep{planck2011-7.7b}.
The fields for {\it Herschel} follow-up observations were selected using a binning of
{\it Planck} cold clumps with respect to the Galactic longitude and latitude, the
estimated dust colour temperature, and the clump mass. At the time of source selection,
distance estimates existed for approximately one third of the sources in C3PO and,
therefore, some sources of unknown mass were also included. The binning ensured
full coverage of the clump parameter space, especially of the high Galactic latitudes
and of the outer Galaxy. Galactic latitudes $|b|<1^{\degr}$ were excluded because that
area is covered by the Hi-GAL programme \citep{Molinari2010}. Similarly, regions
included in other {\it Herschel} key programmes such as the Gould Belt survey
\citep{Andre2010} and HOBYS \citep{Motte2010} were avoided.
A total of 116 separate fields were observed. The SPIRE maps are on average over
40$\arcmin$ in linear size, with an average area of $\sim1800$\,(arcmin)$^2$. The PACS
maps are smaller, with an average area of $\sim 660$\,(arcmin)$^2$. Most fields contain
more than one {\it Planck} clump, the maps altogether covering $\sim$350 individual
{\it Planck} detections. The range of probed column densities extends from diffuse
fields with ${N({\rm H}_2)} \sim 10^{21}$\,cm$^{-2}$ to cores with $N({\rm
H}_2)>10^{23}$\,cm$^{-2}$. The fields are listed in Table~\ref{table:fields} and the
{\it Herschel} observation numbers are included in Table~\ref{table:nicer}.
\subsection{{\it Herschel} data} \label{sect:Herschel_data}
\subsubsection{{\it Herschel} data reduction} \label{sect:reduction}
The fields were mapped with the SPIRE instrument at wavelengths 250, 350, and
500\,$\mu$m and with the PACS instrument at wavelengths 100 and 160\,$\mu$m. One field
was observed with SPIRE alone (G206.33-25.94, part of the Witch Head Nebula, IC~2118). The
{\it Herschel} observations are discussed in detail in \citep{GCC-III} and
\citep{Montillaud2014}. The SPIRE observations at 250\,$\mu$m, 350\,$\mu$m, and
500\,$\mu$m were reduced with the {\it Herschel} Interactive Processing Environment
HIPE v.10.0.0, using the official pipeline with the iterative destriper and the
extended emission calibration options. The maps were produced with the naive map-making
routine. The PACS data at 100\,$\mu$m and 160\,$\mu$m were processed with HIPE v.
10.0.0 up to level 1, and the maps were then produced with Scanamorphos v20
\citep{Roussel2013}. In the order of increasing wavelength, the resolution of the maps
is 7$\arcsec$, 12$\arcsec$, 18$\arcsec$, 25$\arcsec$, and 36$\arcsec$ for the five
bands.
The raw and pipeline-reduced data are available via the {\it Herschel} Science
Archive, the user-reduced maps are available via ESA site\footnote{\em
http://herschel.esac.esa.int/UserReducedData.shtml}.
The accuracy of the absolute calibration of the SPIRE observations is expected to be
better than 7\%\footnote{SPIRE Observer's manual, \\ {\em
http://herschel.esac.esa.int/Documentation.shtml}}. For PACS we assume a calibration
uncertainty of 15\%. This is a conservative estimate that is compatible with the
differences of PACS and Spitzer MIPS measurements of extended emission
\footnote{http://herschel.esac.esa.int/twiki/bin/view/Public/PacsCalibrationWeb}.
\subsubsection{Estimating intensity zero points}
\label{sect:zeropoint}
To determine the zero point of the intensity scale, we compared the {\it Herschel} maps
with {\it Planck} data complemented with the IRIS version of the IRAS 100\,$\mu$m data
\citep{MAMD2005}. The {\it Planck} and IRIS measurements were interpolated to {\it
Herschel} wavelengths using fitted modified blackbody curves, $B_{\nu}(T_{\rm dust})
\nu^{\beta}$, with a fixed value of the spectral index, $\beta=2.0$. The linear
correlations between {\it Herschel} and the reference data were extrapolated to zero
{\it Planck} (+IRIS) surface brightness to determine offsets for the {\it Herschel}
maps. For SPIRE the uncertainties of these fits are typically $\sim$1\,MJy\,sr$^{-1}$
at 250\,$\mu$m and at longer wavelengths smaller in absolute value. The derived
intensity zero points are independent of {\it Planck} calibration and of any
multiplicative errors in the comparison. For PACS the correlations are often less well
defined, and the zero points were set directly based on the comparison of the average
values of the {\it Herschel} maps and the corresponding interpolated {\it Planck} and
IRIS data.
The zero points were calculated iteratively, including colour corrections calculated
using colour temperatures that were estimated from SPIRE data with a fixed spectral
index of $\beta$=2.0.
\subsubsection{Calculating submillimetre optical depth} \label{sect:tau250}
The {\it Herschel} maps were converted into estimates of dust optical depth at
250\,$\mu$m. The surface brightness maps were convolved to a common resolution of
40$\arcsec$ , and colour temperatures were calculated by fitting the spectral energy
distributions (SEDs) with modified blackbody curves with a constant opacity spectral
index of $\beta=$2.0. The 250\,$\mu$m optical depth was obtained from
\begin{equation}
\tau(250\mu{\rm m}) = \frac{I_{\nu}(250\mu{\rm m})}{B_{\nu}(T)},
\end{equation}
using the fitted 250\,$\mu$m intensity $I_{\nu}(250\mu{\rm m})$ and the colour
temperature $T$. The calculations were made with 250--500\,$\mu$m data and
160--500\,$\mu$m data. The fits were weighted according to 15\% and 7\% error estimates
for PACS and SPIRE surface brightness measurements, respectively (see
Sect.~\ref{sect:reduction}).
The assumed value of $\beta$=2.0 may be appropriate for dense clumps, although at lower
column densities the average value is lower, $\beta\sim 1.8$ \citep{Boulanger1996,
planck2011-7.13}, and the value of $\beta$ may further depend on the Galactic location
and the wavelength range \citep[e.g.][]{planck2013-XIV}. If the true value of $\beta$
were 1.8 instead of 2.0, our colour temperature estimates would be higher by $\sim$1\,K
and the $\tau(250\mu$m) values lower by $\sim$30\%. Furthermore, if the values of
$\beta$ were correlated with column density, the slope of $\tau(250\mu$m) vs.
$\tau_{J}$ would be similarly affected.
We return to these effects in Sects.~\ref{sect:biasbias} and \ref{sect:discussion}.
To estimate the statistical uncertainty of $\tau(250\mu$m) values, we used Markov chain
Monte Carlo (MCMC) runs. The prior distribution of temperature values is flat but
limited between 5.0\,K and 35\,K. In addition to the relative errors quoted above, we
included the uncertainty of the intensity zero points. These are typically much smaller
than the assumed relative errors, but may be important at low column densities,
especially at 160\,$\mu$m. The zero-point errors are systematic but are included simply
as another component of statistical noise. Their effect is thus reflected in the error
estimates of individual pixels. The error distribution of $\tau(250\mu$m) is nearly
Gaussian, and we used the standard deviation of the MCMC $\tau(250\mu$m) samples as the
error estimates. These estimates were calculated separately for each pixel of the
$\tau(250\mu$m) maps.
Because of line-of-sight temperature variations, the derived $\tau(250\mu {\rm m})$
estimates probably systematically underestimate the true values \citep{Shetty2009a,
Malinen2011}. We cannot directly determine the magnitude of these errors but, with some
assumptions, we can use radiative transfer modelling to estimate the magnitude of the
bias. The simulations, described in detail in Appendix~\ref{sect:Herschel_simu}, were
used to derive bias maps that are taken into account when the data were correlated with
$\tau_{J}$ values.
\subsection{Near-infrared data} \label{sect:NIR_data}
We used the Two Micron All Sky Survey \cite[2MASS,][]{Skrutskie2006} to derive estimates
of dust column density that are independent of dust emission. We used the method NICER
\citep{Lombardi2001} and the standard extinction curve \citep{Cardelli1989} to convert
the reddening of the background stars to estimates of J-band optical depth, $\tau_{J}$.
Because the calculations involve only near-infrared bands, the results are expected to be
insensitive to the value of the ratio of total to selective extinction, $R_{\rm V}$
\citep{Cardelli1989}.
The shape of the NIR extinction curve is believed to be relatively stable, even at
high extinctions \citep[e.g.][]{Draine2003ARAA, Indebetouw2005, Lombardi2006_Pipe,
RomanZuniga2007, Ascenso2013, Wang2014}. Some variations are observed with Galactic
location and/or density, but generally only at a level of 5\% of the NIR power-law index
\citep[e.g.][]{Stead2009, Fritz2011}.
This question is discussed in more detail in Sect.~\ref{sect:comparison}. The
$\tau(J)$ values are derived using both the J-H and H-K colours but, with the
extinction curve used, we have the correspondence of $E_{J-K}=0.65\,\tau(J)$. Flags
in the 2MASS catalogue were used to avoid galaxies ({\em ext\_key} not null or {\em
gal\_contam} not zero) and sources with uncertain photometry ({\em ph\_qual} worse
than {\em C}).
Five of our fields are fully covered in the VISTA Hemisphere Survey, VHS
\citep{McMahon2013}, which has more than ten times the sensitivity of 2MASS (in $H$ band
VHS has a 5-$\sigma$ detection threshold of 19.0, compared to a 2MASS point source
catalogue completeness limit of $\sim$16\ mag).
One of these fields is too distant to obtain a reliable extinction map, but the data for
the four other fields were analysed and the results compared with those obtained with
2MASS data. The fields are G4.18+35.79 (LDN~134), G21.26+12.11, G24.40+4.68, and
G358.96+36.75. For the fields G21.26+12.11 and G24.40+4.68, only J- and Ks-band data
exist. The data are available in VISTA Survey
Archive\footnote{http://horus.roe.ac.uk/vsa/index.html} , and the VISTA Data Flow System
pipeline processing and science archive are described in \citet{Irwin2004} and
\citet{Hambly2008}.
Extinction maps are produced by averaging extinction estimates of individual stars with
a Gaussian weighting function with FWHM=180$\arcsec$. We also tested a
higher resolution of FWHM=120$\arcsec$. For distant sources, the extinction of the
target clouds cannot be reliably reproduced because of the poor resolution and the
increasing number of foreground stars. This is the main factor that limits the number
of fields where the ratio $\tau(250\mu {\rm m})/\tau(J)$ can be reliably estimated. The
extinction measurements can be significantly biased even in nearby fields if these
contain steep column density variations. No special steps were taken to eliminate
the contamination by foreground stars \citep[see, e.g.,][]{Schneider2011}, apart from
the sigma clipping procedure that is part of the NICER method and was performed at
3$\sigma$ level. The reliability of the extinction maps and the bias caused by sampling
problems and the presence of foreground stars was examined with simulations (see
Sect.~\ref{sect:NIR_simu}). The results of these simulations are used to derive maps of the expected uncertainty and the bias of the ${\tau(J)}$ values for each field.
\subsection{Correlations between sub-millimetre and NIR opacity}
\label{sect:correlations}
The ratio $k$ of sub-millimetre opacity ${\rm \tau(250\mu m)}$ and the NIR opacity
$\tau_{J}$ was estimated for all 116 fields. The $\tau(250\mu$m) maps were convolved to
the 3$\arcmin$ resolution of the $\tau_{J}$ maps. The $\tau(250\mu$m) and the
${\tau(J)}$ data were read at 90$\arcsec$ steps (half-beam sampling), excluding the map
borders where the result of the convolution to 3$\arcmin$ resolution is poorly
defined. For local background subtraction, only areas where the signal was more
than 2$\sigma$ above the average value of the reference area were used
(see~\ref{sect:zeropoint}). Here $\sigma$ is the standard deviation of the values in
the reference region. This is a conservative limit because part of the fluctuations is
caused by real surface brightness variations and not by noise alone.
For $\tau_{J}$ the error estimates were provided by the NICER routine. For ${\rm
\tau(250\mu m)}$ these were obtained from MCMC calculations (see
Sect.~\ref{sect:tau250}). The comparison between the different cases (for example,
regarding the use of 160\,$\mu$m data, background subtraction, or gradient corrections)
provides information on the uncertainty caused by some sources of systematic errors.
The $\tau(250\mu$m) vs. ${\tau(J)}$ points of individual fields and samples of fields
were fitted with a linear model to derive the ratio $\tau(250\mu {\rm m})/\tau(J)$.
These total least-squares fits take into account the uncertainties in both variables.
The fits were made using either all data points or only data below or above a given
$\tau(J)$ limit. To reduce the bias caused by these cuts, the data were divided with
the help of a preliminary linear fit to all data points (see
Sect.~\ref{sect:apparent} for
details). The limiting value of $\tau(J)$ thus corresponds to a position on this line,
and the cut was performed using a line that is perpendicular in a coordinate system
where the average uncertainties of the two variables are equal.
The $\tau(250\mu {\rm m})/\tau(J)$ ratios were also calculated for alternative versions
of the $\tau(250\mu{\rm m})$ data, using local background subtraction or using ancillary
data in an attempt to correct for possible large-scale errors in the surface brightness
data. These alternative data are discussed in Appendix~\ref{sect:alternative}.
\section{Results} \label{sect:results}
\subsection{Apparent ${\rm \tau(250\mu m)}$/$\tau(J)$ values} \label{sect:apparent}
We calculated $\tau(J)$ and ${\rm \tau(250\mu m)}$ maps of the 116 fields as described in
Sects.~\ref{sect:Herschel_data} and~\ref{sect:NIR_data}. The correlations between
$\tau(J)$ and ${\rm \tau(250\mu m)}$ were calculated at a resolution of 180$\arcsec$. In
addition to the full range of column densities, the relationships were examined
separately below and above the limit of $\tau(J)$=0.6 (see
Sect.~\ref{sect:correlations}). This corresponds to visual extinctions $A_{\rm V} \sim
2.3$\,mag and $A_{\rm V} \sim 2.0$\,mag for the $R_{\rm V}$ values of 3.1 and 5.0,
respectively \citep{Cardelli1989}. Instead of a higher limit, we selected the relatively
low number of ${\tau(J)}=0.6$ to maximise the number of fields where a linear fit could
also be made above the $\tau(J)$ threshold. The $\tau(250\mu {\rm m})$ values were
derived from {\it Herschel} data with either 250--500\,$\mu$m or 160--500\,$\mu$m (see
Appendix~\ref{sect:alternative} for analysis with additional alternative data sets).
In a given field, the number of points either below or above the $\tau(J)$ limit is often
insufficient to determine any reliable value for the slope $k={\rm \Delta \tau(250\mu
m)}$/${\Delta \tau(J)}$. In a few fields no reliable value of $k$ can be determined at
all, mainly because of the low quality of the $\tau(J)$ data. This especially affects the
most distant fields because of the contamination by foreground stars and because the
structures are too small to be resolved with the 3$\arcmin$ beam. The formal errors of
the $k$ parameter were used to exclude the clearly unreliable fits. The criterion
$\delta k/k<0.1$ leaves in the default case 106 fits to all data in a field,
103 fits below $\tau(J)$=0.6, and 38 fits above ${\tau(J)}$=0.6. These fits appear
relatively reliable also based on visual inspection.
Figure~\ref{fig:corr_sample} shows an example of the recovered dependence between
$\tau(J)$ and ${\rm \tau(250\mu m)}$ values, including linear fits to the three $\tau(J)$
ranges. In this example, the slope appears to become steeper as $\tau(J)$ increases. This
might be an indication of an increase in the dust submillimetre opacity,
which in turn might be attributed to grain growth
\citep[e.g.,][]{Ossenkopf1994,Stepnik2003,Ormel2011,Ysard2013}. However, before drawing
any such conclusions, we must consider the systematic effects that affect the two
parameters. Figure~\ref{fig:corr_all} shows a summary of all the ${\rm \Delta \tau(250\mu
m)}$/${\Delta \tau(J)}$ values where ${\rm \tau(250\mu m)}$ values are based on {\it
Herschel} 250-500\,$\mu$m data. Before any bias corrections (see below), the values are
seen to cluster around $\sim 2.0 \times 10^{-3}$, with some tendency for higher values in
the higher $\tau(J)$ range.
\begin{figure}
\includegraphics[width=8.8cm]{G95_76+8_17-1_corr_A_PO_3_fwhm180.jpg}
\caption{
Relation between ${\rm \tau(250\mu m)}$ and $\tau(J)$ in the field G95.76+8.17. The black
solid line is a linear weighted total least-squares fit to all data points. The blue and
red points and lines of the corresponding colour show the data and the fits below and
above the threshold of $\tau(J)=$0.6, the dashed line indicates the division. The values
of the slopes $k$ are given in the plot. Error bars are shown for a set of random data
points.
}
\label{fig:corr_sample}
\end{figure}
\begin{figure*}
\includegraphics[width=18.5cm]{corr_all_fwhm180_NB3_PO.jpg}
\caption{
Slopes $k= \Delta \tau(250\mu {\rm m})$/$\Delta \tau(J)$ for all cases with
uncertainties $\delta k/k<0.1$. The black, blue, and red symbols correspond to values
derived for the full $\tau(J)$ range and for data below and above the limit of
$\tau(J)$=0.6. The values of ${\rm \tau(250\mu m)}$ have been derived from SPIRE data
without the subtraction of the local background. Neither $\tau(250\mu$m) nor $\tau(J)$
has been corrected for the expected bias.
}
\label{fig:corr_all}
\end{figure*}
Figure~\ref{fig:hist} summarises the statistics of the dust opacity measurements as
histograms, including all fits where the formal error of the slope of the least-squares
fit ${\rm \tau(250\mu m)}$ vs. $\tau(J)$ is below 10\%.
\begin{figure}
\includegraphics[width=8.5cm]{tauJ_tau250_histo_fwhm180.jpg}
\caption{
Comparison of ${\rm \Delta \tau(250\mu m)}$/${\Delta \tau(J)}$ values in three $\tau(J)$
intervals (three frames), obtained using either three of four {\em Herschel} bands in
deriving the ${\rm \tau(250\mu m)}$ values. The two histograms without hatching
(thick outlines) show all fields where the estimated uncertainty is below 10\%. The two
hatched histograms contain only the intersection with better than 10\% accuracy with both
three and four bands (79, 83, and 14 fields for the three panels, respectively).
}
\label{fig:hist}
\end{figure}
We need to observe a sufficient number of background stars for each resolution
element, even at high column densities. This means that $\tau(J)$ estimates and the
comparison with submillimetre emission can be made only at a low resolution
(2--3$\arcmin$). Averaged over such large areas, the statistical uncertainty of {\it
Herschel} data is very small. In Sect.~\ref{sect:bias_tauJ} we show that the bias is probably also dominated by the errors in $\tau(J)$.
In the following, we rely mainly on the {\em Herschel} data set that consists of
observations 250--500\,$\mu$m (the ``default'' data set).
There are three reasons. First, in theory, the inclusion of the 160\,$\mu$m data reduces
statistical uncertainty of the colour temperature estimates but increases systematic
errors caused by line-of-sight temperature variations \citep{Shetty2009a, Malinen2011}.
Second, because of the smaller (and, for parallel mode, different) area covered by the
PACS observations, the use of the 160\,$\mu$m band significantly reduces the area where
correlations with ${\tau(J)}$ can be calculated.
Third, 160\,$\mu$m data may be affected by additional systematic effects related to the
relative calibration of the two instruments, uncertainties in the zero-point
determination (interpolation between IRAS and {\it Planck} channels and the contribution
of stochastically heated grains in the IRAS 100${\rm \mu m}$ band) and to imperfections
in the map making that could be increased by the smaller size of the PACS maps (see
Sect.~\ref{sect:grad}). We are particularly interested in the coldest regions where
250--500\,${\rm \mu m}$ data provide adequate constraints on the dust temperature.
\subsection{Bias in $\tau(J)$ values} \label{sect:bias_tauJ}
Bias in $\tau(J)$ values is very likely a significant problem, especially for distant fields in which all high column density structures are not resolved and the
results begin to be affected by foreground stars. Both effects decrease the
$\tau(J)$ estimates, especially towards column density peaks. We estimated the
extent of the problem with simulations using the stellar statistics in
low-extinction areas near each field. The contamination by foreground stars was
evaluated with the help of the Besan\c{c}on model of the Galactic stellar
distribution \citep{Robin2003}. We used the {\it Herschel} column density maps as a
model of the column density structure, simulated the distribution of foreground and
background stars, analysed the simulated observations with NICER routine, and
compared the results with the known input $\tau(J)$ map. The procedure is described
in detail in Appendix~\ref{sect:NIR_simu}. We obtained for each field a map of the
expected systematic relative error in $\tau(J)$ that gives a multiplicative
correction factor for the $\tau(J)$. The simulations do not consider the effect of
cloud structures at scales below 18$\arcsec$ but the procedure probably provides a
reasonable estimate of the magnitude of the effect.
We repeated the analysis of the previous section and replaced the original ${\tau(J)}$ maps
with bias-corrected estimates. Figure~\ref{fig:hist_bias} compares the ${\rm \Delta
\tau(250\mu m)}$/${\Delta \tau(J)}$ distributions for the default case with and without
bias correction. The statistics include all fits for which the formal error of the slope is
below 10\%. This corresponds to Fig.~\ref{fig:hist}, but the number of points is
different. Because the bias corrections depend on the cloud distance, fields without
distance estimates had to be dropped. However, in the $\tau(J)>0.6$ interval the number
of fields fulfilling the $\delta k/k<0.1$ criterion has doubled.
\begin{figure}
\includegraphics[width=8.5cm]{tauJ_tau250_histo_deb_fwhm180.jpg}
\caption{
Comparison of ${\rm \Delta \tau(250\mu m)}$/${\Delta \tau(J)}$ distributions without bias
corrections (``Default'') and with bias corrections applied either to $\tau(J)$ or to
both $\tau(J)$ and ${\rm \tau(250\mu m)}$. The three frames correspond to different
ranges of $\tau(J)$ values.
}
\label{fig:hist_bias}
\end{figure}
\begin{figure*}
\begin{center}
\includegraphics[width=17.0cm]{tau_vs_distance_fwhm180.jpg}
\end{center}
\caption{
Slope values $k={\rm \tau(250\mu m)}$/$\tau(J)$ of Fig.~\ref{fig:corr_all} as a
function of estimated distance ({\em frames a, b}), galactocentric distance ({\em frames
c, d}), and galactic height ({\em frames e, f}). The left frames show the original
slopes, the right frames the slopes after the bias corrections of $\tau(250\mu {\rm
m})$ and $\tau(J)$. The black, blue, and red colours correspond to the
full $\tau(J)$ range and to data below and above $\tau(J)$=0.6,
respectively. The solid curves with the
same colours are the weighted moving averages (window sizes 30\% in distance, 800\,pc in
Galactocentric distance, and 100\,pc in Galactic height). All $\tau(250\mu {\rm m})$
values are calculated with SPIRE bands alone.
}
\label{fig:vs_distance}
\end{figure*}
\subsection{Bias in ${\rm \tau(250\mu m)}$ values} \label{sect:bias_tau250}
The systematic errors in ${\rm \tau(250\mu m)}$ values were estimated with radiative
transfer modelling. The line-of-sight temperature variations are expected to be the
main source of error that, in standard analysis, leads to overestimation of the
mass-averaged dust temperature and, subsequently, to underestimation of ${\rm
\tau(250\mu m)}$ \citep[e.g.][]{Ysard2012}.
We constructed for each field a three-dimensional radiative transfer model that covered
a projected area of 30$\arcmin \times 30\arcmin$ with a 10$\arcsec$ pixel size. The
modelling assumed spatially constant dust properties, the dust model \citep{Draine2003}
corresponding to $R_{\rm V}$=5.5 (see Appendix~\ref{sect:Herschel_simu} for details).
The density distribution and external heating were adjusted until the model exactly
reproduced the observed 350\,$\mu$m surface brightness and, for the area above median
column density, the average 250$\mu$m/500\,$\mu$m ratio. The model-predicted surface
brightness maps were analysed as in the case of the actual observations, to produce maps
of ${\rm \tau(250\mu m)}$. To estimate the bias, these values were compared to the
actual ${\rm \tau(250\mu m)}$ values of the model to derive multiplicative
correction factors.
The results depend on the assumed cloud structure in the line-of-sight direction
\citep{Juvela2013_colden}. In our models, the line-of-sight density distribution only has
one peak. This enhances temperature contrasts and increases our bias estimates. On the
other hand, the densest observed cores are probably even more compact, and in their
case we may be underestimating the bias. If the clouds contain embedded sources, the
actual bias may again locally be very different and often lower than predicted by our
models. Although the bias estimation is more difficult than for $\tau(J)$, the
models should again provide a reasonable estimate of the magnitude of the effect. The
relative systematic errors are smaller in ${\rm \tau(250\mu m) }$ than in $\tau(J)$ so
that their effect on the final result is less strong.
The ${k = \rm \Delta \tau(250\mu m)}$/${\Delta \tau(J)}$ values were re-calculated
including bias corrections in both variables. The resulting histograms are included in
Fig.~\ref{fig:hist_bias}. The bias correction makes the distributions significantly
narrower. Figure~\ref{fig:hist_bias} also shows that the corrections are much stronger for
$\tau(J)$ than for $\tau(250\mu{\rm m})$. As a result, the average value of ${\rm \Delta
\tau(250\mu m)}$/${\Delta \tau(J)}$ now decreases with increasing $\tau(J)$. The
number of fields fulfilling the $\delta k/k<0.1$ criterion has doubled to 76 fields (using
SPIRE bands). Compared to the original data, the median values of $k\times 10^4$ have
decreased from 20.2, 21.1, and 23.4 to 15.3, 16.0, and 12.2, the numbers corresponding to the full $\tau(J)$ range, data below $\tau(J)=0.6$, and data above
$\tau(J)=0.6, $ respectively. The strong change in the $k$ values suggests that the uncertainty of $k$ is
probably often several tens of per cent, especially in the ${\tau(J)>0.6}$ interval.
Therefore, Fig.~\ref{fig:hist_bias} does not exclude a systematic
increase of $k$ as the function of $\tau(J)$, if that becomes visible only at high
column densities.
Figure~\ref{fig:vs_distance} displays the slopes $k={\rm \Delta \tau(250\mu m)}$/${\Delta
\tau(J)}$ as function of distance and Galactic location. The left frames show the
relations without and the right frames with the bias corrections applied to $\tau(J)$ and ${\rm \tau(250\mu m)}$. Only fits with $\delta k/k<0.1$ are included. The
original data showed some trends, including an increase in $k$ as a function of
distance and galactocentric distance. The first is visible especially in the high
$\tau(J)$ interval, but was expected because $\tau(J)$ values of distant sources can be
severely underestimated. After bias corrections, the scatter of the $k$ values is
significantly reduced. This suggests that the corrections are of correct magnitude. The
distance dependence has changed so that in the corrected data there is a slight decrease
in $k$ values as a function of distance. This could point to some over-correction of the
$\tau(J)$ estimates, although the bias correction should not only depend on distance, but
even mainly on the cloud structure. However, the decrease of $k$ can be an indication of
selection effects or direct resolution effects. For example, higher $k$ values might be
found in individual dense clumps that are only resolved at short distances.
There is little difference between the $k$ values found in the three $\tau(J)$ intervals.
In the next section we examine in more detail the global $\tau(J)$ dependence of the
$\tau(250\mu {\rm m}) / \tau(J)$ ratios, especially regarding the highest observed column
densities.
\subsection{Global relation ${\rm \tau(250\mu m)}$ vs. $\tau(J)$}
\label{sect:global}
To further test the hypothesis that ${\tau(250\mu{\rm m}) / \tau(J)}$ ratios may change
systematically as a function of column density, we carried out non-linear fits
${\tau(250\mu m) = A + B \times \tau(J) + C \times \tau(J)^2}$. The fits were first
performed using the combined data of all fields. To reduce the mismatch in the zero
levels of individual fields, we subtracted from each $\tau(J)$ and ${\rm \tau(250\mu m)}$
map the local background using the off regions listed in Table~\ref{table:fields}. The
off regions are not completely void of emission but provide a common reference point for
the quantities.
Thus, the relation is expected to develop via the origin for each field separately, the parameter $A$
being close to zero for the combined data as well.
The sign of the fitted parameter $C$ indicates the possible increase or decrease of
$k={\Delta \tau(250\mu {\rm m})/ \Delta \tau(J)}$ as the function of column density.
Figure~\ref{fig:MCMC_TLS} shows the results obtained with the bias-corrected data. We
included data from all fields in which individual linear fits had $\delta k/k<0.2$, thus
relaxing the previous constraint of $\delta k/k<0.1$. The second-order polynomial was
fitted to all data and separately to data points with $\tau(J)>1.0$. In the previous
section a threshold value of $\tau(J)=0.6$ was used. However, some 70\% of all points are
below $\tau(J)=0.6$ and, when included, they dominate the fits that systematically
underestimate the data above $\tau(J)\sim 5$. With the combined data set, there are
enough high column density data points so that the lower limit can be moved upwards. The
use of the $\tau(J)=1.0$ threshold enables an adequate fit to all data with higher
$\tau(J)$ values.
The $\tau(J)$ calculations employ a different off region for each field. These may
contain different amounts of extinction, which leads to small relative shifts along the
$\tau(J)$ axis. Based on dust emission, the extinction in the off regions is typically
$\tau(J)$=0.2--0.4. The uncertainty of the relative zero points contributes to the
scatter in Fig.~\ref{fig:MCMC_TLS}, but the effect is weaker
than the total
dispersion and the non-linearity seen at high extinctions.
To prevent the $\tau(J)$ cut itself from biasing the fits, the data were selected using
lines perpendicular to a linear least-squares line fitted to all data (cf.
Sect.~\ref{sect:correlations}). Thus, the quoted $\tau(J)$ limits correspond to a point
on the fitted line, and the cut itself is perpendicular to the fitted line. All fits
take into account the uncertainties in both variables, which are assumed to be
uncorrelated. The error distributions of the parameters $A$-$C$ were calculated with an
MCMC.
The sign of the parameter $C$ depends on the range of $\tau(J)$ values but is less
dependent on the field selection, for example regarding the $\delta k/k$ limit that was
used to select the fields. Most fields with high $\tau(J)$ values (and thus with a
wide dynamical range) have low values of $\delta k/k$. When all points are included,
the value of the parameter $C$ is negative, but the fit is very poor at high $\tau(J)$.
When pixels $\tau(J)<1.0$ are excluded, $C$ becomes positive (see
Fig.~\ref{fig:MCMC_TLS}a), which points to an increase in the submillimetre opacity above
$\tau(J) \sim 1$. Systematic additive errors in either parameter might also explain the
different behaviour at very low $\tau(J)$. When the fit is made using all data
$\tau(J)>0.6$ (not shown), the parameter $C$ is marginally positive, but beyond
$\tau(J)=10$ the fitted line is below all the data points. The second frame of
Fig.~\ref{fig:MCMC_TLS} shows the fits when pixels with colour temperatures above 14\,K
are excluded. The values of $C$ are now higher and positive even when data
$\tau(J)<1.0$ are not excluded.
The best fit to the high $\tau(J)$ end of the relation is still obtained by excluding data
with $\tau(J)<1.0$, this results in the relation
\begin{equation}
\tau(250\mu {\rm m}) = 0.73\times 10^{-3} + 1.25\times 10^{-3}\,\tau(J) + 0.11\times
10^{-3}\, \tau(J)^2
.\end{equation}
The formal error estimates of the parameters $A$-$C$ are of the order of 5\%, probably
lower than the systematic uncertainties. All data beyond $\tau(J)\sim 5$ are affected
by large bias corrections and, consequently, the value of $C$ also depends on the
accuracy of these corrections. Thus, Fig.~\ref{fig:MCMC_TLS} strongly suggests but does
not yet provide a final proof of the variations of the ratio $\tau(250\mu {\rm m})/\tau(J)$ . The
positive offset $A$=0.7$\times 10^{-3}$ results from the facts that at low column
densities the relation is linear, the curvature increases only beyond ${\tau(J)\sim 5}$,
and the lowest data points ${\tau(J)<1.0}$ are not part of the fit.
\begin{center}
\begin{figure}
\includegraphics[width=8.5cm]{plot_MCMC_TLS_escale.jpg}
\caption{
Fit of ${\tau(250\mu {\rm m}) = A + B \times \tau(J) + C \times \tau(J)^2}$ to the
combined data of all fields in which individual linear fits showed a strong correlation with
$\delta k/k<0.2$. The blue and the green lines correspond to fits to the
full column density range and to data points $\tau(J)>1.0$ alone,
respectively. In the second frame,
only data with colour temperatures below 14\,K are used.
}
\label{fig:MCMC_TLS}
\end{figure}
\end{center}
Figure~\ref{fig:MCMC_TLS_samples} shows the error distributions of the parameters A--C.
The fit was made to data $\tau(J)>1.0$ for all fields with a
linear fit accuracy $\delta
k/k<0.2$. In most fields, the formal error estimates of $\delta \tau(J)$ and $\delta
\tau(250\mu {\rm m})$ are smaller than the actual scatter of points. Therefore we used the residuals of the linear fits before the MCMC calculation to determine a
scaling factor, typically 2.0--3.0, that makes the error estimates in each field
consistent with the actual scatter. Even after this increase of uncertainties, MCMC gives a
100\% probability for a positive value of C. In reality, the result is not that strong
because the uncertainty may be dominated by systematic errors. The sign of $C$ was already
seen to change depending on the range of $\tau(J)$ values fitted. The result also depends
on a relatively small number of fields with data above $\tau(J) > 5$. Therefore, we must
consider the $\tau(250\mu {\rm m})$ vs. ${\tau(J)}$ relation in individual
fields in more detail.
\begin{figure}
\includegraphics[width=8.5cm]{plot_MCMC_TLS_samples_1_000_99_000_escale.jpg}
\caption{
Distributions of the parameters of the fit $\tau(250\mu {\rm m}) = A + B \times \tau(J)
+ C \times \tau(J)^2$. The fit is limited to data with $\tau(J)>1$.
}
\label{fig:MCMC_TLS_samples}
\end{figure}
\begin{center}
\begin{figure*}
\begin{center}
\includegraphics[width=6.0cm]{single_001.jpg}
\includegraphics[width=6.0cm]{single_002.jpg}
\includegraphics[width=6.0cm]{single_003.jpg}
\includegraphics[width=6.0cm]{single_004.jpg}
\includegraphics[width=6.0cm]{single_005.jpg}
\includegraphics[width=6.0cm]{single_006.jpg}
\includegraphics[width=6.0cm]{single_007.jpg}
\includegraphics[width=6.0cm]{single_008.jpg}
\includegraphics[width=6.0cm]{single_009.jpg}
\includegraphics[width=6.0cm]{single_010.jpg}
\includegraphics[width=6.0cm]{single_011.jpg}
\includegraphics[width=6.0cm]{single_012.jpg}
\end{center}
\caption{
Fits of $\tau(250\mu {\rm m})$ vs. $\tau(J)$ in selected fields ordered by increasing
distance. The red and blue points (dust temperature above and below 14\,K) with error
bars are the bias-corrected data points, where $\tau (250\mu {\rm m})$ is based on SPIRE
data.
The slopes of linear fits are listed in the upper left corner for (1) the default data set
based on SPIRE data alone (``Def.''), (2) 160--500\,$\mu$m data (``$\lambda$=4''), (3)
SPIRE data but without bias corrections (``-deb''). The linear fit of the default case
is shown with a black line.
The non-linear fits are shown with solid blue lines (MCMC) and solid red lines
(bootstrapping) with associated shaded 68\,\% confidence regions. The dashed magenta
lines correspond to different bias correction of $\tau(J)$ using distances $d-\delta
d$ and $d+ \delta d$.
The parameters from bootstrapping are given in the lower right corner. The non-linear
fit to data without bias corrections is plotted with a solid green curve (without error
region) with the parameter $C$ given at the bottom of the figure.
The zero points of the ${\tau(J)}$ axes are not absolute.
}
\label{fig:indi_1}
\end{figure*}
\end{center}
\begin{center}
\begin{figure*}
\begin{center}
\includegraphics[width=6.0cm]{single_013.jpg}
\includegraphics[width=6.0cm]{single_014.jpg}
\includegraphics[width=6.0cm]{single_015.jpg}
\includegraphics[width=6.0cm]{single_016.jpg}
\includegraphics[width=6.0cm]{single_017.jpg}
\includegraphics[width=6.0cm]{single_018.jpg}
\includegraphics[width=6.0cm]{single_019.jpg}
\includegraphics[width=6.0cm]{single_020.jpg}
\includegraphics[width=6.0cm]{single_021.jpg}
\end{center}
\caption{
Continued{\ldots}
}
\label{fig:indi_2}
\end{figure*}
\end{center}
\subsection{Correlations in selected fields} \label{sect:individual}
Figure~\ref{fig:MCMC_TLS} showed hints of an increase of the ${\tau(250\mu {\rm m})/\tau(J)}$
values as the column density increases, but the global statistics may be confused by the
mix of different fields. Furthermore, the sign of the $C$ parameter is determined by the
highest ${\tau(J)}$ points that originate in a small number of individual fields. Each
field may be affected by different systematic effects related to the surface brightness
zero points, distance uncertainty (via bias correction), and differences in the local
radiation field. Several diffuse fields are even entirely below $\tau(J)\sim 1.0$.
Therefore, we also need to examine the fields individually.
Three criteria were used to select a subset of fields. We required that (1) the
uncertainty of the fitted parameter $C$ is below ${\rm 0.3\times 10^{-4}}$, (2) there are
at least ten data points (selected from the maps at 90$\arcsec$ steps) with $\tau(J)$
above 0.6, and (3) the bias corrections change the slope of the
linear fit of ${\rm \tau(250\mu
m)}$ vs. $\tau(J)$ by less than 30\%. The first two criteria ensure that there are enough
data points at large $\tau(J)$ with a small scatter to gain some
insight about
the column density dependence of the $\Delta \tau(250\mu {\rm m})/\Delta \tau(J)$. The
third criterion excludes distant fields for which the uncertainty of the bias correction
of $\tau(J)$ renders the results uncertain, even for the apparently well-defined
relation between ${\tau(J)}$ and ${\rm \tau(250\mu m)}$.
The selection leaves 23 fields with distances mostly in the range 100-500\,pc. There are
two exceptions, G216.76-2.58 and G111.41-2.95, for which the estimated distances are
2.4\,kpc and 3.0\,kpc. We kept the two fields in the sample even
though the results are known to be unreliable because of the
large distance.
To avoid underestimating the fit errors, we scaled the error estimates of $\tau(J)$
and $\tau(250\mu {\rm m}) $ up to correspond to the actual scatter of points (see
Sect.~\ref{sect:global}). All observations, sampled at 90$\arcsec$ steps, are fitted
with a linear model ${\tau (250\mu {\rm m})=b + k \times \tau(J)}$ using total least-squares. We continued to use as the default data set one with $\tau(250\mu {\rm m})$
derived from SPIRE bands alone, including bias corrections in $\tau(250\mu {\rm
m})$ and $\tau(J)$. However, for comparison, we also examined results obtained with four
{\em Herschel} bands (160--500\,$\mu$m; including the bias corrections) and, finally,
with three {\em Herschel} bands but without any bias corrections.
The non-linear fits were made using MCMC (with $2\times 10^5$ samples per field) and bootstrap sampling (2000 realisations per field). Both fits used total least-squares\footnote{Distance between a ($\tau(J)$, $\tau(250\mu {\rm m})$) point and the
model curve is measured in a coordinate system where the error region of the point is
circular. We used the smallest distance to the curve, ignoring the marginal effect
resulting from the curvature of the model curve.} and the error estimates of the
individual points. The parameter values and the uncertainties derived with these two
methods should be similar, except for rare cases in which the result depends on a small
number of influential points, which are always present in the MCMC calculation, but not in
all bootstrap samples.
Figures~\ref{fig:indi_1}-\ref{fig:indi_2} show the results of these fits. Each frame
shows the values of the linear slopes for the three cases discussed above. The
parameters of the non-linear fits are shown together with the error estimates
calculated with the bootstrap method. In addition to the fit to the default data set
(solid red curve), the dashed magenta lines show the effect of the distance
uncertainty. Using the distance uncertainties $\delta d$ listed in
Table~\ref{table:fields}, we also calculated the bias corrections for $\tau(J)$ for
distances $d-\delta d$ and $d+\delta d$. Thus, the upper dashed line corresponds to
distance $d-\delta d$ and a smaller bias correction.
Figure~\ref{fig:indipara} shows the linear slopes $k$ and the values of parameter $C$ for
this sample of fields. The fits were performed to all data, without a $\tau(J)$
threshold. If the data below $\tau(J)=0.6$ were removed, the median slope $\tau(250\mu
{\rm m})/\tau(J)=1.6\times 10^{-3}$ did not change appreciably (by less than 0.1$\times
10^{-3}$). The median value of $C$ increased to $2.0\times 10^{-4}$, which in that
case is still lower than the scatter. The differences between the fields are larger than the estimated formal uncertainties (including the statistical errors of
$\tau(J)$ given by NICER and $\tau(250\mu{\rm m})$ derived from the uncertainty of the
surface brightness measurements). The error bars only reflect the statistical errors of
the fits and do not include the uncertainty of the bias correction, for example.
Figure~\ref{fig:indi_1} demonstrates that the uncertainty of the distances can be a
significant source of error. In Fig.~\ref{fig:indipara}, the shaded areas show the
difference between the values obtained with distances $d - \delta d$ and $d + \delta d$,
as listed in Table~\ref{table:fields}. These were estimated directly by repeating the
analysis using these two distance values. A smaller distance corresponds to smaller bias
correction in $\tau(J)$ and, thus, to a steeper slope $k$ and typically a higher value of
$C$.
In some cases, the value of $C$ obtained with the default distance $d$ is
outside the shaded region, showing that the effect is not always this simple. The
distance uncertainty is not yet enough to explain all the scatter in $k$ and especially
in $C$.
A change in the distance estimate results at first approximation in a nearly linear
scaling of $\tau(J)$ values (see Fig.~\ref{fig:indi_1}, comparison of the dashed
magenta lines).
In reality, the situation may be more complex. In particular, if a field contains cloud
structures at different distances, this might result in large errors in both $k$ and
$C$. Figure~\ref{fig:indipara} also shows that in spite of the small formal errors of
the least-squares fits, we cannot constrain the opacity values in the last two fields
with distances exceeding 2\,kpc.
\begin{center}
\begin{figure}
\includegraphics[width=8.5cm]{plot_indi_para_fwhm180_TAUJ_MIN_-99_0.jpg}
\caption{
Linear slope $k$ (upper frame) and the parameters $C$ (lower frame) for the 23 selected fields. The values obtained without bias corrections are shown with black symbols. The
values obtained with corrected $\tau(J)$ and ${\rm \tau(250\mu m)}$ data are shown with
red symbols, the shaded area corresponding to the uncertainty of the bias correction that
is due to the uncertainty of the distance estimates. The dashed lines show the median
values corresponding to the black and red symbols.
The fields are arranged in order of increasing distance, and the ${\rm \tau(250\mu m)}$
values are based on SPIRE data alone.
}
\label{fig:indipara}
\end{figure}
\end{center}
In the sample of Fig.~\ref{fig:indipara}, the median value of $C$ is close to zero with
a number of fields with negative values. The positive values of $C$ in
Fig.~\ref{fig:MCMC_TLS} are due to a small number of fields, and the increase of
$\tau(250\mu {\rm m})/\tau(J)$ values was only visible above $\tau(J)\sim 5$. There are
only 20 fields with any data points above $\tau(J)=5$. Only six fields have ten or more
data points above this limit:
G6.03+36.73, G70.10-1.69, G82.65-2.00 G92.04+3.93 G107.20+5.52, and G202.02+2.85. Of
these, only G6.03+36.73 is included in the sample of Fig.~\ref{fig:indipara}. All the
others were excluded because the bias correction changed the slope $k$ by more than
30\%. Thus, a clear steepening of the relation $\tau(250\mu {\rm m}$ vs. $\tau(J)$ is
seen exclusively in fields with the highest column densities, which for the same
reason also have the largest uncertainty regarding the bias corrections.
\subsection{Maps of $\tau(250\mu {\rm m})/\tau(J)$ ratio}
We also examined the ratios $\tau(250\mu {\rm m})/\tau(J)$ in the form of maps. This is
useful if $k$ changes in small regions that have little effect when all data of a field
are fitted. Unlike in Fig.~\ref{fig:indi_1}, where the offset between $\tau(250\mu {\rm
m})$ and $\tau(J)$ is a free parameter, the appearance of the ratio maps depends on the
consistency of the $\tau(250\mu {\rm m})$ and $\tau(J)$ zero points. Because we do not
have an absolute zero point for $\tau(J)$, we used the reference areas listed in
Table~\ref{table:fields} and subtracted from $\tau(250\mu{\rm m})$ and $\tau(J)$
the average value found in the reference area. This limits the region where a reliable
ratio can be calculated, excluding regions of low column density. The details of the
calculations are given in Appendix~\ref{sect:ratiomaps}, where we also show the figures of
selected fields. As an example, Fig.~\ref{fig:ratio_map} shows the field
G4.18+35.79 (LDN~134), where the ratio $\tau(250\mu {\rm m})/\tau(J)$ is strongly
correlated with column density.
In Fig.~\ref{fig:ratio_map}, the increase of $k$ remains clear even in maps of
$(\tau(250\mu {\rm m}) \pm \delta \tau(250\mu {\rm m})/(\tau(J) \pm \delta \tau(J))$.
The error estimates $\delta \tau(250\mu {\rm m})$ include the statistical errors due to
{\it Herschel} photometry and uncertainty of the surface brightness zero point (see
Sects.~\ref{sect:tau250} and~\ref{sect:zeropoint}).
The parameter $\delta \tau(J)$ corresponds to the uncertainty of the $\tau(J)$ zero
point (see Appendix~\ref{sect:ratiomaps}).
For $\tau({\rm J})$ the formal error estimates calculated with NICER are below
10\%.
For high-opacity sources like G4.18+35.79 the bias correction of $\tau(J)$ is very
important. If no background stars are visible through some part of the core, the
values of $k$ naturally remain more uncertain (see, however,
Sect.~\ref{sect:biasbias}).
Conversely, the zero-point uncertainty only becomes important in diffuse regions but
might even reverse the correlation with column density.
The ratio maps are also affected by the assumption of a constant value of $\beta$ and
of potential errors in the bias corrections. However, we argue in
Sect.~\ref{sect:biasbias} that these are mainly multiplicative errors (that do not affect
the morphology of the maps) and/or tend to decrease the variations seen in the ratio
maps. Therefore, we are confident that the increase of submillimetre opacity that is
seen in some of the maps is real.
Based on the maps, the submillimetre opacity is correlated with column density in
the fields G4.18+35.79, G6.03+36.73, G111.41-2.95, G161.55-9.30, G151.45+3.95, and
G300.86-9.00 (see Appendix~\ref{sect:ratiomaps}). In G4.18+35.79 and G6.03+36.73 the
values rise to close to $4 \times 10^{-3}$. In some fields the background subtraction
reduces the available map area to such an extent that no conclusions can be drawn.
\begin{center}
\begin{figure}
\includegraphics[width=8.8cm]{ratio_G4_18+35_79_TAUJZP.jpg}
\caption{
Field G4.18+35.79 (LDN~134). The upper frames show $\tau(250\mu {\rm m})$ (frame a) and
the ratio $\tau(250\mu {\rm m}) / \tau(J)$ (frame b). The lower frames show the lower
(frame c) and upper (frame d) limits of $\tau(250\mu {\rm m})/\tau(J)$ calculated as
$(\tau(250\mu {\rm m})+\delta \tau(250\mu {\rm m}))/(\tau(J)-\delta \tau(J))$ and
$(\tau(250\mu {\rm m})-\delta \tau(250\mu {\rm m}))/(\tau(J)+\delta \tau(J))$. The areas
not covered by {\it Herschel} observations and regions with a
SN below 0.5 have been
masked. In frame $a$, the solid black contour and the dashed white contour correspond to
$\tau(250\mu{\rm m}= \delta \tau(250\mu{\rm m})$ and $\tau(J)=\delta \tau(J)$. The maps have a resolution of 180\,$\arcsec$ and $\tau(J)$ is derived
using 2MASS data.
}
\label{fig:ratio_map}
\end{figure}
\end{center}
\begin{center}
\begin{figure*}
\sidecaption
\includegraphics[width=12cm]{kC_tables.jpg}
\caption{
Comparison of parameters $k$ (upper frames) and $C$ (lower frames) obtained with data at resolutions of
180$\arcsec$ and 120$\arcsec$ . Frames $a$ and $c$ show fits to all data,
frames $b$ and $d$ fits to values ${\tau(J)}>0.6$ alone. The open circles and crosses show
the estimates obtained with 2MASS data at resolutions of 180$\arcsec$ and 120$\arcsec$ . The filled triangles show the results for VISTA observations, the larger
triangles corresponding to a resolution of 180$\arcsec$ , the smaller to a resolution of 120$\arcsec$
. The fields are the same as in Fig.~\ref{fig:indipara}, with the addition of
G358.96+36.75, which has fewer data points above ${\tau_J>0.6}$ and
for which parameter $C$ could not be fitted with extinction maps
with a resolution of 120$\arcsec$ .
}
\label{fig:kC_tables}
\end{figure*}
\end{center}
So far, all NIR extinction maps were calculated at 180$\arcsec$ resolution. Depending
on the number of background stars, extinction map could be derived at a higher
resolution and possibly with smaller bias. This especially applies to the four fields
for which VISTA observations are available.
We recalculated the extinction maps at 120$\arcsec$ resolution, repeating Monte Carlo
simulations to estimate the bias of ${\tau(J)}$. The smaller beam increases the noise
per resolution element, but does not yet cause holes in the extinction maps. The
analysis was repeated for the four fields with VISTA data with resolutions of 180$\arcsec$ and
120$\arcsec$ .
The results are summarised in Fig.~\ref{fig:kC_tables}. The resolution has no strong
systematic effect on the parameters. The largest differences appear when parameter $C$ is
estimated excluding low column density points. However, even in that case the
difference between the results with a resolution of 120$\arcsec$ and 180$\arcsec$ is smaller than the
effect of excluding low $\tau(J)$ data from the fits. When VISTA data are available,
the results are close to those obtained with 2MASS.
Because we used the same {\it Herschel} data and bias corrections derived in the
same way, the results are not independent. However, because the uncertainty of
$\tau(J)$ is expected to be one of the most significant sources of error, this gives
us some confidence that the observed differences between the fields are real.
The ${\tau(250\mu {\rm m})/\tau(J)}$ ratio in the field G4.18+35.79 was shown in
Fig.~\ref{fig:ratio_map}. Figure~\ref{fig:VHS_map} shows the corresponding figure
obtained with VISTA NIR data and a spatial resolution of 120$\arcsec$. The highest
value of ${\tau(250\mu {\rm m})/\tau(J)}$ has increased from $3.6 \times 10^{-3}$ in
Fig.~\ref{fig:ratio_map} to $6.7 \times 10^{-3}$. This is mainly attributed to the
increased spatial resolution, although also at the 180$\arcsec$ resolution the peak
value is $\sim$25\% higher than in Fig.~\ref{fig:ratio_map}. Even in VISTA data there
are only $\sim$10 stars within the $2\arcmin \times 2\arcmin$ area centred on the
$\tau(250\mu{\rm m}$ maximum, and therefore the peak value of ${\tau(250\mu {\rm
m})/\tau(J)}$ is subject to some uncertainty.
\begin{center}
\begin{figure}
\includegraphics[width=8.8cm]{ratio_G4_18+35_79_TAUJZP_VSA_120.jpg}
\caption{
Ratio ${\tau(250\mu {\rm m})/\tau(J)}$ in field G4.18+35.79 at a resolution of
120$\arcsec$, based on VISTA NIR data. Frame $a$ shows a map of $\tau(250\mu {\rm m}),$
frame $b$ a map of the ratio $\tau(250\mu {\rm m}) / \tau(J)$. Frames $c$ and $d$
are estimated lower and upper limits of $\tau(250\mu {\rm m})/\tau(J)$ (cf
Fig.~\ref{fig:ratio_map}).
}
\label{fig:VHS_map}
\end{figure}
\end{center}
\subsection{Potential systematic errors} \label{sect:biasbias}
Because of the significance of the bias corrections (Sects.~\ref{sect:bias_tauJ} and~\ref{sect:bias_tau250}), we tried to characterise the effects that systematic
errors in these corrections could have on the $\tau(250\mu {\rm m})/\tau(J)$ ratios.
Furthermore, the assumption of a constant dust emissivity spectral index may be
incorrect. Below we examine the possible systematic effects caused by these factors.
The bias correction made to the ${\rm \tau(250\mu m)}$ values is in itself small (see
Fig.~\ref{fig:hist_bias}) and, consequently, the errors made in that correction are expected to be small. The correction was derived from radiative transfer models with
dust properties corresponding to $R_{\rm V}$=5.5 \citep{Draine2003}. If the submillimetre
dust emissivity is in fact higher, the dust column density will be overestimated and
models will also overestimate the cloud opacity at visual and NIR wavelengths, thus
exhibiting stronger temperature variations than the real clouds. Because the ${\rm
\tau(250\mu m)}$ bias is related to the line-of-sight temperature variations, we could in
this case systematically overestimate the bias in ${\rm \tau(250\mu m)}$. To check the
magnitude of the effect, we repeated the modelling using a dust model with higher long-wavelength emissivity. We used a dust model from \citet{Ossenkopf1994} with coagulated
grains with thin ice mantles accreted in 10$^5$ years at a density of 10$^6$\,cm$^{-3}$.
Compared to NIR (the wavelengths contributing to most of the heating deep inside a
cloud), this dust model has an emissivity higher by $\sim$50\% at SPIRE wavelengths. The
results are shown in Fig.~\ref{fig:MOD_BIAS}, the open circles corresponding to this
alternative modelling. Because of the smaller estimated bias, these should be below the
$k$ values of our previous analysis (red symbols). The median value of $k$ is 1.58$\times
10^{-3}$ and thus practically unchanged. The strongest change is seen for
G6.03+36.73 (LDN~183), where the estimate has been reduced by more than 20\%. This is a
source with very high column density and thus, in its central parts, a very uncertain
estimate of ${\rm \tau(250\mu m)}$. However, the uncertainty of ${\rm \tau(250\mu m)}$
bias must also be considered in connection with the uncertainty of the ${\tau(J)}$
correction, which could compensate for some of the change (see below).
The ${\tau(J)}$ bias corrections are more significant than the ${\rm \tau(250\mu m)}$
bias corrections. The estimation of the ${\tau(J)}$ bias is, in principle, more reliable because it only
depends on the assumption of the ${\tau(J)}$ structure of the clouds. In our
calculations (see Sect.~\ref{sect:bias_tauJ}), $\tau(J)$ was derived from {\it Herschel}
observations, dividing ${\rm \tau(250\mu m)}$ by the constant factor of $k=$1.5$\times
10^{-3}$ to obtain a template map of $\tau(J)$. There are two possible sources of error.
First, if the targets contain much structure below the 18$\arcsec$ resolution of the
${\rm \tau(250\mu m)}$ maps, we will underestimate the bias and our $k$ estimates will be
too high. We cannot directly estimate this error, but it is expected to be a small
fraction of the total bias estimate. This is because $18\arcsec \ll 3\arcmin$ and, at
least for the closest fields, {\it Herschel} already resolves most of the
cloud structure.
The second potential source of error is again connected with dust emissivity. If the
local ratio between ${\rm \tau(250\mu m)}$ and ${\tau(J)}$ is higher than 1.5$\times
10^{-3}$ (strongly increased submillimetre opacity), we have
overestimated the cloud opacity at NIR wavelengths in the modelling, the bias correction of ${\tau(J)}$ is
too large, and we underestimate the true value of $k$. Thus, a change in the value of $k$
that is used to estimate the ${\tau(J)}$ bias will change the recovered value of $k$ in
the same direction. To examine this potential problem more quantitatively, we repeated
the bias estimation using $k$ values of 1.0$\times 10^{-3}$ and 3.0$\times 10^{-3}$. The
resulting values of $k$ and $C$ parameters are shown in Fig.~\ref{fig:MOD_BIAS} as
triangles. The initial assumption of $k=$1.0$\times 10^{-3}$ leads to a recovered median
value of $k=$1.26$\times 10^{-3}$. The initial assumption of $k=$3.0$\times 10^{-3}$
leads to a recovered median value of $k=$1.92$\times 10^{-3}$. In both cases the input
and output values are inconsistent, unlike in our previous analysis,
where an assumption of 1.5$\times 10^{-3}$ led to a recovered value of 1.6$\times
10^{-3}$. Furthermore, an error in the assumed value of $k$ leads to a systematic error
in the recovered value that is about half of the original error, even lower if the true value
of $k$ was initially overestimated.
The previous test shows that the estimates of $k$ will be biased towards the selected
value of 1.5$\times 10^{-3}$. Our previously recovered median value of 1.6$\times
10^{-3}$ is thus not significantly affected (bias lower than 0.05), but the effect can be
stronger for individual fields. For example, in G4.18+35.79 the estimate was 1.8$\times
10^{-3}$ , but the true value is probably higher by $\sim$10\%. The calculations could
be iterated, field by field, to carry out the bias correction self-consistently with the
final $k$ estimate. However, the errors are typically below 10\%, and rough estimates of
their magnitude can be seen in Fig.~\ref{fig:MOD_BIAS}.
There is a specific consequence of the way the $\tau(250\mu {\rm m})$ and ${\tau(J)}$
bias corrections are implemented. If a cloud included regions of such a high opacity that
no 2MASS stars were visible through the cloud, the ratio of $\tau(250\mu {\rm m}$ and
$\tau(J)$ would normally be overestimated. The resulting apparent increase of
submillimetre opacity could thus be an artefact resulting from errors in extinction
values. However, in our analysis we also correct ${\tau(J)}$ in this case based on
the assumed opacity derived using the $\tau(250\mu {\rm m})$ input map and the extinction
calculated with simulated 2MASS stars. If the gap in the distribution of background stars
is increased, the ratio $\tau(250\mu {\rm m})/\tau(J)$ does not continue to increase, but
instead tends towards the assumed ratio, 1.5$\times 10^{-3}$. Higher values should thus
not be the result of gaps in extinction data.
To investigate the potential effects of a spatially varying spectral index, we repeated
the analysis using an ad hoc $\beta(T)$ law to introduce $\beta$ variations in all of our
maps. We took the temperatures calculated with $\beta=2.0$ and fixed new $\beta$ values
pixel by pixel using a functional dependence $\beta=2.0 \times (T/15.0)^{-0.24}$. We then
repeated the full analysis, starting with the zero point and colour corrections and
continuing with the calculation of colour temperatures and submillimetre opacity. We did
not solve the ($T$, $\beta$) values (which are very susceptible to noise effects), but
simply assumed that $\beta$ could vary in a systematic way so that the
values are higher when the dust temperature is lower. The parameters of the $\beta(T)$
formula were selected so that $\beta$ changes from $\sim 1.8$ in warm regions with $T\sim
23$\,K to $\sim 2.2$ in the coldest spots $T\sim 10$\,K \citep[cf][]{Dupac2003,
Desert2008, planck2011-7.7b, Paradis2010, Veneziani2010, Juvela2011}. The average $\beta$
is still close to the original $\beta=2.0,$ and we mainly examined the effects of
correlated changes of $\beta$ rather than the effects of absolute $\beta$ values that can
be estimated more directly.
The crosses in Fig.~\ref{fig:MOD_BIAS} show the slopes $k=\Delta \tau(250\mu {\rm m}/
\Delta\tau(J)$ and the parameters $C$ obtained with the spatially varying $\beta$. The
values are typically within $\sim$10\% of the values obtained with $\beta=2.0$. The
strongest changes of $k$ are seen in the two closest fields, G4.18+35.79 and G6.03+3673,
which have well-resolved clumps with very low temperatures. The increase of the slope
values is $\sim$20\%. Note that this is mostly consistent with the general dependence
between $\tau(250\mu {\rm m})$ and $\beta$ and not necessarily an effect of the spatial
variation of $\beta$. In this respect, it is very interesting to note that the effect on
the parameter $C$ is weak. In other words, if the variations of $\beta$ are as assumed
above, this will be reflected in the slope $\tau(250\mu {\rm m})/\tau(J),$ but without a
noticeable non-linearity in the $\tau(250\mu {\rm m})$ vs. $\tau(J)$ relation.
All the above suggests an uncertainty of 10--20\% in $k$ and $\sim$0.1 units in $C$. In
particular, if the increase of dust opacity is associated with values $\beta>2.0$, our
highest $k$ estimates could still systematically underestimate the true values of $k$
because of the lower assumed value of $\beta$ and because the $\tau(J)$ correction biases
the $k$ values towards 1.5$\times 10^{-3}$.
\begin{center}
\begin{figure}
\includegraphics[width=8.5cm]{plot_indi_para_fwhm180_TAUJ_MIN_-99_0_MOD.jpg}
\caption{
As Fig.~\ref{fig:indipara}, but comparing our default $\tau(250\mu {\rm m})/\tau(J)$
estimates (red solid circles) and results from alternative analyses: $\tau(250\mu {\rm
m}$ estimates derived assuming a spatially varying spectral index (crosses), $\tau(250\mu
{\rm m}$ bias estimated with \citet{Ossenkopf1994} dust model (open circles), $\tau(J)$
bias estimated using $k=1.0\times 10^{-3}$ (triangles pointing upwards), and $\tau(J)$
bias estimated using $k=3.0\times 10^{-3}$ (triangles pointing downwards).
}
\label{fig:MOD_BIAS}
\end{figure}
\end{center}
\section{Discussion} \label{sect:discussion}
We have examined the submillimetre opacity by correlating the 250\,$\mu$m optical depth
${\rm \tau(250\mu m)}$ with the near-infrared optical depth, $\tau(J)$, assuming the
latter to be an independent tracer of the total dust column density.
Because the comparison was made at a resolution of 2$\arcmin$ , it is not sensitive to
dense cores ($A_{\rm V}>>10$\,mag), at which both the ${\rm \tau(250\mu m)}$ and $\tau(J)$
estimates would become very uncertain. Nevertheless, corrections for systematic bias in
${\rm \tau(250\mu m)}$ and especially in $\tau(J)$ are important.
The sample consists of the heterogeneous set of 116 Galactic fields that were mapped
with {\it Herschel} as part of the {\em Galactic Cold Cores} project. The main
objectives were to estimate the typical ratio of ${\rm \tau(250\mu m)/\tau(J)}$ and to
search for variations of this quantity, between the fields and as function of column
density. Such variations were then related to differences in the properties of
interstellar dust grains. For the present sample, high column densities also imply low
dust temperatures and thus conditions where submillimetre dust opacity is expected to be
enhanced by grain aggregation.
The limited resolution means that we did not probe the full range of opacity
variations, if these are partly limited inside compact cores.
\subsection{Main results and their reliability} \label{sect:summary}
By restricting the analysis to $\sim$20 fields for which the results appeared most reliable, we
derived a median value of $k={\rm \tau(250\mu m)/\tau(J) = 1.6 \times 10^{-3}}$ (see
Sect.~\ref{sect:individual} and Fig.~\ref{fig:indipara}). In Fig.~\ref{fig:indipara},
50\% of the most nearby fields had positive values of $C$, the multiplier of the second-order term, indicating some degree of positive correlation between column density and
submillimetre emissivity measured by $\tau(250\mu {\rm m})$. In the maps of the ratio
$k$, the same tendency was very clear in only six cases. The low percentage is partly
caused by the noise in ${\tau(J)}$, whose magnitude is strongly correlated with
the cloud distances. For two of the best examples, G4.18+35.79 and G6.03+36.73, the ratio
${\rm \tau(250\mu m)/\tau(J)}$ increases to $\sim 4\times 10^{-3}$, almost a factor of
three higher than the median value. The peak values are uncertain because of the large bias
corrections, but because of the low spatial resolution used, the strongest effect can be
even greater.
No dependence on either Galactocentric distance or on Galactic height was observed
(Fig.~\ref{fig:vs_distance}). The reliability of the estimates decreases with distance,
but nevertheless, our sample extends over more than $\sim$4\,kpc in Galactocentric
distance. No trends are seen; if they were present at 10\% level, they should still be visible
over the scatter of individual data points. The result might be affected by systematic
errors in the bias correction. However, these probably depend either on the distance or
on the morphology of the field and at first approximation are
not expected to be different in the
inner and in the outer Galaxy.
In Fig.~\ref{fig:indipara}, the only clear trend is the decrease of $k$ as a function
of distance (also visible as larger scatter around the Galactocentric distance of
8.5\,kpc). As mentioned in Sects.~\ref{sect:bias_tauJ} and~\ref{sect:bias_tau250},
there are two possible explanations. First, the bias correction of $\tau(J)$ might be
overestimated so that as the distance increases, the error increases and $k$ becomes
underestimated. This is probably not the main reason because the bias correction
depends as much on column density values and column density gradients as on distance.
The trend probably is a combination of selection and resolution effects. At
1\,kpc the 180$\arcsec$ resolution corresponds to almost 1\,pc in linear scale.
Therefore, we measure the mean cloud properties for the most distant fields.
In the nearby fields individual clumps are resolved and the slopes $k$ reflect more
the contrast between diffuse regions and compact, core-sized objects. Thus, the trend
would be compatible with the hypothesis that $\tau(250\mu {\rm m})/\tau(J)$ increases
in the densest and coldest regions of interstellar clouds.
The global non-linear fits of Fig.~\ref{fig:MCMC_TLS} also indicated an increase of
the ratio $k=\tau(250\mu {\rm m})/\tau(J)$ as the function of column density. The plots
show clear deviations from a linear dependence beyond $\tau(J)\sim 4$. For $\tau(J)\sim
10$, $k$ is already twice as high at low column densities. The results are dependent
on a few fields with the highest column densities for which the bias
corrections reach about ten per cent. However, if the distance
dependence of
Fig.~\ref{fig:vs_distance} means that the bias correction of ${\tau(J)}$ is probably overestimated and not underestimated, the true values of $k$ might be even
higher.
Figure~\ref{fig:indipara} concentrated on selected fields for
which linear and non-linear
fits were more reliable than on average. Within this subset, no clear distance
dependence was visible in $k$ values, but the scatter is larger than the estimated
uncertainties. The linear slopes are sensitive to the highest column densities within a
field. The parameter $C$ of the non-linear fits is expected to
be even more sensitive to these
data points, which probe the variations of $\tau(250\mu m)/\tau(J)$ inside the fields. For
the whole sample, the median value of $C$ was very close to zero.
Nevertheless, it may be significant that when the two fields at kiloparsec
distances are excluded, the value of $C$ appears to decrease somewhat systematically with the
distance of the field. There is some preference for the most nearby fields to have positive values; here, the
densest parts of the clouds are resolved.
Two of the first fields with high $k$ values are G6.03+36.73 and G4.18+35.79, better
known as LDN~183 and LDN~134. At the {\it Herschel} scale, the clouds have a simple
morphology, each consisting of a single column density maximum that is partially resolved
by the 180$\arcsec$ beam. This is illustrated by Figs.~\ref{fig:ratio_map} and
\ref{fig:ratiomaps}. The ratio $k=\tau(250\mu {\rm m})/\tau(J)$ closely follows the
morphology of the column density distribution, making these the best examples of clumps
with increased submillimetre opacity. The highest values are $\sim 4\times 10^{-3}$,
almost three times the average value over all fields.
In our results, some of the main sources of uncertainty are the corrections made for
the expected bias in $\tau(J)$ and, to lesser extent, in ${\rm \tau(250\mu m)}$.
According to Fig.~\ref{fig:indipara}, the net effect of all bias corrections is a
$\sim$20\% decrease in the value of $k$. This number applies to nearby fields but is
more dramatic if all fields are taken into account. Figure~\ref{fig:vs_distance} shows
that for fields at $\sim$1\,kpc distance the correction is almost a factor of two. The
corrections have been remarkably successful in decreasing the scatter of $\tau(250\mu
m)/\tau(J)$ values, which strongly suggests that they are approximately of the
correct magnitude.
In the statistical sense, estimating the ${\tau(J)}$ bias is straightforward,
using a model of the Galactic stellar distribution and the higher resolution {\it
Herschel} data as a template for the column density structure of the field (see
Appendix~\ref{sect:NIR_simu} and Sect.~\ref{sect:biasbias}).
The correction is large, but {\em \textup{on average}}, the correction itself probably does not suffer
from major systematic errors. Errors could arise either from incorrect distance estimates
or from the use of an incorrect model of the NIR opacity distribution. The distances
are uncertain, and through the ${\rm \tau(J)}$ bias, their effect on the main parameters is
shown in Fig.~\ref{fig:indipara} (the grey bands) and in Fig.~\ref{fig:indi_1}. For a
sample of fields, this is mainly a statistical and not a systematic error.
The model of $\tau(J)$ distribution in each field was derived from {\it Herschel} data
at 18$\arcsec$ resolution. If there is still significant structure below this scale,
our correction of $\tau(J)$ values will be too low. The difference of the
2-3$\arcmin$ scale and the 18$\arcsec$ scale is so large, however, that most of the
effects of column density variations are already included. Thus, after the distance,
the other main error in the $\tau(J)$ bias correction arises from scaling the {\it
Herschel} estimates of ${\rm \tau(250\mu m)}$ into a template of the J-band opacity. We
showed in Sect.~\ref{sect:biasbias} that this amounts to an uncertainty
of $\sim$10\% in the
final listed values of $k=\Delta \tau(250\mu {\rm m})/ \Delta \tau(J)$. For the sample
in Fig.~\ref{fig:indipara}, the value of $k$ assumed during the bias correction is
consistent with the recovered value (to be precise, the assumed value of $k=1.5\times 10^{-3}$
, the recovered value of $k=1.6\times 10^{-3}$). For individual fields, the values are
biased towards the initial assumed value. Figure~\ref{fig:MOD_BIAS} showed that if bias
correction assumed a value of $k=1.0\times 10^{-3}$, the recovered median value was
still higher than $k=1.2\times 10^{-3}$. This shows that $k$ cannot be significantly
overestimated because of an erroneous bias correction in $\tau(J)$. Thus, the result
that the average ratio $k=\Delta \tau(250\mu m)/ \Delta \tau(J)$ is clearly higher than the
normal value found in diffuse medium remains robust.
Because of the bias in the $\tau(J)$ correction, our estimates are probably conservative
for fields where values $k>1.5\times 10^{-3}$ were obtained. The dependence on the
assumed value of $k$ decreases as the true value of $k$ increases. This is because
higher $k$ means a lower NIR opacity and lower overall bias in $\tau(J)$. Nevertheless,
for fields like G4.18+35.79, we might systematically underestimate $k$ by
$\sim$10\% because of this error in the $\tau(J)$ bias correction. At the highest column
densities {\it Herschel} emission data themselves will underestimate the cloud opacity, which leads to the corrections discussed in\
Sect. \ref{sect:bias_tau250}. For the optical
depth ratio, the effect of ${\rm \tau(250\mu m)}$ bias is weaker than that of
$\tau(J)$. Nevertheless, the errors made in the corrections of ${\rm \tau(250\mu m)}$ and $\tau(J)$
(for example, those associated with a change of submillimetre opacity) would
partly cancel each other out.
In the analysis we assumed that apart from the problems associated with the
sampling provided by the background stars, NIR reddening is an independent and reliable
measure of column density. Unlike in the optical range, the NIR extinction curve is
often assumed to be constant for a wide range of column densities \citep{Cardelli1989,
MartinWhittet1990, Roy2013}. Nevertheless, some cloud-to-cloud variations are observed,
the ratio $E_{J-H}/E_{H-K}$ ranging from values lower than 1.5 to higher than 2.0 \citep{Racca2002,
Draine2003ARAA}. Clear changes take place at high optical depths, above $A_{\rm V} \sim
20$\,mag, but only in the form of the flattening of the MIR extinction curve, at
wavelengths above 3\,$\mu$m \citep{Indebetouw2005, Cambresy2011, Ascenso2013}. Recently,
\citet{Whittet2013} studied cloud LDN~183 (which is also included in our sample).
Comparison with the 9.7\,$\mu$m silicate absorption feature suggested that the NIR
colour excess might not be a perfect tracer of the total dust column. The ratio between
$E_{J-K}$ and the 9.7\,$\mu$m feature was observed to increase as the column density
exceeded $A_{\rm V} \sim$20\,mag. The observation might partly
arise because the
9.7\,$\mu$m feature is dampened by the formation of ice mantles \citep[see
also][]{Chiar2007}. However, if we assume that NIR extinction indeed \textup{{\em
\textup{overestimates}}} the total dust column in this object, the ratio of ${\rm \tau(250\mu
m)}$ relative to column density would increase by $\sim$20\% \citep[see][Fig.
12]{Whittet2013}. This is still a weaker effect than the increase by more than a
factor of two that was observed in $\tau(250\mu {\rm m})/\tau(J)$. We recall
that at high column densities, above $A_{\rm V}\sim 20$\,mag, the values of ${\rm
\tau(250\mu m)}$ are also uncertain and can be underestimated by a significant fraction
(\citet{Pagani2004} discussed a similar limitation at the nearby wavelength of
200\,$\mu$m).
Finally, we recall that $\tau(250\mu{\rm m})$ values were derived using a fixed value of
the spectral index, $\beta=2.0$. By assuming $\beta=1.8$ instead, the $\tau(250\mu{\rm
m})$ and $k$ values would decrease by up to 30\%. Conversely, if the value of $\beta$
increased towards dense and cold regions \citep{Dupac2003, Desert2008, planck2011-7.7b,
Paradis2010, Veneziani2010, Juvela2011}, we underestimate the
dust opacity changes if we use a constant value of $\beta$ . In this sense (and regarding possible bias in
${\tau(J)}$ corrections),
Figs.~\ref{fig:MCMC_TLS}--8 give conservative estimates of the
possible increase of submillimetre dust opacity. Clearly, the assumed
values of $\beta$ must also be taken into account when comparing our
results with other studies (see Sect. \ref{sect:comparison}). On the
other hand, our tests indicate that spatial variations of $\beta$
probably do not have a strong additional effect on $k$ (see
Sect.~\ref{sect:biasbias}).
\subsection{Comparison with other studies} \label{sect:comparison}
The submillimetre dust opacity has previously been studied in relation to both dust
extinction \citep{Terebey2009, Flagey2009, Juvela2011, Martin2012, Roy2013,
Malinen2013, Malinen2014} and to H{\sc I} \citep{Boulanger1996, Lagache1999,
planck2011-7.12, planck2011-7.13, Martin2012}. \citet{Boulanger1996} compared COBE
observations of FIR dust emission with the hydrogen 21\,cm observations. At high
latitudes, where H{\sc I} is a good tracer of the full gas column density, the derived
value was $\tau/N_H = 1.0 \times 10^{-25} \, (\lambda/250\mu{\rm
m})$\,cm$^{2}$\,H$^{-1}$. Similar values
were obtained in the first studies using {\it Planck} data \citep{planck2011-7.12}.
In the most recent {\it Planck} papers, somewhat lower values were reported, corresponding to
$\tau(250\mu {\rm m})/N_{\rm H} \sim 0.55 \times 10^{-25}$\,cm$^{2}$\,H$^{-1}$
\citep{planck2013-p06b, planck2013-XVII}. The change is associated with the revised
calibration of the highest frequency channels and, correspondingly, a lower value of
$\beta$. In \citet{planck2013-XVII} the spectral index above 353\,GHz was found to be
$\beta \sim 1.65$, lower than the value of 1.8 assumed in \citet{planck2011-7.12} or
the value of 2.0 used in \citet{Boulanger1996}. The higher dust opacity value of
\citet{Boulanger1996} is thus largely explained by the higher value of $\beta$.
The {\it Planck} result for the diffuse medium can be compared to our result most directly
by converting their $N_{\rm H}$ to $\tau(J)$. We can use the conversion factor $N({\rm
H}_2)/A_{\rm V}= 9.4 \times 10^{20}$\,cm$^{-2}$\,mag$^{-1}$ derived by \citet{BSD} at low
extinction, $E(B-V)$<0.5 \citep[see also][]{Nozawa2013}. Some studies \citep{Rachford2009,
planck2013-p06b, Liszt2014a} have found $N_{\rm H}/E(B-V)$ values that are 10--30\% higher
than in \citet{BSD}, these results apply partly to even more diffuse lines of sight
($E(B-V)\la 0.1$). On the other hand, \citet{Gudennavar2012} examined a sample of lines of
sight with $E(B-V)$ extending to values higher than one. The result, $N({\rm
H})/E(B-V)=(6.1\pm 0.2)\times 10^{21}$\,H\,cm$^{-2}$\,mag$^{-1}$, was close to that of
\citet{BSD}. With the \citet{BSD} relation and $R_{\rm V}=3.1$ extinction curve
\citep{Cardelli1989}, the {\it Planck} result $\tau(250\mu {\rm m})/N_{\rm H} \sim 0.55
\times 10^{-25}$\,cm$^{2}$\,H$^{-1}$ \citet{planck2013-p06b} corresponds to ${\tau(250\mu {\rm
m})/\tau(J)}=0.41\times 10^{-3}$. Our median value of ${\tau(250\mu {\rm
m})/\tau(J)}=1.6\times 10^{-3}$ is thus 3.9 times higher, and our highest local values,
4.5$\times 10^{-3}$ in G4.18+35.79, are higher by one order of magnitude. In
\citet{planck2013-p06b} the estimated value of $N({\rm H}/E(B-V))$ was $\sim$20\% higher
than in \citet{BSD}. With this value, our median value of ${\tau(250\mu {\rm m})/\tau(J)}$
would be $\sim 3.3$ times higher than the {\it Planck} estimate.
We can alternatively convert our result into $\tau(250\mu {\rm m})/N_{\rm H}$ , but in
dense regions the shape of the extinction curve (dependence on $R_{\rm V}$) and
the ratio between visual extinction and total column density are more uncertain. Using
the $R_{\rm V}=3.1$ and $N({\rm H}_2)/A_{\rm V}$ from \citet{BSD}, our median value
corresponds to $\tau(250\mu {\rm m})/N_{\rm H} = 2.16 \times
10^{-25}$\,cm$^{2}$\,H$^{-1}$ (again, of course, 3.9 times the {\em Planck} value).
However, $R_{\rm V}$ is expected to be higher than 3.1 in dense clouds. Using $R_{\rm
V}=5.5$ instead of $R_{\rm V}=3.1$ in converting $\tau(J)$ into visual extinction, the
values of $\tau(250\mu {\rm m})/N_{\rm H}$ would decrease by $\sim 15$\% (change in
$A_{\rm V}/E(J-K)$). However, the scaling we used above (corresponding to the $R_{\rm
V}$=3.1 extinction curve and the ratio of $N({\rm H}_2)/A_{\rm V}$ taken from
\citet{BSD}) corresponds to $N({\rm H})/E(J-K)=11.0\times
10^{21}$\,cm$^{-2}$\,mag$^{-1}$, which is very close to the value of $11.5\times
10^{21}$\,cm$^{-2}$\,mag$^{-1}$ that \citet{Martin2012} derived in Vela molecular
cloud using 2MASS NIR data up to $E(J-K)\sim0.55$\,mag ($E(B-V)\sim 1.1$\,mag).
When $R_{\rm V}$ is modified, the NIR extinction curve remains practically unchanged
and the differences take place between the optical and NIR wavelengths. Compared to shorter (optical) and longer (MIR) wavelengths, the extinction curve is
considered relatively constant in the NIR regime \citep[e.g.][]{Indebetouw2005,
Lombardi2006_Pipe, RomanZuniga2007, Ascenso2013, Wang2014} and the power-law slope of the
NIR extinction curve typically only varies at a level of $\sim$5\%
\citep{Stead2009, Fritz2011}. In very dense cores the formation of ice mantles and the
grain growth could have an additional impact. \citet{RomanZuniga2007} examined the
cloud Barnard 59 up to $A_{\rm V}=59$\,mag, but found no significant changes in the NIR
extinction law (compatible with $R_{\rm V}=5.5$). Similarly, \citet{Lombardi2006_Pipe}
found no changes in the Pipe nebula in the reddening NIR law up to $E(H-K)\sim
1.5$\,mag ($A_{\rm V}\sim 9$\,mag). Therefore, it is not likely that our results are significantly biased because of the assumed NIR extinction curve. At the resolution
of our extinction maps, practically all our data points are lower
than$A_{\rm V} \sim
10$\,mag and are thus not affected by extreme optical depths. However,
\citet{Whittet2013} observed a decrease in L183 in the ratio of $E(J-K)$ and the
9.7\,$\mu$m silicate absorption feature. Above $E(J-K)\sim 1.0$ ($A_{\rm J} \sim
1.7$\,mag, $A_{\rm V} \sim 5.0$\,mag for $R_{\rm V}=5.0$), the ratio deviated from
diffuse medium value by $\sim 20$\%. Stronger deviations were
only seen beyond
$E(J-K)\sim 3$ ($A_{\rm J} \sim 5$\,mag, $A_{\rm V} \sim 15$\,mag for $R_{\rm
V}=5.0$). This is above the range probed by our measurements. It is not clear that
changes in this ratio would only be caused by the NIR extinction curve. However, if,
in some sources, the extinction curve does flatten above $A(J)\sim 2$, this might
contribute to the observed increase of $\tau(250\mu{\rm m})/\tau(J)$ (possibly at the
20\% level).
The effect of the bias corrections on $\tau(250\mu{\rm m})/\tau(J)$ is typically
$\sim$20\% or weaker, and their uncertainty is smaller. The largest uncertainties in
$\tau(250\mu {\rm m})/N_{\rm H}$ are caused by $\beta$ and possibly by $R_{\rm V}$.
With $\beta=1.8$ and $R_{\rm V}$=5.5, the $\tau(250\mu{\rm m})/\tau(J)$ values would
be 40\% lower than with $\beta=2.0$ and $R_{\rm V}$=3.1. At this lower limit, our
median value of $\tau(250\mu{\rm m})/\tau(J)$ would not be 3.9 times, but $\sim$2.3
times the value found in the diffuse high-latitude sky. The 40\% uncertainty may be a
realistic 1 $\sigma$ lower limit for the linear fits that include all pixels in the
maps. However, it is a conservative estimate for the clumps where the average value of
$\beta$ is expected to be clearly higher than 1.8. In fact, preliminary results
indicate that the average value of $\beta$ in our fields (including the more diffuse
regions) is close to $\beta=1.9,$ and that with $\beta=2.0$ we underestimate the
$\tau(250\mu {\rm m})$ values of many clumps \citep{Juvela2014_TB}.
Increased far-infrared and submillimetre opacity has been reported by many authors
\citep[][etc.]{Kramer2003, Lehtinen2004, delBurgo2005, Ridderstad2010, Bernard2010,
Suutarinen2013}. One famous example is the Taurus filament L1506, for which the models of
\citet{Stepnik2003} suggested an increase by more than a factor of three. With the recent
detailed modelling of the dust properties and the structure of this filament,
\citet{Ysard2013} estimated the increase of the 250\,$\mu$m opacity to be $\sim$2.
\citet{Martin2012} studied the relation in the Vela cloud, comparing BLAST and IRAS data with
the reddening of 2MASS stars. The properties of the examined areas corresponded to the
average properties of our fields, with column densities extending to $10^{22}$\,cm$^{-2}$
and with typical dust temperatures of $\sim 15\,K$. They found a very similar range of
dust opacities, $\tau(250\mu m)/N_{\rm H} = (2-4)\times 10^{-25}$\,cm$^2$\,H$^{-1}$
{(assuming $\beta=1.8$)}.
In the Orion A cloud, a comparison of {\it Herschel} and 2MASS data led to the detection of
a dependence of $N^{0.28}$ on the 250\,$\mu$m opacity \citep{Roy2013}. The range of column
densities in Orion A was similar to our fields, and the derived dust opacities were mainly in
the range of $\tau(250\mu m)/N_{\rm H} = (1-3)\times 10^{-25}$\,cm$^2$\,H$^{-1}$. These
estimates were derived with $\beta=1.8$ and would become $\sim 20$\% higher (depending on
the details of the fitting) if a value of $\beta=2.0$ were used.
More recently, \citet{Lombardi2014} derived ratios $A({\rm K})/\tau(850\mu m)$ of
2640\,mag and 3460\,mag for the Orion A and B molecular clouds, respectively. The 850\,$\mu$m
optical depth was derived from {\em Herschel} observations rescaled using a comparison
with {\em Planck} measurements, and the NIR extinction was calculated with the method NICEST
\citep{Lombardi2009}. The modified blackbody fits used $\beta$ values that were estimated
in \citet{planck2013-p06b} at a resolution of 30$\arcmin$. With the reported average value
of $\beta=1.8$ and adopting a ratio 0.40 between $A({\rm K})$ and $A({\rm J})$, the
results correspond to ${\rm \tau(250\mu m)/\tau(J)}$ values of $1.56\times 10^{-3}$ and
$1.19\times 10^{-3}$ for Orion A and B, respectively. These fits were made for data ${\rm
\tau(850\mu m)}< 2\times 10^{-4}$ (${\rm \tau(250\mu m)} \la 1.8\times 10^{-3}$). The
optical depth range is similar to many of our fields, and the result for Orion A is close
to our median value for the fits concerning entire fields. By adopting $\beta=1.8$, our
median value would fall between the Orion A and Orion B estimates of
\citet{Lombardi2014}.
\subsection{Implications for dust evolution}
As shown by simulations, an observed increase of the dust far-IR/submm opacity towards
dense regions cannot be due to radiative transfer effects, but must originate in
intrinsic variations of dust properties \citep{Malinen2011, JuvelaYsard2012,
Ysard2012}. Theoretical studies have shown that such an increase, coupled with a
decrease in dust temperature, can be explained by the formation of large aggregate
particles \citep{Ossenkopf1994, Stognienko1995, Ormel2011, Kohler2011, Kohler2012}.
\citet{Kohler2012} showed that the coagulation of just four big grains of the diffuse
ISM type, coated by smaller carbon grains, already leads to an increase in the opacity
at 250\,$\mu$m of a factor 2.6. These authors also showed that these aggregates can
form within a typical cloud lifetime of 10 million years \citep{Walmsley1991}.
Consequently, we interpret the observed increase in ${\rm \tau(250\mu m)}$ in our cold
core sample as grain growth in dense molecular regions.
In our sample, the clouds LDN~183 and LDN~134 (G6.03+36.73 and G4.18+35.79) were the
most convincing examples of increased submillimetre dust opacity. LDN~183 has been
studied thoroughly in both continuum and line emission \citep[e.g.][]{Juvela2002,
Pagani2005, Pagani2007}. A slight increase of 200\,$\mu$m opacity was already reported
based on ISOPHOT observations \citep{Juvela2002}, but ISOPHOT data $\lambda \le 200\mu$m
and, to some extent even {\it Herschel} observations, are not sufficient to fully probe
the inner parts of the cloud where the high visual extinction is approaching 100\,mag
and the dust temperature drops well below 10\,K \citep{Pagani2004, Pagani2014}. LDN~183
was the first object where enhanced mid-infrared (MIR) light scattering, the so-called
coreshine phenomenon, was detected in Spitzer data \citep{Steinacker2010, Pagani2010}.
The effect was also seen in \citet{GCC-III}, where WISE MIR observations were analysed
and both LDN~183 and LDN~134 were found to be sources of strong MIR emission. If the
signal is interpreted as scattering of the interstellar radiation field, it seems to
imply the presence of very large, micrometre-sized dust particles \citep{Steinacker2010,
Steinacker2014, Lefevre2014}. Thus, our detection of increased submillimetre opacity in
these clouds agrees with the evidence provided by MIR wavelengths.
\section{Conclusions} \label{sect:conclusions}
We have examined dust optical depths by comparing measurements of submillimetre dust
emission and the reddening of the light of background stars in the near-infrared. The
goal was to measure the value of dust submillimetre opacity and to search for variations
that might be correlated with the physical state and the environment of the cloud. The
study led to the following conclusions:
\begin{itemize}
\item For a subsample of 23 fields with well-defined correlation between the
two variables, we obtained a median ratio of $\tau(250\mu {\rm m})/\tau(J) = (1.6 \pm 0.2)
\times 10^{-3}$. This is more than three times the value that was derived from {\it
Planck} data for the diffuse medium at high Galactic latitudes. Assuming $\beta=1.8$
instead of $\beta=2.0$, the value decreases by 30\%, but is still more than twice the
diffuse value.
\item The conversion to $\tau(250\mu {\rm m})/N_{\rm H}$ involves more assumptions.
Using the $R_{\rm V}=3.1$ extinction curve and the $N({\rm H}_2)/A_{\rm V}$ ratio of
\citet{BSD}, our median estimate corresponds to $\tau(250\mu {\rm m})/N_{\rm H} = 2.16
\times 10^{-25}$\,cm$^{2}$\,H$^{-1}$.
\item
The fit to all data above $\tau(J)=1$ gives a relation $\tau(250\mu {\rm m}) \sim
1.25 \times 10^{-3} \, \tau(J) + 0.11 \times 10^{-3} \, \tau(J)^2$. The positive
second-order coefficient $C=0.11 \times 10^{-3}$ is determined by a small number of
fields that, because of the high column density, are subject to large uncertainty in
the bias corrections.
\item
For the same sample, the scatter in the coefficients of the second-order terms $C$ is
$\sim 2 \times 10^{-4}$ and the median value is consistent with zero. Spatial
variations of the ratio $\tau(250\mu {\rm m}) / \tau(J)$ are only seen in a few
fields.
\item
From the maps of $\tau(250\mu{\rm m})/\tau(J)$, we identified six fields
where the ratio appears to increase further at the location of the main column
density peaks. The highest values in the fields G4.18+35.79 and G6.03+36.73 are $\sim
4 \times 10^{-3}$ at the resolution of 180$\arcsec$. Thus, although the densest
clumps are associated with the largest uncertainties, we consider this increase of
submillimetre opacity to be real.
\end{itemize}
\begin{acknowledgements}
This publication makes use of data products from the Two Micron All Sky Survey, which
is a joint project of the University of Massachusetts and the Infrared Processing and
Analysis Center/California Institute of Technology, funded by the National Aeronautics
and Space Administration and the National Science Foundation.
MJ, JMo, VMP, and JMa acknowledge the support of the Academy of Finland Grant No.
250741.
\end{acknowledgements}
|
\section{Introduction}
\noindent Deep convolutional neural network (CNN) has become a keen
tool in addressing large scale artificial intelligence tasks. Though
the study of CNN can be traced back to late 1980s
\cite{lecun89,lecun90}, the recent success of deep CNN is largely
attributed to the concurrent progresses of the two technical
streams. On the one hand, the new deep CNN architecture with
elements such as Dropout \cite{hinton12,KrizhevskySH12}, DropConnect
\cite{WanZZLF13}, Rectified Linear Units-ReLU \cite{NairH10} as well
as new optimization strategies \cite{DeanCMCDLMRSTYN12} have
empowered deep CNN with greater learning capacity. On the other
hand, the rapid advances and democratization of high performance
general purpose vector processing hardware, typified by graphics
processing unit (GPU), unleashes the potential power of deep CNN by
scaling up the network significantly.
Various infrastructures were used in scaling up deep CNNs, including
GPU \cite{CoatesHWWCN13}, distributed CPU based framework
\cite{DeanCMCDLMRSTYN12}, FPGA \cite{farabet09}, etc. Though the
implementation details among those approaches differ, the core
insight underlying the idea of scaling up deep CNN is
parallelization \cite{bengio07} in which vectorization technique is
the fundamental element. While the consecutive distinguished
performance of GPU trained CNNs in the ImageNet visual recognition
challenge \cite{KrizhevskySH12,RussakovskyDHBF13} as well as the
reported results in many studies in the literature justify its
effectiveness \cite{jia14,sermanet13}, the published literature did
not provide sufficient insights on how the vectorization was carried
out in detail. We also found there is no previous study to answer
how different degrees of vectorization influence the performance of
deep CNN, which is, however, crucial in finding the bottlenecks and
helps to scale up the network architecture. We believe these
questions form a significant research gap and the answer to these
questions shall shed some light on the design, tuning and
implementation of vectorized CNNs.
In this paper, we reinterpret the key operators in deep CNNs in vectorized forms with which high parallelism can be easily achieved given basic parallelized matrix-vector operators. To show the impact of the vectorization on the speed of both model training and testing, we developed and compared six implementations of CNNs with various degrees of vectorization. We also provide a unified framework for both high-level and low-level vision applications including recognition, detection, denoise and image deconvolution. Our Matlab Vectorized CNN implementation (VCNN) will be made publicly available on the project webpage.
\subsection{Related Work}
\noindent Efforts on speeding up CNN by vectorization starts with
its inception. Specialized CNN chip \cite{jackel90} was built and
successfully applied to handwriting recognition in the early 90s.
Simard et al. \shortcite{simard03} simplified CNN by fusing
convolution and pooling operations. This speeded up the network and
performed well in document analysis. Chellapilla et al.
\shortcite{chellapilla06} adopted the same architecture but unrolled
the convolution operation into a matrix-matrix product. It has now
been proven that this vectorization approach works particularly well
with modern GPUs. However, limited by the available computing power,
the scale of the CNN explored at that time was much smaller than
modern deep CNNs.
When deep architecture showed its ability to effectively learn
highly complex functions \cite{hinton06}, scaling up neural network
based models was soon becoming one of the major tasks in deep
learning \cite{bengio07}. Vectorization played an important role in
achieving this goal. Scaling up CNN by vectorized GPU
implementations such as Caffe \cite{jia14}, Overfeat
\cite{sermanet13}, CudaConvnet \cite{KrizhevskySH12} and Theano
\cite{bergstra10} generates state-of-the-art results on many vision
tasks. Albeit the good performance, few of the previous papers
elaborated on their vectorization strategies. As a consequence, how
vectorization affects design choices in both model training and
testing is unclear.
Efforts were also put in the acceleration of a part of the deep CNN
from algorithmic aspects, exemplified by the separable kernels for
convolution \cite{denton14} and the FFT speedup
\cite{mathieu2013fast}. Instead of finding a faster alternative for
one specific layer, we focus more on the general vectorization
techniques used in {\it all} building blocks in deep CNNs, which is
instrumental not only in accelerating existing networks, but also in
providing guidance for implementing and designing new CNNs across different platforms, for
various vision tasks.
\section{Vectorization of Deep CNN}
\noindent Vectorization refers to the process that transforms the
original data structure into a vector representation so that the
scalar operators can be converted into a vector implementation. In this section, we introduce vectorization strategies for different layers in Deep CNNs.
Figure \ref{fig:fig1} shows the
architecture of a typical deep CNN for vision tasks. It contains all of the essential parts of modern CNNs. Comprehensive introductions on
CNN's general architecture and the recent advances can be found in
\cite{lecun98} and \cite{KrizhevskySH12}.
We mark the places where vectorization plays
an important role. ``{\bf a}" is the convolution layer that
transforms the input image into feature representations, whereas
``{\bf b}" is the one to handle the pooling related operations.
``{\bf c}" represents the convolution related operations for feature
maps. We will see shortly that the vectorization strategies between
``{\bf a}" and ``{\bf c}" are slightly different. ``{\bf d}" involves operations in the fully connected network. Finally, ``{\bf e}" is the
vectorization operation required to simultaneously process multiple input samples
(e.g. mini-batch training). It is worth noting that we need to
consider both forward pass and back-propagation for all these
operations.
\begin{figure}[t]
\centering
\includegraphics[width=1.0\linewidth]{figures/cnn.eps}
\caption{Convolutional Neural Network architecture for visual recognition.}
\label{fig:fig1}
\end{figure}
\subsection{Vectorizing Convolution}
\label{sec:conv_vect} We refer to the image and intermediate feature maps as $f$, one of the
convolution kernels as $w_i$, the convolution layer can be typically
expressed as
\begin{eqnarray}
f^{l+1}_i = \sigma(w^l_i*f^{l}+b_i^l),
\end{eqnarray}
where $i$ indexes the $i^{th}$ kernel. $l$ indexes the layer. $b^l$ is the bias weight. $*$ is the convolution operator. For vision tasks, $f$ can be 2- or 3-dimension. The outputs from previous layer can be deemed as one single input $f^{l}$. $\sigma$ is the nonlinear function which could be ReLU, hyperbolic tangent, sigmoid, etc. Adding bias weight and applying nonlinear mapping are element-wise operations which can be deemed as already fully vectorized, i.e. the whole feature vector can be processed simultaneously. Contrarily, the convolution operators involve a bunch of multiplication with conflict memory access. Even the operators are parallized for each pixel, the parallelism \cite{ragan13} to be exploited is rather limited: compared to the number of computing units on GPU, the number of convolution in one layer is usually smaller. A fine-grained parallelism on element-wise multiplication is much preferred, leading to the vectorization process to unroll the convolution.
In what follows, all the original data $f$, $b$ and $w$ can be viewed as data vectors. Specifically, we seek vectorization operators $\varphi_c()$ to map kernel or feature map to its matrix form so that convolution can be conducted by matrix-vector multiplication. However, a straight forward kernel-matrix, image-vector
product representation of convolution is not applicable here, since the kernel matrix is a sparse block-Toeplitz-Toeplitz-block one, not suitable for
parallelization due to the existence of many zero elements. Thanks to the duality of kernel and feature map in convolution, we can construct a {\it dense} feature-map-matrix and a kernel-vector. Further, multiple kernels can be put together to form a matrix so as to generate multiple feature map outputs simultaneously,
\begin{eqnarray}
[f_i^{l+1}]_{i} = \sigma(\varphi_c(f^{l})[w_i^l]_i+[b_i^l]_i).
\end{eqnarray}
Operator $[~]_i$ is to assemble vectors with index $i$ to form a matrix.
\begin{figure}[t]
\centering
\includegraphics[width=1.0\linewidth]{figures/conv1.eps}
\caption{Strategy to vectorize convolution. Illustration of the way to covolve a 3x3 input image with three 2x2 kernels and generates three 2x2 feature maps.}
\label{fig:fig2}
\end{figure}
{\bf Backpropagation} The training procedure requires the backward propagation of gradients through $\varphi_c(f^{l})$. Note that $\varphi_c(f^{l})$ is in the unrolled matrix form, different form the outputs of previous layer $[f^{l}_i]_i$. An inverse operator $\varphi_c^{-1}()$ is thus required to transform the matrix-form gradients into the vector form for further propagation. Since $\varphi_c()$ is a one-to-many mapping, $\varphi_c^{-1}()$ is a many-to-one operator. Fortunately, the gradient update is a linear process which can also be processed separately and combined afterwards.
{\bf Matlab Practice} Our matlab implementation to vectorize the input image is shown in Fig. \ref{fig:fig2}. Specifically, we first crop the image patches base on the kernel size and reorganize
them into columns, as indicated by the dotted bounds. Convolution kernels are
arranged by rows in another matrix. We can see that the product of
these two matrices will put all the convolved feature maps in the
resulting matrix, one feature map per row. $\varphi_c()$ here can
be efficiently implemented by the \textit{im2col()} functions in
Matlab on both GPU \footnote{We used a custom version of
\textit{im2col()} for GPU.} and CPU. We note that the feature map vector here are transpose of $f_i$, simply because the function of \textit{im2col()}.
\paragraph{\it Convolution with the feature map } An alternative practice is needed to
handle convolution of the feature map (e.g.
``{\bf c}" in Fig. \ref{fig:fig1}). This is because we need to first combine $[f^{l}_i]_i$ to a higher dimensional $f^{l}$ and then perform convolution.
One example is illustrated in Fig. \ref{fig:fig3}, where we need to combine 3 $3\times3$ feature maps to a $3\times3\times3$ one and apply $\varphi_c()$. In practice, we could first apply three times $\varphi_c()$ to $f^{l}_i$ and then combine the results. We found it less efficient since the actual number of feature map is much larger. To exploit more parallelism, we try to reorganize the data and apply only once the vectorization operator $\varphi_c()$.
\begin{figure}[t]
\centering
\includegraphics[width=1.0\linewidth]{figures/conv2.eps}
\caption{Strategy to vectorize convolution with a feature map. Illustration of the way to convolve a 3x3x3 feature map.}
\label{fig:fig3}
\end{figure}
In Fig. \ref{fig:fig3}, we first reshape column vectors $[f_i]_i$ (a) back to a 2D matrix
and put them side by side (b). This operation is cheap
since it just changes the representation of the data but does not
rearrange it. It allows us to vectorize all the feature
maps in a holistic way (c). Since only valid region is considered during convolution, we set all the redundant columns to zeros. The final $\varphi_c(f)$ is obtained by rearranging the intermediate result in (c). Since redundant columns are involved, we use Matlab function \textit{accumarray()} to handle many-to-one rearrangement.
One may note that the operator $\varphi^{-1}_c()$ is a many-to-one mapping as well. So it can be efficiently implemented by \textit{accumarray()}, for backpropagation.
\subsection{Vectorizing Pooling}
It is inefficient to carry out the pooling separately for each
feature map, so the goal here is to simultaneously process those
separate operations by vectorization. The pooling operator can be abstracted as
\begin{eqnarray}
f^{l+1} = \sigma(\varphi_p(f^{l})+b^l),
\end{eqnarray}
where $\varphi_p()$ is a many-to-one mapping with a defined operation corresponding to max- or average- pooling.
\begin{figure}[t]
\centering
\includegraphics[width=1.0\linewidth]{figures/pool.eps}
\caption{Strategy to vectorize pooling. Illustration of pooling for one 4x4 feature map.}
\label{fig:fig4}
\end{figure}
Due to the information loss, the inverse operator $\varphi^{-1}_p()$ is not well defined. We simply use nearest neighbor upscaling for approximation during backpropagation.
The pooling operations, both average pooling and max pooling, can be
thought of as a vector accumulation process guided by a pre-defined
index map, which can be similarly implemented by
\textit{accumarray()}. The only difference for max pooling is a
\textit{max} function is involved. For overlapping pooling, we could
insert the overlapped elements into the feature map and then apply
the same pooling strategy.
\subsection{Vectorizing Fully Connected Layers}
The fully connected layers (i.e. ``{\bf d}" in Fig. \ref{fig:fig1}), can be written in a dense matrix-matrix multiplication. Thus both the feed forward and backpropagation are naturally vectorized, in a unified matrix-matrix form.
\subsection{Vectorization for Mini-batches}
Our vectorization strategy can be directly extended to support
mini-batch training. Given a batch of samples indexed by $j$, the mini-batch training with a convolution layer is given by
\begin{eqnarray}
[f_i^{l+1}]_{i} = \sigma([\varphi_c(f^{l}_{,j})]_j[w_i^l]_i+[b_i^l]_i),
\end{eqnarray}
where $[~]_j$ is to assemble the matrix of different samples.
Figure \ref{fig:fig5} shows the Matlab implementation of batch mode for the same operation as
in Fig. \ref{fig:fig2} with the batch size of 2. Both samples in the input batch are vectorized and the
outputs are arranged horizontally. We can show that the product of
the same kernel matrix as in Fig. \ref{fig:fig2} and this matrix is able to simultaneously generate feature maps for both
samples. Note that if an input sample of a convolutional layer has
multiple channels, we could treat it as a multi-channel feature map as shown in Fig. \ref{fig:fig3}.
\begin{figure}[t]
\centering
\includegraphics[width=1.0\linewidth]{figures/batch.eps}
\caption{Strategy to vectorize mini-batch operations.}
\label{fig:fig5}
\end{figure}
\section{Experiments and Analysis}
The goal of the experiments presented in this section is to understand
the role of vectorization in training and testing CNNs as well as its limitation.
In order to make our experiment results of high validity and
relevancy, we compared the training and testing speed of our fully vectorized
implementation with Caffe and Cudaconvnet, the speed is competitive,
if not faster in all the tested cases.
\subsection{Comparing Different Degrees of Vectorization}
In this section, we seek to understand the role of vectorization by comparing six CNN implementations. These implementations differ in the degree of vectorization, which is illustrated in table 1. Imp-1 is a least vectorized one, Imp-2 naively parallelize the process of batch samples by adding a parallel for-loop to Imp-1 whilst Imp-6 is a fully vectorized implementation guided by the approaches we introduced. When working with one particular implementation we also observe how results change with network scales. The reason is we would like to examine different vectorization strategies with both small scale network and large scale ones and we are particularly interested in large scale CNNs since it is more relevant to recent advances in the field.
We consider three scales. Scale 1 is a small model, it is very similar to the standard LeNet \cite{lecun98}, but with more feature maps in the convolutional layers\footnote{This setting of conv layer is the same in Caffe's MNIST example. We have 2 fully connected hidden layers.}, ReLU nonlinearity and cross-entropy error. Scale 2 is a large network with about 60 million trainable parameters which is comparable to AlexNet \cite{KrizhevskySH12}. However, the architecture is tailored for the purpose of this study. First, we would like to keep the number of conv layer and the number of fully connected layer balanced so that we shall have a fair performance breakdown. It also allows us to directly compare the results with Scale 1. Second, to enable a fair comparison, we would like to have a unified vectorization scheme for convolution throughout the network. Thus unlike AlexNet which uses a stride 4 convolution in the first conv layer and stride 1 thereafter, all convolution operations use the same stride of 1 in our model. The consequences of those are, compare to AlexNet, scale 2 tends to have more feature maps in the conv layer but smaller size for the input images. We set the number of output units to 1000 as in AlexNet. Scale 3 is a larger network with 10,000 output units. This pushes the number of trainable parameters to 94 million, keeping other settings the same as scale 2. The performance on GPU\footnote{GeForce GTX 780 Ti, same card was used in the rest of the experiments.} (in terms of the number of images to be processed per second) of the six CNNs with different network scales during training is illustrated in table 2.
\begin{table}[t]
\centering \small
\begin{tabular}{l*{6}{c}r}
Vec ele & Fu-co & Conv & Pool & Feat & Batch\\
\hline
Imp-6 & \checkmark & \checkmark & \checkmark & \checkmark & \checkmark \\
Imp-5 & \checkmark & \checkmark & \checkmark & \checkmark \\
Imp-4 & \checkmark & \checkmark & \checkmark \\
Imp-3 & \checkmark & \checkmark \\
Imp-2 & \checkmark & & & & \checkmark \\
Imp-1 & \checkmark
\end{tabular}
\caption{Different vectorized elements included in the six CNNs.
Fu-co: fully connected layer, Conv: convolutional layer, Pool:
pooling layer, Feat: feature map pachification, Batch: vectorize for
batch. A tick indicates the element is vectorized}
\label{tab:table1}
\end{table}
We are able to observe several insights from the figures. First of all, the results indicates that vectorization is vital to CNN's speed. We can see that Imp-1 is very slow and a naive parallelization (Imp-2) seems work poorly on GPU. Especially for the large scale networks, Imp-1 and Imp-2 are simply too slow to be practical. When training a small network, a fully vectorized CNN (Imp-6) is more than 200 times faster than the naive parallelization version during training and more than 100 times faster during testing. This acceleration is going to be more significant for bigger networks since Imp-1 and 2 scale poorly.
Second, all vectorization element we introduced contribute significantly to the final performance, during both training and testing. One interesting insight is the contribution of vectorizing pooling and feature map patchification seems to increase with the scale of the network. For instance, in table 2, Imp-4 (vectorize for pooling) has a 1.9x speed up than Imp-3 under scale 1 but a 4.5x speed up and a 4.3x speed up under scale 2 and scale 3 respectively. Same phenomenon happens for testing. This strongly indicates that the vectorization strategy for those two elements scales well with the size of the network.
\begin{table}[!htb]
\centering \small
\begin{tabular}{l*{6}{c}r}
\#img/sec & Imp-1 & Imp-2 & Imp-3 & Imp-4 & Imp-5 & Imp-6 \\
\hline
Scale 1 & 1 & 6.1 & 15.3 & 29.5 & 85.4 & 1312.1 \\
Scale 2 & n/a & n/a & 2.4 & 11 & 42.3 & 188.7 \\
Scale 3 & n/a & n/a & 2.3 & 10 & 34.3 & 161.2 \\
\multicolumn{7}{l}{%
\begin{minipage}{5.5cm}%
\scriptsize $*$ batch size is 100 for scale 1 and 200 for scale 2 and 3.%
\end{minipage}%
}\\
\end{tabular}
\caption{Training performance of the six CNNs (\#images to be
processed per second). Scale1: small model, 10 output units; Scale2:
large model, 1000 output units, Scale3: larger model, 10000 output
units.} \label{tab:table2}
\end{table}
On the other hand, we also observe that vectorizing batch processing brings more than 10x speed up for small models but only 3x to 5x speed up for large scale models. The contribution of vectorizing batch processing to the performance seems to decrease when scaling up the network though the speed up remains significant. We further investigate this phenomenon in the next section which leads to a strategy to achieve optimal training and testing speed.
\subsection{In Search of Optimal Speed}
We investigated the puzzle of decelerating speed up by scrutinizing the performance against different batch sizes. The results are presented in table 3 and 4 for training and testing respectively.
\begin{table}[!htb]
\centering
\small
\begin{tabular}{l*{6}{c}r}
\#img/sec & b=1 & b=100 & b=200 & b=300 & b=400 \\
\hline
Scale 1 & 88.5 & 1312 & 1450.9 & 1574.2 & 1632.8 \\
Scale 2 & 41.9 & 136.9 & 188.7 & 192.3 & 106.3 \\
Scale 3 & 34.3 & 123.5 & 161.3 & 163.9 & 91 \\
\end{tabular}
\caption{Training performance of Imp-6 against different batch sizes (\#images to be processed per second).}
\label{tab:table4}
\end{table}
\begin{table}[!htb]
\centering
\small
\begin{tabular}{l*{6}{c}r}
\#img/sec & b=1 & b=100 & b=200 & b=400 & b=600 \\
\hline
Scale 1 & 151.5 & 1812.6 & 1878.4 & 2023.5 & 2192.2 \\
Scale 2 & 75.8 & 222.2 & 270.2 & 285.7 & 103.1 \\
Scale 3 & 74 & 212.8 & 256.4 & 277.8 & 89.2 \\
\end{tabular}
\caption{Test performance of Imp-6 against different batch sizes (\#images to be processed per second).}
\label{tab:table5}
\end{table}
In table 3, we can see that for the small model (scale 1) the acceleration brought by each adjacent batch size increase is 14x, 1.1x, 1.08x and 1.03x. The acceleration obtained via the increase of batch size seems to be rapidly vanishing. For the large model (scale 2), the first three acceleration ratio are 3.2x, 1.3x and 1.02x, demonstrating the same vanishing trend. Further increase in batch size even leads to a performance degradation instead. Same situation occurs for the larger model (scale 3). Though the ability of processing 192 images/second for training and 285 images/second for testing with our commodity GPU for the scale 2 network is promising, this result still indicates that there is some scaling limitation within the vectorization for batch processing. Similar results in table 4 seems to further suggest that such limitation is shared between training and testing. In order to completely understand the rationale under the hood, we have to resort to a detailed performance breakdown.
\subsubsection{Performance Breakdown and Limitation. }
We decompose the whole training procedure into the following components. They are 1) conv layers; 2) pooling layers; 3) fully connected layers; 4) others (e.g. ReLU, cost). We distinguish the statistics between forward pass and back-propagation, therefore 8 components to look at.
\begin{figure}[t]
\centering
\begin{tabular}{@{\hspace{0mm}}c@{\hspace{1mm}}c@{\hspace{1mm}}c@{\hspace{1mm}}c@{\hspace{1mm}}c}
\includegraphics[width=0.35\linewidth]{figures/a.eps}&
\includegraphics[width=0.35\linewidth]{figures/b.eps}&
\includegraphics[width=0.35\linewidth]{figures/c.eps}
\\
(a) & (b) & (c)
\end{tabular}
\caption{Performance break down. (a) Scale 3 network, batch size = 1. (b) Scale 3 network, batch size = 200. (c) Scale 3 network, batch size = 300. \textit{conv}: conv layers, \textit{pool}: pooling layers, \textit{full}: fully connected layers, \textit{other}: other operations, \textit{\_f}: forward pass, \textit{\_b}: back-propagation. }
\label{fig:fig6}
\end{figure}
Figure 6 illustrates the performance break down (in terms of the
proportion of computing time in processing one batch) during
training of the two representative cases from our largest network
(scale 3) in the experiment. Batch size is 1 for Fig. 6(a), 200 for
Fig. 6(b) and 300 for Fig. 6(c). We can observe from Fig. 6(a) that
44\% of the overall time was used in processing the fully connected
layers. It was in fact the biggest consumer of the computing time
for this batch size. We also see that the time spent on full\_b is
significantly more than full\_f. This makes sense because it
involves larger matrix multiplication and larger transform matrix
than that in the forward pass. The second largest consumer of time
is the convolution layers and we can see that the time spent in
forward pass and back-propagation is reasonably balanced.
However, the situation we found in Fig. 6(b) and Fig. 6(c) is very
different. One obvious character is, when increasing the batch size,
the time costs by the conv layers is now considerably more than the
fully connected layers. While the proportion between full\_f and
full\_b among the three batch sizes roughly remains the same, we
found conv\_f spent much more time than conv\_b for large batch
sizes. This indicates the scaling limitation is within the conv\_f
when vectorizing for batch processing. A further scrutiny on this
issue shows that the limitation is caused by the following two
factors namely the memory overhead in handling multiple samples and
the overhead caused by invoking patchification on bigger samples.
While there might be alternative strategies to vectorize batch
processing, we argue that the aforementioned overhead is hard to be
completely avoided.
\subsubsection{Finding the Optimal Speed. }
We found the observations from Fig. 6 are also valid for scale 1 and
scale 2 networks, but with an important difference. For small
networks like the scale 1 network, the acceleration brought by batch
processing shall be valid for very big batch sizes (e.g. 1000)
whilst for large networks batch size needs to be chosen carefully or
else the speed degradation like we saw in table 3 and 4 shall occur
before the network hits the GPU memory ceiling. This suggests that
given a network design choosing an appropriate batch size may be
vital in achieving the optimal speed. Based on our scale 2 network,
we select 10 other networks by randomly adjusting several parameters
such as filter size, number of feature maps, number of output units
and sigmoid function, etc. We run these networks for both training
and testing by adjusting the batch sizes to see if this contention
is generally applicable for large networks.
\begin{figure}[t]
\centering
\begin{tabular}{@{\hspace{0mm}}c@{\hspace{1mm}}c@{\hspace{1mm}}c@{\hspace{1mm}}c@{\hspace{1mm}}c}
\includegraphics[width=0.5\linewidth]{figures/line_chart1.eps}&
\includegraphics[width=0.5\linewidth]{figures/line_chart2.eps}\\
(a) & (b)
\end{tabular}
\caption{Speed of 10 randomly selected networks. X axis, batch size. Y axis, number of images to be processed per second. (a) for training. (b) for testing.}
\label{fig:fig7}
\end{figure}
Figure 7 confirms our aforementioned contention for large
networks and makes the importance of choosing an appropriate batch size
obvious. First, it suggests that the optimal batch size among
different network parameters is usually quite different. Directly
adopting a batch size from a previous set of network parameters may
lead to significantly inferior speed. Second, it also suggests
that the optimal batch size between the training stage and the
testing stage is also different, even if for the same network. A
naive adoption of the batch size from the training stage is often
not optimal and leads to considerable speed loss. These findings has
direct implications in building real-time systems in which
optimization for model testing is the key.
\section{Unification of High{/}Low Level Vision Tasks }
Despite the rapid adoption of deep CNN in addressing various kinds of high level computer vision tasks typified by image classification and object localization, other problems such as detecting objects of different shapes in real-time seem still a problem under investigation. On the other hand, we observed that there are a few very recent studies \cite{xu14,Eigen14} successfully used deep CNN in various low level vision tasks such as image deblurring and denoising, etc. Though the domain knowledge required to build those new networks substantially differ from that used in addressing high level vision tasks, same vectorization principles presented in this paper will apply.
More interestingly, the same vectorization principle across those tasks actually gives us a chance (perhaps for the first time) to unify both high level vision tasks and low level vision tasks in a single computational framework. In this section, we introduce the application of our VCNN implementation in tasks seemingly of distinct fields namely, image denoising and deblurring (low level vision) as well as multi-object detection (high level vision).
\subsection{CNN for Image Processing }
Image processing tasks do not require pooling and fully connected layers in general. To verify the effectiveness of the proposed vectorized framework, we implemented a network architecture by simply removing the pooling and fully connected layers from Fig. \ref{fig:fig1} and trained the network with synthesized clear-noisy image pairs. One of the denoise result is given in Fig. \ref{fig:fig8}. Another sample application of our vectorized CNN is the recent proposed image deconvolution \cite{xu14}. Result is shown in Fig. \ref{fig:fig9}.
\begin{figure}[!htb]
\centering
\includegraphics[width=1.0\linewidth]{figures/denoise.eps}
\caption{Application in image denoising. }
\label{fig:fig8}
\end{figure}
\begin{figure}[!htb]
\centering
\includegraphics[width=1.0\linewidth]{figures/deblur.eps}
\caption{Application in image deconvolution. }
\label{fig:fig9}
\end{figure}
\subsection{Novel Training Scheme for Multi-object Detection }
Conventional image classifiers are usually trained by image samples with equal sizes. This imposes a critical limitation when applying it in detection. For instance, it is reasonable to put a human face sample in a square image, but doing so for non-squared objects (e.g. shoes) tends to include more background content thus introduces more noise which is detrimental to accurate and efficient object detection. One possible alternative is to formulate object detection as a regression problem \cite{szegedy13}, however, it requires a very large amount of data and usually very big models to capture the variety of the possible patterns.
\begin{figure}[!htb]
\centering
\includegraphics[width=1.0\linewidth]{figures/detection.eps}
\caption{Application in real time multi-object detection. Shoes review videos are from Youtube. }
\label{fig:fig10}
\end{figure}
Using VCNN, we were able to train a single image classifier but with heterogeneous input sizes by using vectorization. The key insight is heterogeneous inputs can actually share all the weights in a CNN except the ones in the connection between conv layer and fully connected layer. This approach not only avoids the background noise but also being a lot more lightweight than the regression approach. We successfully applied it in a detection system which runs in real-time. We can show that this approach tends to have less false alarms and works efficiently with multi-scale detection through vectorization.
\section{Conclusion}
In this paper, we elaborate several aspects on vectorization of
deep CNN. First, we present the vectorization steps of all essential parts of implementing deep CNNs. The vectorization steps are further exemplified by Matlab practices. Second, we have developed and compared six CNN implementations
with different degrees of vectorization to analysis the impact
of vectorization on speed. Third, based on the practices, we provide a unified framework for handling both low-level and high-level vision tasks. Experiments on various applications including image denoise, decovolution and real-time object detection demonstrated the effectiveness of the proposed strategies. As
the introduced vectorization techniques are general enough, our future
direction includes optimization for different hardware or cloud platforms.
\bibliographystyle{aaai}
|
\section{Introduction}
Compact-object binaries are amongst the most compelling sources of gravitational waves. In
particular, the ubiquity of supermassive black holes residing in galactic centres
\cite{Magorrian:1997hw} has made the extreme mass ratio regime a prime target for the eLISA mission
\cite{Gair:2004iv,AmaroSeoane:2007aw,Gair:2008bx,eLISA:GWNotes,eLISA}. Meanwhile, the comparable and
intermediate mass ratio regimes are an intriguing target for study by the imminent Advanced LIGO
detector \cite{LIGO}. In order to maximise the scientific gain realised from gravitational-wave
observations, highly accurate models of gravitational-wave sources are essential.
For the case of extreme mass ratio inspirals (EMRIs) --- binary systems in which a compact, solar mass
object inspirals into an approximately million solar mass black hole ---
the demands of gravitational wave astronomy are particularly stringent; the
promise of groundbreaking scientific advances --- including precision tests
of general relativity in the strong-field regime
\cite{Barack:2003fp,Babak:2010ej,Gair:2010bx} and a better census of
black hole populations —-- hinges on our ability to track the phase of their
gravitational waveforms throughout the long inspiral, with an accuracy of better
than 1 part in 10,000 \cite{Gair:2004iv}. This, in turn requires highly
accurate, long time models of the orbital motion.
For the past two decades, these demands have stimulated an intense period of EMRI
research among the gravitational physics community. Despite of the impressive progress made by
numerical relativists towards tackling the two-body problem in general relativity
(for reviews see Refs.~\cite{Hinder:2010vn,Pfeiffer:2012pc,Sperhake:2014wpa}), the
disparity of length scales characterising the EMRI regime is a significant roadblock
for existing numerical relativity techniques. Indeed, to this day EMRIs remain intractable by current
numerical relativity methods, and successful approaches have instead tackled the problem perturbatively
or through post-Newtonian approximations (see Ref.~\cite{Tiec:2014lba} for a review). This article will focus on the first of these;
by treating the smaller object as a perturbation to the larger mass, the so-called ``self-force
approach'' reviewed here has been a resoundingly successful tool for EMRI research.
Within the self-force approach, the smaller mass, $\mu<<M$, is assumed to be sufficiently small that it
may be used as a perturbative expansion parameter in the background of the larger mass, $M$.
Expanding the Einstein equation in $\mu$, we see
the smaller object as an effective point particle generating a perturbation about the background of
the larger mass. At zeroth order in $\mu$, the smaller object merely follows a geodesic of the
background. At first order in $\mu$, it deviates from this geodesic due to its interaction with its
self-field. Viewing this deviation as a force acting on the smaller object, the calculation of this
self-force is critical to the accurate modelling of the evolution of the system. For
the purposes of producing an accurate waveform for space-based detectors, and for producing
accurate intermediate mass ratio inspiral (IMRI) models, it will be necessary to include further effects up to
second perturbative order \cite{Isoyama:2012bx,Burko:2013cca}. Indeed, recent compelling
work \cite{LeTiec:2011bk,Tiec:2013twa} suggests that IMRIs —-- and even comparable mass binaries —-- may be modelled using self-force
techniques.
A na\"ive calculation of the first order perturbation due to a point particle leads to a retarded
field which diverges at the location of the particle. The self-force, being the derivative of the
field, also diverges at the location of the particle and one obtains equations of motion which are not
well-defined and must be regularized. A series of formal derivations of the regularized first order
equations of motion (now commonly referred to as the MiSaTaQuWa equations, named after Mino,
Sasaki, Tanaka \cite{Mino:1996nk} and Quinn and Wald \cite{Quinn:1996am} who first derived them)
for a point particle in curved spacetime have been developed
\cite{Dirac:1938nz,DeWitt:1960fc,Hobbs:1968a,Mino:1996nk,Quinn:1996am,Quinn:2000wa,Detweiler:2002mi,
Harte:2008xq,Harte:2009uy,Harte:2009yr,Galley:2010xn,Galley:2011te}, culminating in a rigorous work by Gralla and Wald
\cite{Gralla:2008fg} and Pound \cite{Pound:2009sm} in the gravitational case and by Gralla, et al.
\cite{Gralla:2009md} in the electromagnetic case. This was subsequently extended to second
perturbative order by Rosenthal \cite{Rosenthal:2005ju,Rosenthal:2005it,Rosenthal:2006nh,Rosenthal:2006iy}, Pound \cite{Pound:2012nt,Pound:2012dk,Pound:2014xva}, Gralla
\cite{Gralla:2012db} and Detweiler \cite{Detweiler:2011tt}. These derivations eliminate the
ambiguities associated with the divergent self-field of a point particle and provide a
well-defined, finite equation of motion. Building upon this foundational progress, several
practical computational strategies have emerged from these formal derivations:
\begin{itemize}
\item \emph{Dissipative self-force approaches:} While the full first-order self-force
is divergent, it turns out that the dissipative component is finite and requires no
regularization. This fact has prompted the development of methods for computing the
dissipative component alone, sidestepping the issue of regularization altogether.
These dissipative approaches fall into two categories:
\begin{enumerate}
\item \emph{Flux methods:} By measuring the orbit-averaged flux of gravitational waves onto the horizon of the
larger black hole and out to infinity, the fact that the field is evaluated far away from the worldline
means that no divergent quantities are ever encountered. This approach yields the time-averaged\footnote{For
the case of inclined orbits in Kerr spacetime, this is more appropriately formulated as a torus-average.} dissipative
component of the self-force \cite{Mino:2003yg,Glampedakis:2002ya,Hughes:2005qb,Sago:2005fn,Drasco:2005is,Drasco:2005kz,Sundararajan:2007jg,Fujita:2009us}.
\item \emph{Local/instantaneous dissipative self-force:}
The time averaging element of flux methods can be eliminated by instead computing the local, instantaneous
dissipative component of the self-force from the half-advanced-minus-half-retarded field
\cite{Mino:2003yg,Gralla:2005et,Hinderer:2008dm,Flanagan:2012kg}.
\end{enumerate}
Both methods, however, fundamentally rely on neglecting potentially important
conservative effects which can significantly alter the orbital phase of the
system.
\item The \emph{mode-sum} approach: Introduced in Refs.~\cite{Barack:1999wf,Barack:2001gx}, and
having since been successfully used in many applications, the approach relies on the decomposition
of the retarded field into spherical harmonic modes (which are finite, but not differentiable at
the particle), numerically solving for each mode independently and subtracting
analytically-derived ``regularization parameters'', then summing over modes.
\item The \emph{effective source} approach: Proposed in \cite{Barack:2007jh} and
\cite{Vega:2007mc}, the approach implements the regularization before solving the wave equation.
This has the advantage that all quantities are finite throughout the calculation and one can directly
solve a wave equation for the regularized field.
\item The \emph{worldline convolution} approach: First suggested in
\cite{Poisson:Wiseman:1998,Anderson:2005gb}, one computes the regularized retarded field as a
convolution of the retarded Green function along the past worldline of the particle. Although the
approach is the most closely related to the early formal derivations, it is only recently that it
has been successfully applied to calculations in black hole spacetimes.
\end{itemize}
For a comprehensive review of the self-force problem, see
Refs.~\cite{Poisson:2011nh,Detweiler:2005kq,Barack:2009ux,Blanchet:2011zz}. In this paper, I will review the various
approaches and give details of how each one is implemented in practice, highlighting the advantages
and disadvantages in each case.
This paper follows the conventions of Misner, Thorne and Wheeler \cite{Misner:1974qy}; a ``mostly
positive'' metric signature, $(-,+,+,+)$, is used for the spacetime metric, the connection
coefficients are defined by
$\Gamma^{\lambda}_{\mu\nu}=\frac{1}{2}g^{\lambda\sigma}(g_{\sigma\mu,\nu}
+g_{\sigma\nu,\mu}-g_{\mu\nu,\sigma}$), the Riemann tensor is
$R^{\alpha}{}_{\!\lambda\mu\nu}=\Gamma^{\alpha}_{\lambda\nu,\mu}
-\Gamma^{\alpha}_{\lambda\mu,\nu}+\Gamma^{\alpha}_{\sigma\mu}\Gamma^{\sigma}_{\lambda\nu}
-\Gamma^{\alpha}_{\sigma\nu}\Gamma^{\sigma}_{\lambda\mu}$, the Ricci tensor and scalar are
$R_{\alpha\beta}=R^{\mu}{}_{\!\alpha\mu\beta}$ and $R=R_{\alpha}{}^{\!\alpha}$, and the Einstein
equations are $G_{\alpha\beta}=R_{\alpha\beta}-\frac{1}{2}g_{\alpha\beta}R=8\pi T_{\alpha\beta}$.
Standard geometrized units are used, with $c=G=1$. Greek indices are used for four-dimensional
spacetime components, symmetrisation of indices is denoted using parenthesis [e.g. $(\alpha
\beta)$], anti-symmetrisation is denoted using square brackets (e.g. $[\alpha \beta]$) and indices
are excluded from symmetrisation by surrounding them by vertical bars [e.g. $(\alpha | \beta |
\gamma)$]. Latin letters starting from $i$ are used for indices summed only over spatial dimensions
and capital letters are used to denote the spinorial/tensorial indices appropriate to the field
being considered. Either $x$ or $x^\mu$ are used when referring to a spacetime field point and
$z(\tau)$ or $z^\mu(\tau)$ are used when referring to a point on a worldline parametrised by proper
time $\tau$. Finally, a retarded (or source) point is denoted using a prime, i.e. $z'$.
\section{Equations of Motion}
The formal equations of motion of a compact object moving in a curved spacetime are now well
established up to second perturbative order. Writing the perturbed spacetime in terms of a
background plus perturbation, $g_{\alpha\beta} = g^{(0)}_{\alpha \beta} + h_{\alpha \beta}$, the
equations of motion essentially amount to those of an accelerated worldline in the background spacetime,
with the acceleration given by a well-defined regular field which is sourced by the worldline.
To order $\mu$, this
coupled system of equations for the worldline and its self-field are commonly referred to as the
MiSaTaQuWa equations and are given (in Lorenz gauge, assuming a Ricci-flat background spacetime) by
\begin{subequations}
\begin{align}
\label{eq:field-gravity}
\Box \bar{h}^{\rm ret}_{\alpha\beta} + 2 C_{\alpha}{}^{\gamma}{}_{\beta}{}^{\delta } \bar{h}^{\rm ret}_{\gamma\delta} &= - 16 \pi \mu \int g_{\alpha'(\alpha} u^{\alpha'} g_{\beta)\beta'} u^{\beta'} \delta_4(x, z(\tau')) d \tau' \\
\label{eq:accel-gravity}
\mu \, a^\alpha &= \mu\, k^{\alpha\beta\gamma\delta} \bar{h}^{\rm R}_{\beta\gamma;\delta}
\end{align}
\end{subequations}
with
\begin{equation*}
k^{\alpha\beta\gamma\delta} = \frac12 g_{(0)}^{\alpha\delta} u^\beta u^\gamma - g_{(0)}^{\alpha\beta} u^\gamma u^\delta -
\frac12 u^\alpha u^\beta u^\gamma u^\delta + \frac14 u^\alpha g_{(0)}^{\beta \gamma} u^\delta +
\frac14 g_{(0)}^{\alpha \delta} g_{(0)}^{\beta \gamma}.
\end{equation*}
Here, $\mu$ is mass of the object, $g_{\alpha'\alpha}$ is the bivector of parallel transport,
$C_{\alpha\beta\gamma\delta}$ is the Weyl tensor of the background
spacetime, and we use the trace-reversed metric perturbation
$\bar{h}_{\alpha\beta} = h_{\alpha\beta} - \frac{1}{2} g^{(0)}_{\alpha\beta} h$, $h = h^\gamma{}_{\gamma}$.
One can also consider compact objects possessing other types of charge. For example,
a particularly simple case is that of a scalar charge $q$ with mass $m$ and scalar field $\Phi$,
in which case the equations of motion are given by\footnote{In the scalar case, it is important to distinguish
between the self-force $F^\alpha = \nabla^\alpha \Phi$ and the self-acceleration, which is given
by projecting to self-force orthogonal to the worldline,
$a^\alpha = (g_{(0)}^{\alpha \beta} + u^{\alpha} u^{\beta}) F_{\beta}$. In the electromagnetic and
gravitational cases the self-force has no component along the worldline and the two may be used
interchangeably.}${}^{,}$\footnote{We assume that the mass $m$ is small and ignore its effect on the equations
of motion.}
\begin{subequations}
\begin{align}
\label{eq:field-scalar}
(\Box - \xi R) \Phi^{\rm ret} &= - 4 \pi q \int \delta_4(x, z(\tau')) d \tau' \\
\label{eq:accel-scalar}
m a^\alpha &= q \big(g_{(0)}^{\alpha\beta}+u^\alpha u^\beta\big) \Phi^{\rm R}_{,\beta} \\
\label{eq:mdot-scalar}
\frac{dm}{d\tau} &= - q \, u^\alpha \Phi^{\rm R}_{,\alpha}.
\end{align}
\end{subequations}
Here, $R$ is the Ricci scalar of the background spacetime and $\xi$ is the coupling to scalar curvature.
Similarly, for an electric charge, $e$, one obtains equations of motion which
are given in Lorenz gauge by
\begin{subequations}
\begin{align}
\label{eq:field-em}
\Box A^{\rm ret}_\alpha - R_\alpha^\beta A^{\rm ret}_\beta &= - 4 \pi e \int g_{\alpha\alpha'} u^{\alpha'} \delta_4(x, z(\tau')) d \tau' \\
\label{eq:accel-em}
m a^\alpha &= e \big(g_{(0)}^{\alpha\beta}+u^\alpha u^\beta\big) A^{\rm R}_{[\gamma, \beta]} u^\gamma,
\end{align}
\end{subequations}
where $R_{\alpha\beta}$ is the Ricci tensor of the background spacetime and $A^\mu$ is the vector potential.
The key component in all instances
is the identification of the appropriate regularized field on the worldline. Detweiler and Whiting
identified a particularly elegant choice for the regularized field, written in terms of the
difference between the retarded field and a locally-defined singular field,
\begin{equation}
\label{eq:singular-regular-split}
\Phi^{\rm R} = \Phi^{\rm ret} - \Phi^{\rm S}, \quad
A^{\rm R}_\alpha = A^{\rm ret}_\alpha - A^{\rm S}_\alpha, \quad
h^{\rm R}_{\alpha\beta} = h^{\rm ret}_{\alpha\beta} - h^{\rm S}_{\alpha\beta}.
\end{equation}
In addition to giving the physically-correct self-force, the Detweiler-Whiting regular field has the
appealing feature of being a solution of the homogeneous field equations in the vicinity of the
worldline. Most computational strategies essentially amount to differing ways of representing this
singular field\footnote{Some methods \cite{Lousto:2008mb,Kol:2013tfa} rely on alternative
prescriptions for the singular field than that proposed by Detweiler and Whiting.}
and obtaining the regularized field on the worldline.
\section{Numerical regularization strategies}
In a numerical implementation, it is essential to avoid the evaluation of divergent quantities. In
the case of self-force calculations, both the retarded and singular fields diverge on the worldline
so one must avoid evaluating them there. Several strategies for doing so have emerged over the
years (see Table \ref{table:methods} for a summary).
One option is to only ever evaluate finite, dissipative quantities (e.g. the retarded
field far from the worldline or the half-advanced-minus-half-retarded field on
the worldline).
This is the basis of the dissipative methods mentioned in the
introduction. Since these methods effectively avoid the problem of
regularization, they will not be discussed further here; we will return to them
in Sec.~\ref{sec:evolution}. It is worth noting, however, that the methods
typically used by dissipative calculations are essentially the same as those
used by mode-sum regularization for computing the retarded field, but without
the additional regularization step.
This leaves three regularization strategies which allow the regularized field to be computed on the
worldline without encountering numerical divergences: worldline convolution, mode-sum
regularization, and the effective source approach.
\begin{table}
\begin{tabular}{|c|m{2mm}||m{3.5cm}|m{3.5cm}|m{3.5cm}|}
\hline
\multicolumn{2}{|c||}{Case} & Worldline & Mode-sum & Effective Source \\
\hline
\parbox[t]{2mm}{\multirow{2}{*}{\rotatebox[origin=c]{90}{Scalar\qquad \,}}}
&
\rotatebox[origin=c]{90}{\,Schwarzschild\,}
&
circular (apprx)~\cite{Anderson:2005gb};\hfill\break
generic\hfill\break(quasilocal)
\cite{Casals:2009xa,Ottewill:2007mz};\hfill\break
generic~\cite{Casals:2013mpa,Wardell:2014kea,Zenginoglu:2012xe};\hfill\break
static~\cite{Casals:2012qq};\hfill\break
accelerated~\cite{Ottewill:2008uu};
&
radial~\cite{Barack:2000zq};\hfill\break
circular~\cite{Burko:2000xx,Detweiler:2002gi,DiazRivera:2004ik,Canizares:2009ay};\hfill\break
eccentric~\cite{Haas:2007kz,Haas:2006ne,Canizares:2010yx,Heffernan:2012su,Thornburg:2010tq};\hfill\break
static~\cite{Casals:2012qq};
&
circular~\cite{Barack:2007jh,Vega:2007mc,Dolan:2010mt,Lousto:2008mb,Vega:2009qb,Warburton:2013lea};\hfill\break
eccentric~\cite{Vega:2013wxa};\hfill\break
evolving~\cite{Diener:2011cc};
\\
\cline{2-5}
&
\rotatebox[origin=c]{90}{Kerr}
&
generic~\cite{Ottewill:2007mz};\hfill\break
accelerated~\cite{Ottewill:2008uu};
&
circular~\cite{Warburton:2010eq};\hfill\break
equatorial~\cite{Heffernan:2012vj,Warburton:2011hp};\hfill\break
inclined circular~\cite{Warburton:2014bya};\hfill\break
accelerated~\cite{Linz:2014pka};\hfill\break
static~\cite{Burko:2001kr,Ottewill:2012aj};
&
circular~\cite{Dolan:2011dx};\hfill\break
eccentric~\cite{Thornburg:Capra17};
\\
\hline
\parbox[t]{2mm}{\multirow{2}{*}{\rotatebox[origin=c]{90}{EM\qquad\,}}}
&
\rotatebox[origin=c]{90}{\,Schwarzschild\,}
&
static~\cite{Casals:2012qq};
&
static~\cite{Casals:2012qq};\hfill\break
eccentric~\cite{Haas:2011np,Heffernan:2012su};\hfill\break
static~(Schwarzschild-de~Sitter)~\cite{Kuchar:2013bla};\hfill\break
radial~(Reissner-\hfill\break
Nordstr\"om)~\cite{Zimmerman:2012zu};
&
---
\\
\cline{2-5}
&
\rotatebox[origin=c]{90}{\,Kerr\,}
&
---
&
equatorial~\cite{Heffernan:2012vj};\hfill\break
accelerated~\cite{Linz:2014pka};
&
---
\\
\hline
\parbox[t]{2mm}{\multirow{2}{*}{\rotatebox[origin=c]{90}{Gravity\quad\,}}} & \rotatebox[origin=c]{90}{\,Schwarzschild\,}
&
generic\hfill\break
(quasilocal)~\cite{Anderson:2004eg};
&
radial~\cite{Barack:2002ku};\hfill\break
circular~\cite{Barack:2005nr,Dolan:2013roa,Dolan:2014pja,Field:2010xn,Keidl:2010pm,Merlin:2014qda,Sago:2008id,Shah:2010bi,Akcay:2010dx};\hfill\break
eccentric~\cite{Heffernan:2012su,Barack:2008ms,Barack:2010tm,Sago:2009zz,Field:2009kk,Hopper:2010uv,Hopper:2012ty,Akcay:2013wfa,Pound:2013faa,Osburn:2014hoa};\hfill\break
osculating~\cite{Warburton:2011fk};
&
circular~\cite{Dolan:2012jg};
\\
\cline{2-5}
&
\rotatebox[origin=c]{90}{\,Kerr\,}
&
circular\hfill\break
(quasilocal)~\cite{Anderson:2005gb};\hfill\break
branch cut~\cite{Kavanagh:Capra17};
&
equatorial~\cite{Heffernan:2012vj};\hfill\break
accelerated~\cite{Linz:2014pka};\hfill\break
circular~\cite{Pound:2013faa,Shah:2012gu};
&
circular~\cite{Dolan:Capra16};\hfill\break
generic~\cite{Wardell:2011gb};
\\
\hline
\end{tabular}
\caption{Summary of regularization methods employed by self-force calculations in black hole spacetimes.}
\label{table:methods}
\end{table}
\subsection{Worldline convolution}
The worldline convolution method relies on a split of the regularized self-force into an
``instantaneous'' piece and a history-dependent term. The instantaneous piece is easily calculated
from local quantities evaluated at the particle's position,
\begin{subequations}
\label{eq:inst}
\begin{align}
\Phi_{,\alpha}^{\text{inst}} &= q \bigg[\frac{1}{2}\Big(\xi - \frac{1}{6}\Big)R u_\alpha + \Big(g^{(0)}_{\alpha\beta} + u_\alpha u_\beta\Big) \Big(\frac{1}{3} \dot{a}^\beta + \frac{1}{6} R^{\beta}{}_{\gamma} u^\gamma\Big)\bigg], \\
A_{[\beta;\alpha]}^{\text{inst}} u^\beta &= e \Big(g^{(0)}_{\alpha\beta} + u_\alpha u_\beta\Big) \Big(\frac{1}{3} \dot{a}^\beta + \frac{1}{6} R^{\beta}{}_{\gamma} u^\gamma\Big), \\
\bar{h}_{\alpha\beta;\gamma}^{\text{inst}} u^\beta u^\gamma & = \bar{h}_{\alpha\beta;\gamma}^{\text{inst}} u^\alpha u^\beta = 0.
\end{align}
\end{subequations}
The history-dependent term is much more difficult to calculate, as it is given in terms of
a convolution of the derivative of the retarded Green function along the worldline's entire
past-history (see Fig.~\ref{fig:convolution}),
\begin{subequations}
\label{eq:hist}
\begin{align}
\Phi_{,\alpha}^{\text{hist}} &= q \int_{-\infty}^{\tau^-} \nabla_\alpha G^{\rm ret}[x, z(\tau')] \,d\tau',\\
A_{\alpha;\beta}^{\text{hist}} &= e \int_{-\infty}^{\tau^-} \nabla_\beta G^{\rm ret}_{\alpha\alpha'}[x, z(\tau')] u^{\alpha'}\,d\tau',\\
\bar{h}_{\alpha\beta;\gamma}^{\text{hist}} &= 4\mu \int_{-\infty}^{\tau^-} \nabla_\gamma G^{\rm ret}_{\alpha\beta\alpha'\beta'}[x, z(\tau')] u^{\alpha'} u^{\beta'} \,d\tau'.
\end{align}
\end{subequations}
The covariant derivatives here are taken with respect to the first argument of the retarded Green
function. Likewise, the regularized self-field can be obtained from a worldline convolution of the
retarded Green function itself; similar formulae can also be derived for higher derivatives.
\begin{figure}[htb!]
\begin{center}
\includegraphics[width=6cm]{figs/matchedexpansion}
\end{center}
\caption{
Schematic representation of the worldline convolution method. The equations shown are for the case
of the self-force on a scalar charge, but are representative of similar equations in the
electromagnetic and gravitational cases. Reproduced from Ref.~\cite{Casals:2013mpa}.
}
\label{fig:convolution}
\end{figure}
The retarded Green function appearing in these equations is a solution of the wave equation with a
delta-function source,
\begin{subequations}
\label{eq:BoxG}
\begin{align}
\label{eq:BoxG-scalar}
(\Box - \xi R) G^{\rm ret} &= - 4 \pi \, \delta^4 (x,x'), \\
\Box G^{\rm ret}_{\alpha\alpha'}
- R_\alpha{}^\beta G^{\rm ret}_{\beta\alpha'} &=
- 4 \pi \, g_{\alpha\alpha'} \delta^4 (x,x'), \\
\Box G^{\rm ret}_{\alpha \beta\alpha' \beta'}
+ 2 R_\alpha{}^\gamma{}_\beta{}^\delta G^{\rm ret}_{\gamma \delta\alpha' \beta'} &=
- 4 \pi \, g_{\alpha\alpha'} g_{\beta\beta'} \delta^4 (x,x'),
\end{align}
\end{subequations}
with boundary conditions such that the solutions correspond to purely outgoing radiation at infinity and no radiation emerging from the horizon.
The regularization (i.e. subtraction of the Detweiler-Whiting singular field) is formally achieved
by the limiting procedure in the upper limit of integration, i.e. by cutting off the integration at
$\tau^-$, slightly before the coincidence point $z = x$. In practice, this is done by taking the
integration all the way up to coincidence, but excluding the \emph{direct} contribution to the
Green function at coincidence (i.e. the term proportional to $\delta(\sigma)$ in the Hadamard form for
the Green function). Although the retarded Green function also diverges at certain other points
along the past worldline (in particular at null-geodesic intersections where the particle sees null
rays emitted from its past), it turns out that these are all integrable singularities (of the form
$1/\sigma$ and $\delta(\sigma)$) and the integral may be accurately evaluated to give a finite and
accurate value for the regularized self-force.
Despite having been proposed in the early days of EMRI self-force calculations
\cite{Poisson:Wiseman:1998}, the worldline convolution method was largely ignored for a long time by
numerical implementations (the notable exception being investigative studies
\cite{Anderson:2005gb,Capon:1998} which did not complete a full calculation of the
self-force). A likely reason is that the method relies on knowledge of the retarded Green function
for all points on the particle's past worldline. While methods have existed for decades for
computing portions of the retarded Green function, it turns out to be very difficult to obtain it
accurately everywhere that it is needed for the worldline convolution.
Thankfully, recent progress has led to two practical computational strategies, both of which have
been successfully applied to compute the self-force in black hole spacetimes. The first of these is
based on a frequency-domain decomposition of the Green function, and builds on the rich history of
black hole perturbation theory developed over the past several decades. The second, a time-domain
approach, has been a very recent development and shows a great deal of promise for the future.
Independently of whether a frequency-domain or a time-domain scheme is used, a common problem with
both is that they fail at early times when source and field points are close together; in the
frequency-domain case the convergence is poor at early times, while in the time-domain case
features from the numerical approximation pollute the data at early times. A relatively
straightforward solution which has been successfully applied in both scenarios is to only rely on
their results at late times and to supplement them at early times with a quasilocal Taylor series
expansion. Provided a sufficiently early time can be chosen where both the distant past and
quasilocal calculations converge, this yields a global approximation for the Green function which
is sufficient for producing an accurate result for the self-force. To date this has been shown to
be possible in the case of a scalar charge in Nariai \cite{Casals:2009zh},
Schwarzschild \cite{Casals:2013mpa,Wardell:2014kea} and Kerr \cite{Wardell:Capra17} spacetimes.
Given an approximation for the Green function valid throughout the past worldline, it is then
trivial to numerically integrate Eq.~\eqref{eq:hist} to obtain the self-force
(see Fig.~\ref{fig:convolution}). As mentioned previously, the divergences in the retarded Green function at null-geodesic intersections on the past worldline are all integrable singularities and do not pose a significant obstacle to accurate numerical evaluation of the integrals.
\subsubsection{Quasilocal expansion}
In order to obtain an approximation to the retarded Green function which is valid at early times,
it is convenient to start with the Hadamard form for the Green function (in Lorenz gauge)
\begin{equation}
G^{\mathrm{ret}}_{AB'}(x,x') = \Theta_{-} \Big[ U_{AB'} \delta (\sigma) - V_{AB'} \Theta (-\sigma) \Big] ,
\end{equation}
where $\Theta_{-}$ is analogous to the Heaviside step-function, being $1$ when $x'$ is in the
causal past of $x$, and $0$ otherwise, $\delta(\sigma)$ is the covariant form of the Dirac delta
function, $U_{AB'}$ and $V_{AB'}$ are symmetric bi-spinors/tensors and are regular for
$x'\rightarrow x$. The bi-scalar $\sigma \left(x,x'\right)$ is the Synge world function, which is
equal to one half of the squared geodesic distance between $x$ and $x'$. In particular,
$\sigma(x,x) = 0$ and $\sigma(x,z)<0$ when $x$ and $z$ are timelike separated. Because of the
limiting procedure in the history integral, Eq.~\eqref{eq:hist}, only the term involving $V_{AB'}$
is non-zero, and at all required points along the worldline $\Theta_- = 1 = \Theta(-\sigma)$. The
problem of determining the retarded Green function at early times therefore reduces to finding an
approximation for $V_{AB'}(x,x')$ which is valid for $x$ and $x'$ close together.
Several methods have been developed for computing approximations to $V_{AB'}$. Fundamentally, they
rely on either the use of a series expansion, or on the use of numerically evolved transport
equations (ordinary differential equations defined along a worldline). The series expansion
approach has been the most fruitful to date with results including: leading-order coordinate
expansions in Schwarzschild and Kerr spacetimes for scalar \cite{Ottewill:2007mz,Ottewill:2008uu}
and gravitational cases \cite{Anderson:2004eg}, high-order coordinate expansions in spherically
symmetric spacetimes (including Schwarzschild) \cite{Casals:2009xa}, formal covariant expansions
in generic spacetimes \cite{Avramidi:1986mj,Avramidi:2000pia,Decanini:2005gt,Ottewill:2009uj},
and moderately high-order coordinate expansions in Schwarzschild \cite{Heffernan:2012su} and Kerr
\cite{Heffernan:2012vj} spacetimes. The only numerical calculation I am aware of was done in
\cite{Ottewill:2009uj} for generic spacetimes (with an example application in Schwarzschild
spacetime).
The series expansion method produces an expression for $V(x,x')$ as a power series in the coordinate
distance between $x$ and $x'$. For example, for the scalar case in Schwarzschild spacetime it
takes the form
\begin{equation} \label{eq:CoordGreen}
V(x,x') =
\sum_{i,j,k=0}^{\infty} v_{ijk}(r) ~ (t-t')^{2i} (1-\cos\gamma)^j (r-r')^k,
\end{equation}
where $\gamma$ is the angular separation of the points and the $v_{ijk}$ are analytic functions of
$r$ and $M$. It is straightforward to take partial derivatives of these expressions at either
spacetime point to obtain the derivative of the Green function. Although this series on its own may
be sufficient for use in the quasilocal component of a worldline convolution, it turns out that
some simple tricks allow for a vast improvement in accuracy. It turns out that, because
$V(x,x')$ diverges at the edge of the normal neighbourhood, the series approximation benefits
significantly from Pad\'e resummation which incorporates information about the form of the
divergence \cite{Casals:2009xa}. One minor caveat is that since Pad\'e re-summation is only well defined for
series expansions in a single variable, it is necessary to first expand $r'$ and $\gamma$ in a
Taylor series in $t-t'$, using the equations of motion to determine the higher derivatives
appearing in the series coefficients. Then, with $V(x,x')$ written as a power series in $t-t'$
alone, a standard diagonal Pad\'e approximant provides an accurate representation of the Green
function in the quasi-local region.
\subsubsection{Frequency domain methods}
\label{sec:GF-FD}
Frequency domain methods for computing the retarded Green function rely crucially on the
separability of the wave equation. In the scalar, Schwarzschild case that represents the current
state-of-the art\footnote{More generally, Teukolsky \cite{Teukolsky:1972my,Teukolsky:1973ha} showed that the field
equations may be separated in Kerr spacetime in the gravitational case by making use of the
spin-weighted spheroidal harmonics in place of the spherical harmonics. A series of works --- pioneered
by Regge and Wheeler \cite{Regge:1957td} and improved upon by others
\cite{Zerilli:1971wd,Vishveshwara:1970cc,Moncrief:1974am,Cunningham:1978zfa,Cunningham:1979px,Martel:2005ir,Berndtson:2009hp} --- achieved a
similar separation in the Schwarzschild gravitational case by making use of tensor spherical
harmonics. However, separability alone is not sufficient and there remain some technical issues
which have yet to be solved before solutions of the Teukolsky or Regge-Wheeler equations could be
used in a worldline convolution approach. Most important is the issue of gauge; the MiSaTaQuWa
equations, Eqs.~\eqref{eq:inst} and \eqref{eq:hist}, were derived in the Lorenz gauge, whereas
solutions of the Teukolsky and Regge-Wheeler equations are in a gauge different from Lorenz gauge.
Fortunately, recent work \cite{Pound:2013faa} has resolved many of the conceptual issues associated
with gauge choice.}
\cite{Casals:2013mpa} (also see \cite{Casals:2009zh} for a related calculation in Nariai
spacetime), this can be achieved by writing the Green function as a sum of spherical harmonic and
Fourier modes
\begin{equation}
\label{eq:G-FD}
G^{\rm ret}(x,x')= \frac{1}{r\, r'} \sum_{\ell=0}^{\infty}
P_{\ell}(\cos\gamma)\frac{1}{2\pi}
\int_{-\infty+i \epsilon}^{\infty+i \epsilon} \hat{g}_{\ell}(r,r';\omega) e^{-i\omega (t-t')} d\omega.
\end{equation}
Here, $\epsilon>0$ is a formal parameter to ensure the correct boundary conditions are satisfied
for a retarded Green function. Substituting this into the wave equation,
Eq.~\eqref{eq:BoxG-scalar}, one obtains an independent set of ordinary differential equations for
$\hat{g}_\ell(r,r'; \omega)$, one equation for each $\ell$ and $\omega$,
\begin{equation}
\label{eq:radial}
\left[\frac{d^2}{dr_*^2}+\omega^2-V_\ell(r)\right]\hat{g}_{\ell}(r,r';\omega)=-\delta(r_*-r'_*),
\end{equation}
with
\begin{equation}
V_\ell(r) \equiv \left(1-\frac{2M}{r}\right)\left[\frac{\ell(\ell+1)}{r^2}+\frac{2M}{r^3}\right].
\end{equation}
Here, $r_*\equiv r+2M\ln\left(\frac{r}{2M}-1\right)$ is the radial tortoise coordinate.
Given two linearly-independent solutions, $p(r;\omega)$ and $q(r;\omega)$, of the homogeneous
version of (\ref{eq:radial}), $\hat{g}_\ell(r,r';\omega)$ is given by
\begin{align}
\hat{g}_\ell(r,r';\omega)=\frac{p_\ell(r_<,\omega)q_\ell(r_>,\omega)}{W(p,q)},
\end{align}
where $r_>\equiv \max(r,r'),\ r_<\equiv \min(r,r')$ and $W(p,q)$ is the Wronskian. One is then
faced with the integral over frequencies in the inverse Fourier transform appearing in
\eqref{eq:G-FD}.
\begin{figure}[htb!]
\begin{center}
\includegraphics[width=6.1cm]{figs/GF-ell0-omega}
\includegraphics[width=6.1cm]{figs/GF-ell0-t}
\includegraphics[width=6.1cm]{figs/GF-ell10-omega}
\includegraphics[width=6.1cm]{figs/GF-ell10-t}
\end{center}
\caption{
Spherical harmonic modes of the Schwarzschild scalar Green function. Left: the frequency domain
Green function, $\hat{g}_\ell (r,r'; \omega)$, as a function of real frequency for $r=r'=10M$, $\ell = 0$
(top) and $\ell=10$ (bottom). Right: the corresponding time domain Green function, $g_\ell(r,r';
t-t')$ computed by Fourier transforming the frequency domain Green function using
Eq.~\eqref{eq:GF-fourier} with $\omega_{\rm max} = 8.5$ (blue, solid line), and by using the time
domain methods described in Sec.~\ref{sec:GF-TD} (orange, dashed line).
}
\label{fig:GF-ell}
\end{figure}
This may be achieved through straightforward integration along the real-$\omega$
axis (see Fig.~\ref{fig:GF-ell}); the only caveat is that the formally-infinite integral over frequencies should be cut off at some finite maximum frequency using a smooth window function, for example
\begin{equation}
\label{eq:GF-fourier}
g_{\ell} (r, r'; t-t') = \frac{1}{2\pi} \int_{-\infty+i \epsilon}^{\infty+i \epsilon} \hat{g}_{\ell}(r,r';\omega) e^{-i\omega (t-t')} \big(1-\erf[2(\omega - \omega_{\rm max})]\big)/2 \, d\omega.
\end{equation}
Alternatively, as was done in \cite{Casals:2013mpa}, the integration contour can be deformed into
the complex frequency plane following a proposal by Leaver \cite{Leaver:1986gd,PhysRevD.38.725}. In
Schwarzschild spacetime,
$\hat{g}_\ell(r,r';\omega)$ has simple poles in the lower semi-plane at the quasinormal mode frequencies,
along with a branch cut starting at $\omega=0$ and continuing along the negative imaginary axis
(see Fig.~\ref{fig:contour}). The deformed frequency integral is therefore given by an integral
over a high-frequency arc, an integral around a branch cut, and an integral over the residues at
the poles\footnote{In the Kerr spacetime there are indications that there may be additional branch
cuts to consider \cite{Kavanagh:Capra17}.}.
\begin{figure}[htb!]
\begin{center}
\includegraphics[width=7cm]{figs/contour}
\end{center}
\caption{
Contour deformation in the complex-frequency plane. The residue theorem of complex analysis allows
one to re-express the integral over (just above) the real line of the Fourier modes,
$\hat{g}_\ell(r,r';\omega)$, as an integral over a high-frequency arc plus an integral around a branch cut
and a sum over the residues at the poles.
Reproduced from Ref.~\cite{Casals:2013mpa}.
}
\label{fig:contour}
\end{figure}
The high-frequency arcs can be disregarded as they are likely to only contribute at very early times,
where the quasilocal approximation can be used. There are well-established methods for accurately
computing the location of the poles (quasinormal mode frequencies) and the residues at the poles.
The biggest difficulty in the frequency-domain approach is in evaluating the branch cut integral,
about which very little was known until recently (other than asymptotic approximations for e.g.
large radius or late times). Substantial recent progress has established methods for calculating
this branch cut contribution \cite{Casals:2011aa}; these methods were used in \cite{Casals:2013mpa} to compute the self-force in Schwarzschild spacetime.
The final step in the frequency domain approach is to sum over spherical harmonic modes to produce
the full Green function. Here, the distributional parts of the full Green function can cause poor
convergence in the mode-sum. Fortunately, there is a straightforward solution to this problem: by
smoothly cutting off the mode sum at large $\ell$,
\begin{equation}
G^{\rm ret}(x,x')= \frac{1}{r\, r'} \sum_{\ell=0}^{\ell_{\rm max}}
P_{\ell}(\cos\gamma) g_{\ell}(r,r'; t-t') e^{-\ell^2 / 2 \ell_{\rm cut}^2}
\end{equation}
one obtains a mollified retarded Green which is appropriate for use in a self-force calculation,
and whose sum over $\ell$ converges. Empirically, it has been found that choosing
$\ell_{\rm cut} \approx \ell_{\rm max}/5$ gives good results.
\subsubsection{Time domain approaches}
\label{sec:GF-TD}
The frequency domain approach to computing the retarded Green function has several shortcomings. It
relies on relatively difficult technical calculations and has poor convergence properties at early
times and in scenarios where the worldline is not well-represented by a discrete spectrum of
frequencies. Recent work has shown that the Green function may also be accurately
computed (at least for the purposes of self-force calculations) in the time domain using
straightforward numerical evolutions. A time domain calculation sidesteps issues related to a wide
frequency spectrum and appears to exhibit much better convergence properties at early times.
There are two closely-related proposals for computing the Green function in the time domain.
In \cite{Zenginoglu:2012xe}, Eq.~\eqref{eq:BoxG-scalar} was numerically solved as an initial value problem,
with the delta-function source approximated by a narrow Gaussian. A reformulation of this approach
as a homogeneous problem with Gaussian initial data was subsequently given in \cite{Wardell:2014kea}. It
was found that these ``numerical Gaussian'' approaches are able to approximate the retarded Green
function remarkably well in a large region of the space required by worldline convolutions. The
size of the Gaussian limits the scale of the smallest features which can be resolved, but it turns
out that this is not significantly detrimental to a self-force calculation.
The only regimes where the numerical Gaussian approach is not well suited to computing the
retarded Green function are at very early and very late times. At early times the direct
$\delta(\sigma(x,x))$ term (which must be excluded from the worldline convolution of the retarded Green
function) is smeared out and is difficult to isolate from the rest of the Green function. There is
no conceptual difficulty at late times, but the reality of a numerical evolution is that it can
only be run for a finite time. Fortunately, both of these issues are easily overcome; the former by
using the quasilocal expansion at early times, and the latter by using a late-time expansion of the
branch cut integral (see Fig.~\ref{fig:matched-GF}).
\begin{figure}[htb!]
\begin{center}
\includegraphics[width=8.5cm]{figs/matched-gf}
\includegraphics[width=8.5cm]{figs/partial-field-circ}
\end{center}
\caption{
Top: Matched Green function along an eccentric orbit (eccentricity $e=0.5$ and semilatus rectum $p=7.2$) in Schwarzschild spacetime.
The full Green function (black) is constructed by combining a quasilocal approximation at early
times (orange, dashed line) with a time domain calculation at intermediate times (blue, dashed
line) and a late-time branch cut approximation at late times (green, dashed line).
Bottom: Integrating the retarded Green function (excluding the direct part) up to some time point on
the past worldline gives the contribution to the regularized self-field from all points on the past
worldline up to that time. With an eccentric orbit of period $\approx 317 M$, we see that a good
approximation to the self-field is obtained by including the contributions from less than one full
orbit. Note that the formal divergences of the Green function are integrable and do not cause
significant numerical difficulty in computing the self-force.
}
\label{fig:matched-GF}
\end{figure}
In the time domain approach, each numerical time domain evolution gives the Green function
$G(x_0,x')$ for a single base point $x_0$, and for all source points, $x$. As a result, a single
numerical calculation can only be used to compute the self-force at a single spacetime point, but
it can be computed for any past-worldline ending up at that point. This is in contrast to other
self-force methods, where a single orbit is considered at a time, but a single calculation gives
the self-force at all points on the orbit. The problem of efficiently spanning the parameter space of base points $x_0$ is an ideal application of reduced order methods
\cite{Field:2013cfa,Wardell:Capra17}.
\subsection{Mode-sum regularization} \label{sec:msr}
The mode-sum regularization scheme, first proposed by Barack and Ori \cite{Barack:1999wf}, has
proven highly successful as a computational self-force strategy, having been applied to the
computation of the the self force for a variety of configurations in Schwarzschild
\cite{Barack:2000zq,Burko:2000xx,Detweiler:2002gi,DiazRivera:2004ik,Haas:2006ne,Haas:2007kz,
Canizares:2009ay,Canizares:2010yx,Jaramillo:2011gu,Barack:2007tm,Barack:2002ku,Sago:2008id,Sago:2009zz,Barack:2010tm,
Thornburg:2010tq,Keidl:2010pm,Shah:2010bi,Merlin:2014qda} and Kerr
\cite{Warburton:2010eq,Warburton:2011hp,Shah:2012gu,Warburton:2014bya,Isoyama:2014mja}
spacetimes. The success of the
method hinges on the fact that the first order retarded field diverges in a way which can be
effectively smoothed out by a spectral decomposition in the angular directions. More specifically,
the divergence appears as an odd power of $1/s$, where $s^2 = (g^{\alpha\beta} + u^\alpha u^\beta)
\sigma_{;\alpha} \sigma_{;\beta}$ is an appropriate measure of distance from the worldline. The
result is that the $1/s$ divergence of the field near the worldline turns into an infinite sum of
modes, each of which are individually finite (but possibly discontinuous) on the worldline. The
divergence of the field then manifests itself through the failure of the (infinite) sum over modes
to converge. Conveniently, this odd-in-$1/s$ property of the retarded field also holds for its
derivatives, both the first derivatives required for the self-force and higher derivatives which
are useful for computing higher-multipole gauge invariants \cite{Dolan:2013roa,Dolan:2014pja}. As a
result, an arbitrary number of derivatives of the first order retarded field is represented by an
infinite sum of modes, each of which are individually finite (but possibly discontinuous) on the
worldline. The trade-off is that as the number of derivatives increases the divergence of the sum
over modes becomes increasingly strong.
Since the individual modes are finite, numerical calculations of the retarded field and its
derivatives can be done in the reduced $(t,r)$ space without encountering any numerical
divergences. The remaining piece of the problem is a method for rendering the divergent sum over
modes finite. An analytic decomposition of the angular dependence of the Detweiler-Whiting singular
field yields ``regularization parameters'' which may be subtracted mode-by-mode from the numerical
retarded field values. Provided all parts of the Detweiler-Whiting singular field which don't
vanish on the worldline are included in the calculation, the regularized sum over modes is
convergent on the worldline and one never encounters any numerical divergences.
It is important to note that the use of the Detweiler-Whiting singular field is not merely a
convenience; it also provides well-motivated physical grounds to justify the mode-sum regularization
approach. Other ad-hoc approaches based on identifying the asymptotically divergent contributions
to the mode-sum often produce the correct regularization parameters, but their use is difficult to
justify rigorously and can easily lead to incorrect results. Crucially, there is no way of knowing
whether such an ad-hoc regularization procedure is producing a physically correct result, or if
important contributions are being overlooked. They should therefore not be relied on without
extreme care.
Unfortunately, despite its success in first order calculations, mode-sum regularization alone is
not an effective tool for computing the second order self-force. The reason for this is
straightforward: the second order retarded field contains divergent terms which appears in the form
of even powers of $1/s$. Intuitively, this arises from the fact that the second order field
contains terms which depend quadratically on the first order field. Unlike the odd-in-$1/s$ case,
the angular decomposition of $1/s^2$ leads to individual modes which diverge logarithmically as the
worldline is approached. Fortunately, all is not lost for the mode-sum method as a computational
tool at second order. Provided the leading order logarithmic divergence is subtracted by some other
means (for example, using the effective source approach), regularization parameters may be used to
accelerate the rate of convergence of the mode sum. Such a hybrid scheme complements the generality
of the effective source approach with the computational efficiency of mode-sum regularization.
\subsubsection{Regularization parameters}
The computation of regularization parameters has been addressed by a series of calculations
stemming from the original Barack-Ori derivation. Barack and Ori's original work gave the first two
self-force regularization parameters for scalar, electromagnetic and Lorenz-gauge gravitational
cases in both Schwarzschild and Kerr spacetimes. These are sufficient for computing the regularized
self-force, but they yield mode-sums which have relatively poor quadratic convergence with the
number of modes included. Since the Detweiler-Whiting regularized field is a smooth, homogeneous
function in the vicinity of the worldline, its mode-sum representation converges exponentially.
This spectral convergence is spoiled by the fact that only the portion of the Detweiler-Whiting
singular field which does not vanish on the worldline is subtracted by the leading order
regularization parameters. The mode decomposition effectively contains information about the
extension of the field off the worldline to the entire two-sphere, so the regularized field
contains residual pieces of the Detweiler-Whiting singular field off the worldline, but on a
two-sphere of the same radius. By deriving higher-order regularization parameters one can subtract
these residual pieces order-by-order, leaving a mode-sum which is more and more rapidly convergent
(see Fig.~\ref{fig:modesum}).
The derivation of these higher-order parameters is closely related to the computation of the
quasilocal expansion of the Green function and has been addressed in a series of papers: the first
higher order parameter was given in \cite{Detweiler:2002gi} for the case of a scalar charge on a circular
orbit in Schwarzschild spacetime, and for eccentric geodesic orbits in \cite{Haas:2006ne}. This
was subsequently extended by several further orders for equatorial geodesic motion in the scalar,
electromagnetic and gravitational cases in both Schwarzschild \cite{Heffernan:2012su} and Kerr
\cite{Heffernan:2012vj} spacetimes. Recent work has also produced parameters for accelerated
worldlines \cite{Casals:2012qq,Linz:2014pka}.
\begin{figure}[htb!]
\begin{center}
\includegraphics[width=10cm]{figs/modesum}
\end{center}
\caption{
Spherical harmonic modes of the self-force for a point scalar charge on a
circular orbit of radius $r_0 = 6M$ in Schwarzschild spacetime. By subtracting
analytically determined high-order ``regularization parameters'', the sum over
modes is rendered more and more rapidly convergent. Each regularization parameter,
$F_r^{[n]}$, behaves asymptotically for large-$\ell$ as $1/\ell^{n}$.
}
\label{fig:modesum}
\end{figure}
Within these derivations of high-order regularization parameters, a subtle ambiguity appears
through the elevation of the four-velocity from a quantity defined on a worldline to a quantity
defined everywhere on the two-sphere. A natural, covariant choice is to define this off-worldline
extension through parallel transport, $u^{\alpha'} = g_{\alpha}{}^{\alpha'} u^\alpha$. However, in
practical calculations it is often convenient to make a coordinate choice. For example, a common
choice is to define the extension in terms of ``constant coordinate components'', i.e. to define
$u^{\alpha'}$ such that its components in some coordinate system have a constant value everywhere on
the two-sphere. This is a perfectly valid choice, as is any other choice where $u^{\alpha'}$ agrees
with the actual four-velocity when evaluated on the worldline. The only caveat is that the
regularization parameters beyond the leading two orders change depending on the particular choice
of extension. As a result, in order to use higher-order regularization parameters it is essential
that a compatible choice of off-worldline extension is used in the retarded field calculation.
\subsubsection{Choice of basis and gauge}
There are two other factors which must be considered in the mode-sum scheme: the choice of angular basis functions for the spectral decomposition and the choice of gauge in the electromagnetic and
gravitational cases.
The choice of basis functions is typically motivated by their ability to produce separability of
the retarded field equations; an appropriate choice of basis functions yields an independent set
of equations in $(t,r)$ space for each individual mode. For the case of a scalar charge in Schwarzschild
spacetime, the appropriate basis functions are the standard spherical harmonics. In the Kerr case
the spheroidal harmonics are required for separability. For electromagnetic and gravitational cases,
there are several choices. For Schwarzschild spacetime one can choose between a tensor harmonic
basis and a basis of spin-weighted spherical harmonics. In the Kerr case only the spin-weighted
spheroidal harmonics are known to yield separability \cite{Teukolsky:1972my,Teukolsky:1973ha}.
This choice of angular basis also affects the regularization parameters. The parameters for scalar
spherical harmonic modes, tensor harmonic modes, spin-weighted spherical harmonic modes and
spheroidal harmonic modes are all potentially different. However, it is always possible project the
tensor, spin-weighted, or spheroidal harmonic modes onto the scalar spherical harmonic basis. In
fact, since the regularization parameters are most easily obtained for the scalar spherical
harmonic basis, calculations of the retarded field are typically done using the computationally
most convenient basis and the result is then projected onto the scalar spherical harmonics before
the regularization and sum-over-modes steps are done \cite{Barack:2010tm}.
A more difficult issue in the electromagnetic and gravitational cases is the
choice of gauge. The Detweiler-Whiting singular field is defined in Lorenz
gauge (since it is derived from a Lorenz gauge Green function), but numerical
calculations of the retarded field are more easily done in either radiation or
Regge-Wheeler gauge. The existence of tensor spherical harmonics makes a
Lorenz-gauge calculation possible, if somewhat cumbersome in the Schwarzschild
case \cite{Akcay:2010dx,Akcay:2013wfa,Osburn:2014hoa}. One obtains a coupled
set of 10 equations for the metric perturbation, but there is no coupling
between different different tensor-harmonic modes. Unfortunately, this does not
hold for the Kerr case as there are no known tensor spheroidal harmonics.
Rather than trying to derive tensor spheroidal harmonics, a better approach is
to work with the relatively straightforward Teukolsky equation in radiation
gauge \cite{Shah:2012gu}. The difficulty then is in identifying the appropriate regularization
parameters, particularly since the gauge transformation from Lorenz gauge is
itself often singular. This remained an open problem until recently, when an
understanding of how to apply the mode-sum regularization scheme in radiation
gauge was finally established \cite{Pound:2013faa} and applied in a self-force
calculation \cite{Merlin:2014qda}.
\subsubsection{Mode-sum regularization in the frequency domain}
The mode-sum scheme provides a method for computing the regularized self-force using numerical data
for the modes of the retarded field in the reduced $(t,r)$ space. There are, however, several
possibilities for computing this numerical data. One option relies on a further decomposition of
the time dependence into Fourier-frequency modes, in an analogous way to the frequency-domain Green
function described in Sec.~\ref{sec:GF-FD} above. Using a spin-weighted spheroidal harmonic basis,
this leads to a radial equation for the Teukolsky function,
\begin{equation}
\Delta^{-s} \frac{d}{dr} \Big( \Delta^{s+1} \frac{dR}{dr}\Big) + V(r) R = \mathcal{T}(r)
\end{equation}
with the potential given by
\begin{equation}
V(r) = \frac{K^2 - 2 i s (r - M) K}{\Delta} + 4 i r \omega s - \lambda.
\end{equation}
Here, $\Delta \equiv r^2 - 2 M r + a^2$, $K \equiv (r^2 + a^2) \omega - a m$, $\lambda$ is an
eigenvalue of the spheroidal equation,
\begin{align}
\Big[\frac{1}{\sin \theta} \frac{d}{d\theta} \Big(\sin \theta \frac{d}{d\theta}\Big) &- a^2 \omega^2 \sin^2 \theta - \frac{(m+s\cos \theta)^2}{\sin^2 \theta} \nonumber \\
& - 2 a \omega s \cos \theta + s + 2 m a \omega + \lambda] {}_{s} S_{\ell m} = 0
\end{align}
and $a$ and $M$ are the Kerr spin and mass parameters. For $a=0$, $\lambda = (\ell-s)(\ell+s+1)$
and this reduces to the Schwarzschild wave equation decomposed into spin-weighted spherical
harmonics (in radiation gauge for the gravitational case), while for $s=0$ it reduces to the scalar
wave equation. A decomposition of the point-particle source term yields a source involving
$\delta(r - r_p)$ (and in some cases its derivative), where $r_p$ is the radial location of the
particle. In practice, the frequency domain mode-sum approach proceeds in the same way as for the
Green function: two independent solutions of the homogeneous radial equation are obtained and are
matched at the particle's location\footnote{For eccentric orbits where the particle can not be
considered to be at a single radial location in the frequency domain this matching must be modified
slightly using, for example, the method of extended homogeneous solutions
\cite{Barack:2008ms,Hopper:2010uv}.}. Then, the distributional sources do not introduce any numerical
difficulty as they simply appear as jumps when matching the homogeneous inner and outer
solutions.
The solutions of the Teukolsky equation for a given $(\ell,m,\omega)$ can be obtained either
through straightforward numerical integration of the radial ordinary differential equation
(typically with some modifications to improve numerical accuracy \cite{Sasaki:1981kj,Sasaki:1981sx}) or as an
approximation in the form of an infinite convergent series of hypergeometric functions.
The latter method is based an idea originally developed by Leaver \cite{Leaver:1986JMP} and now commonly
referred to as the ``MST'' method, after Mano, Suzuki, and Takasugi \cite{Mano:1996vt} who reformulated
it into its current form. It provides an efficient and highly accurate method for computing
solutions of the radial equation. For example, recent results have used it to compute solutions
accurate to several hundred decimal places, allowing the solutions to be used to determine
previously unknown high-order post-Newtonian parameters \cite{Shah:2013uya,Shah:2014tka,Bini:2014nfa,Bini:2014ica,Bini:2014zxa}. For a
comprehensive review of the MST approach, see the living review by Sasaki and Tagoshi
\cite{Sasaki:2003xr} and references therein.
The frequency-domain approach is particularly appropriate in scenarios where the worldline is well
represented by a discrete spectrum involving a small number of frequencies. In such cases, mode-sum
regularization is by far the optimal choice and is unparalleled in its accuracy and computational
efficiency \cite{Barton:2008eb}. The prototypical example is a circular orbit, in which only a single frequency is
present. In the case of eccentric equatorial orbits (and inclined circular orbits in the Kerr
case), there are two fundamental frequencies, and also an infinite number of higher harmonics
produced from combinations of the fundamental frequencies. For mildly eccentric orbits this does
not cause a great deal of difficulty. However, for more eccentric orbits (with eccentricities $e
\gtrsim 0.5$) an increasingly large number of frequencies must be included and the competitive
advantage of the frequency domain approach is lost \cite{Barton:2008eb}. Even worse, generic geodesic orbits in Kerr
spacetime have three fundamental frequencies and the computational difficulty is so high that a
calculation has yet to be attempted. Similarly, unbound orbits and other cases such as radial infall are not well suited to frequency domain methods. Apart from these deficiencies, the frequency
domain mode-sum approach has been highly successful for producing results.
\subsubsection{Mode-sum regularization in the time domain}
The mode-sum scheme may also be applied in the time domain by skipping the Fourier decomposition
step and instead solving a set of $1+1$D partial differential equations in $(t,r)$ space. The main
difficulty then is in appropriately handling the distributional source term which has the form
$\delta(r - r_p(t))$. One solution, used in
\cite{Barack:2007tm,Barack:2010tm,Haas:2007kz,Haas:2011np,Sundararajan:2007jg,Sundararajan:2008zm}, is to use a discretised
representation of a delta function and to construct the computational grid such that the worldline
only ever passes \emph{between} grid points.
An alternative approach is to reformulate the problem in an
analogous way to the frequency-domain method. By splitting the computational domain into two
domains --- one either side of the particle --- the delta-function source can be reformulated in
terms of a jump in the fields and their derivatives at the interface of the two domains. This
method is well suited to highly-accurate spectral numerical methods as all of the numerically
evolved fields are smooth functions. It was implemented using discontinuous-Galerkin methods in
\cite{Field:2009kk,Field:2010xn}
and using Chebyshev pseudo-spectral methods in \cite{Canizares:2009ay,Canizares:2010yx}. Eccentric orbits present a small
additional complexity in this case as the particle must always lie on the domain interface. This is
easily achieved by introducing a time-dependent mapping between the computational and physical
coordinates of the system \cite{Canizares:2010yx,Field:2010xn}.
\subsubsection{Limitations of the mode-sum regularization scheme}
Despite its resounding success to date, the mode-sum regularization scheme has some unfortunate
disadvantages which make it ill-suited as a general-purpose method for self-force calculations:
\begin{enumerate}
\item Its application in the Kerr case is only straightforward in the frequency domain, since the
field equations are not separable in the time domain\footnote{It may still be possible to use
mode-sum regularization for the Kerr case in the time domain by decomposing the field equations
into \emph{spherical} harmonics and evolving the resulting infinitely coupled set of $1+1$D
partial differential equations in a similar manner to the Schwarzschild case.}
\item It relies on the use of Lorenz gauge for regularization in the gravitational case. This is
not a major issue in the Schwarzschild case since the tensor spherical harmonics may be used to
decouple the Lorenz gauge field equations in the angular directions (leaving 10 coupled $1+1$D
equations for each $\ell,m$ mode). However, there are no know tensor spheroidal harmonics which
would be required for the Kerr case, and even if there were they would likely not be applicable
in the time domain (again, it is conceivable that a coupled system of equations involving tensor
spherical harmonics could be used, but the coupling would result in considerable complexity).
\item It is not applicable beyond first perturbative order, since the modes of the second
order perturbation diverge logarithmically near the worldline.
\end{enumerate}
The first two issues are not necessarily showstoppers. There have been several attempts at $1+1$D
time-domain implementations using coupled spherical-harmonic modes in the Kerr case
\cite{Dolan:2012yt,Stein:Thesis}, and recent progress on reformulating mode-sum regularization for radiation gauge
has clarified the regularization issue \cite{Pound:2013faa}. The third point, however, appears insurmountable. For these
reasons, among others, the effective source method, described in the next section, was developed.
\subsection{Effective source approach}
Proposed in 2007 as a solution to the shortcomings of mode-sum regularization, the effective source
approach\footnote{Note that the effective source proposed by Lousto and Nakano
\cite{Lousto:2008mb} is similar in spirit, but differs in that it is not derived from the
Detweiler-Whiting singular field.} provides an alternative method for handling the divergence of
the retarded field. Rather than first computing the retarded field and then subtracting the
singular piece as a post-processing step, one can instead work directly with an equation for the
regular field. This idea --- independently proposed by Barack and Golbourn \cite{Barack:2007jh} and
by Vega and Detweiler \cite{Vega:2007mc} --- has the distinct advantage of involving only regular
quantities from the outset, making it applicable in a wider variety of scenarios than the mode-sum
scheme. In particular, since it does not rely on a mode decomposition of the retarded field, it can
be used by any numerical prescription for solving the retarded field equations, whether in the
frequency domain (where the method is really just a generalisation of mode-sum regularization) or
in the time domain as a $1+1$D, $2+1$D or even $3+1$D problem.
The basic idea is to use the split of the retarded field into regular and singular pieces,
Eq.~\eqref{eq:singular-regular-split}, to rewrite the field equations, Eqs.~\eqref{eq:field-gravity},
\eqref{eq:field-scalar} and \eqref{eq:field-em} as equations for the Detweiler-Whiting regular
field,
\begin{subequations}
\label{es:effsource-field}
\begin{align}
(\Box&-\xi R)\Phi^{\rm R} = (\Box-\xi R)(\Phi^{\rm ret} - \Phi^{\rm S}) \nonumber \\
&= -4\pi\rho - (\Box-\xi R) \Phi^{\rm S},
\end{align}
\begin{align}
(\Box \delta_\alpha{}^\beta &- R_\alpha^\beta) A^{\rm R}_\beta =
(\Box \delta_\alpha{}^\beta - R_\alpha^\beta) (A^{\rm ret}_\beta-A^{\rm S}_\beta) \nonumber \\
&= - 4 \pi e \int g_{\alpha\alpha'} u^{\alpha'} \delta_4(x, z(\tau')) d \tau' - (\Box \delta_\alpha{}^\beta - R_\alpha^\beta) (A^{\rm S}_\beta),
\end{align}
\begin{align}
(\Box & \delta_\alpha{}^\gamma \delta_\beta{}^\delta + 2 C_{\alpha}{}^{\gamma}{}_{\beta}{}^{\delta }) \bar{h}^{\rm R}_{\gamma\delta} =
(\Box \delta_\alpha{}^\gamma \delta_\beta{}^\delta + 2 C_{\alpha}{}^{\gamma}{}_{\beta}{}^{\delta }) (\bar{h}^{\rm ret}_{\gamma\delta} - \bar{h}^{\rm S}_{\gamma\delta}) \nonumber \\
&= - 16 \pi \mu \int g_{\alpha'(\alpha} u^{\alpha'} g_{\beta)\beta'} u^{\beta'} \delta_4(x, z(\tau')) d \tau' - (\Box \delta_\alpha{}^\gamma \delta_\beta{}^\delta + 2 C_{\alpha}{}^{\gamma}{}_{\beta}{}^{\delta }) (\bar{h}^{\rm S}_{\gamma\delta}).
\end{align}
\end{subequations}
If the singular field used in the subtraction is exactly the Detweiler-Whiting singular field, then
the two terms on the right hand side of this equation cancel and the regularized field would be a
homogeneous solution of the wave equation. Unfortunately, one typically does not have access to an
exact expression for the singular field. Indeed, the Detweiler-Whiting singular field is only
defined through a Hadamard parametrix which is not even defined globally. Instead, the best one can
typically do is a local expansion which is valid only in the vicinity of the worldline. Borrowing
the language of Barack and Golbourn, we refer to an approximation to the singular
field as a ``puncture'' field, $\Phi^{\rm S} \approx \Phi^{\rm P}$, $A_\alpha^{\rm S}
\approx A_\alpha^{\rm P}$, $\bar{h}_{\alpha\beta}^{\rm S} \approx \bar{h}_{\alpha\beta}^{\rm P}$.
Then, the corresponding approximate regular field --- referred to as the ``residual field'' --- is
no longer a solution of the homogeneous equation, but instead is a solution of the sourced equation
with an \emph{effective source} (see Fig.~\ref{fig:effsource}) which is defined to be the right-hand
side of Eq.~\eqref{es:effsource-field} with the puncture field substituted for the singular field
\begin{subequations}
\begin{align}
S^{\rm eff} &= - 4 \pi q \int \delta_4(x, z(\tau')) d \tau' - (\Box-\xi R)\Phi^P, \\
S_\alpha^{\rm eff} &= - 4 \pi e \int g_{\alpha\alpha'} u^{\alpha'} \delta_4(x, z(\tau')) d \tau' - (\Box \delta_\alpha{}^\beta - R_\alpha^\beta) (A^{\rm P}_\beta), \\
S_{\alpha\beta}^{\rm eff} &= - 16 \pi \mu \int g_{\alpha'(\alpha} u^{\alpha'} g_{\beta)\beta'} u^{\beta'} \delta_4(x, z(\tau')) d \tau' - (\Box \delta_\alpha{}^\gamma \delta_\beta{}^\delta + 2 C_{\alpha}{}^{\gamma}{}_{\beta}{}^{\delta }) (\bar{h}^{\rm P}_{\gamma\delta}).
\end{align}
\end{subequations}
\begin{figure}[htb!]
\begin{center}
\includegraphics[width=6cm]{figs/effsrc}
\includegraphics[width=6cm]{figs/effsrc-smooth}
\end{center}
\caption{
The effective source for a scalar particle on a circular orbit in Schwarzschild spacetime. The
source is generically non-smooth in the vicinity of the worldline (left), but the smoothness can be
improved by incorporating higher-order parts of the Detweiler-Whiting singular field into the
source.
}
\label{fig:effsource}
\end{figure}
Note that the presence of a distributional component of the source on the worldline is merely a
formal prescription; in practice the puncture field is chosen so that it exactly cancels this
distributional component on the worldline. This effective source is then finite everywhere, but has
limited differentiability on the worldline. This makes it well-suited to numerical implementations
since no divergent quantities are ever encountered. The only numerical difficulty arises from the
non-smoothness of the source in the vicinity of the worldline (see Fig.~\ref{fig:effsource}), which
leads to numerical noise in the computed self force. The noise can be reduced by making the source
smoother using higher-order parts of the Detweiler-Whiting singular field. As shown in
Fig.~\ref{fig:src-smoothness}, at the same numerical resolution a higher order source ($C^2$ in
this case) eliminates the vast majority of the numerical noise that is present when using a lower
order source ($C^0$ in the case in the figure). The cost of this improved accuracy is a source which
is considerably more complicated, and costly to compute in a numerical code.
\begin{figure}[htb!]
\begin{center}
\includegraphics[width=8cm]{figs/smooth-src}
\end{center}
\caption{
Radial self-force for a scalar particle on an eccentric orbit in Schwarzschild spacetime, computed
with a $3+1$D implementation of the effective source scheme. Here, the independent variable $\chi$ is a
``relativistic anomaly parameter'' defined through $r = p M / (1+e \cos \chi)$, with $p$ the
semilatus rectum and $e$ the eccentricity.
The high-frequency errors using a continuous
but non-differentiable source (blue) are dramatically decreased by using a twice differentiable source (orange)
obtained from a higher-order approximation to the
Detweiler-Whiting singular field. For reference, a highly accurate value computed using frequency-domain
mode-sum regularization is also included (dashed, black). Figure based on version presented in Ref.~\cite{Diener:APS}.
}
\label{fig:src-smoothness}
\end{figure}
An additional level of complexity arises from the fact that the puncture field is defined only in
the vicinity of the worldline. To avoid ambiguities in its definition far from the worldline, one
must ensure that the puncture field goes to zero there. This is most easily achieved by multiplying
it by a window function, $\mathcal{W}$, with properties such that it only modifies the puncture
field in a way that its local expansion about the worldline is preserved to some chosen order.
In a first-order calculation of the self-force, it suffices to choose $\mathcal{W}$ such that
$\mathcal{W}(x_p) = 1$, $\mathcal{W'}(x_p) = 0$, $\mathcal{W''}(x_p) = 0$ and $\mathcal{W} = 0$ far
away from the worldline.
\begin{figure}[htb!]
\begin{center}
\includegraphics[width=8cm]{figs/l1m1}
\end{center}
\caption{
The $\ell=1, m=1$ spherical-harmonic mode of the residual field, $\Phi^{\rm res}$, for a scalar point particle on a circular orbit in Schwarzschild spacetime. This was produced in
\cite{Warburton:2013lea} using a frequency-domain effective source approach.
}
\label{fig:residual}
\end{figure}
The residual field (see Fig.~\ref{fig:residual}) then obeys
\begin{subequations}
\begin{align}
(\Box-\xi R)\Phi^{\rm res} &= S^{\rm eff} \label{eq:field-residual-scalar} \\
(\Box \delta_\alpha{}^\beta - R_\alpha^\beta) A^{\rm res}_\beta &= S_{\alpha}^{\rm eff} \label{eq:field-residual-em} \\
(\Box \delta_\alpha{}^\gamma \delta_\beta{}^\delta + 2 C_{\alpha}{}^{\gamma}{}_{\beta}{}^{\delta }) \bar{h}^{\rm res}_{\gamma\delta} &= S_{\alpha\beta}^{\rm eff} \label{eq:field-residual-gravity}
\end{align}
\end{subequations}
and has the properties
\begin{subequations}
\begin{gather}
\Phi^{\rm res} (x_p) = \Phi^{\rm R}(x_p), \quad \nabla_\alpha \Phi^{\rm res}(x_p) = \nabla_\alpha \Phi^{\rm R}(x_p),
\nonumber \\
\Phi^{\rm res}(x) = \Phi^{\rm ret}(x) \quad \text{for} \quad x \not\in \operatorname{supp}(\mathcal{W}),
\end{gather}
\begin{gather}
A_{\alpha}^{\rm res} (x_p) = A_{\alpha}^{\rm R}(x_p), \quad \nabla_\alpha A_{\alpha}^{\rm res}(x_p) = \nabla_\alpha A_{\alpha}^{\rm R}(x_p),
\nonumber \\
A_{\alpha}^{\rm res}(x) = A_{\alpha}^{\rm ret}(x) \quad \text{for} \quad x \not\in \operatorname{supp}(\mathcal{W}),
\end{gather}
\begin{gather}
h_{\alpha\beta}^{\rm res} (x_p) = h_{\alpha\beta}^{\rm R}(x_p), \quad \nabla_\alpha h_{\alpha\beta}^{\rm res}(x_p) = \nabla_\alpha h_{\alpha\beta}^{\rm R}(x_p),
\nonumber \\
h_{\alpha\beta}^{\rm res}(x) = h_{\alpha\beta}^{\rm ret}(x) \quad \text{for} \quad x \not\in \operatorname{supp}(\mathcal{W}).
\end{gather}
\end{subequations}
As the residual field coincides with the retarded field far from the particle we can use the usual
retarded field boundary conditions when solving Eqs.~\eqref{eq:field-residual-scalar}, \eqref{eq:field-residual-em} and \eqref{eq:field-residual-gravity}. The details of a
numerical implementation of the effective source approach are then much the same as for the Green
function and mode-sum regularization schemes. One can use either a frequency domain or a time domain
method for solving the field equations, the key differences now being that there is no restriction
to $1+1$D, and that the effective source must be included as a source term. A more
thorough review of the effective source approach --- including a detailed description of methods for obtaining explicit expressions for the effective source --- can be found in Refs.~\cite{Wardell:2011gb,Vega:2011wf}.
The effective source approach has been successfully applied in the frequency domain
\cite{Warburton:2013lea}, and in the time domain in $1+1$D \cite{Vega:2007mc}, $2+1$D
\cite{Dolan:2010mt,Dolan:2011dx,Dolan:2012jg} and $3+1$D
\cite{Vega:2009qb,Diener:2011cc,Vega:2013wxa} contexts. It has also been formulated --- but not yet
implemented --- in the gravitational case to second order in the mass ratio \cite{Pound:2014xva,Gralla:2012db};
for the second order problem it is currently the only viable computational strategy. Since the
effects of the second-order metric perturbation will be very small --- being suppressed by two orders
of the mass ratio relative to test body effects --- it is likely that a highly accurate numerical
scheme will be required, suggesting a frequency domain treatment of the problem where one
encounters \emph{ordinary differential equations} (ODEs) which are relatively easy to solve
numerically to high accuracy.
\section{Evolution schemes}
\label{sec:evolution}
The calculation of the self-force is only the first stage in the production of an inspiral model.
Another critical component is a scheme for evolving the orbit using this self-force information.
Various approaches have been proposed, each of which brings with it its own advantages and
disadvantages. Approaches which make approximations in the self-force used to drive the inspiral
can give substantial decreases in the computational cost of an inspiral simulation, but come at the
cost of ignoring potentially relevant physical effects.
\subsection{Dissipation driven inspirals}
A straightforward model for the inspiral can be obtained from energy balance considerations. Using
a flux calculation --- in which fluxes of energy and angular momentum are obtained by evaluating
the point-particle retarded field near the horizon and out at infinity -- the entire issue of
regularization is avoided and one obtains an approximation to the contributions to the inspiral
coming from dissipative self-force effects. However, this is inadequate for capturing all relevant
effects from the first-order self-force. Being a dissipative approximation, it completely misses all
conservative corrections to the motion. Furthermore, there is no well-defined way of associating
fluxes far from the worldline with an instantaneous local self-force on the worldline; a flux model
can only be used to drive an inspiral in a time-averaged sense, ignoring some potentially important
dissipative contributions. Nevertheless, the relative straightforwardness and computational
efficiency of their implementation have made flux models a compelling approach to assessing
qualitative features of self-force driven inspirals. These ``kludge'' models have been used to
produce kludge waveforms for EMRI systems which capture at least some of the
qualitative physical effects
\cite{Glampedakis:2002ya,Hughes:2005qb,Drasco:2005kz,Sundararajan:2007jg,Sundararajan:2008zm}.
\begin{figure}[htb!]
\begin{center}
\includegraphics[width=6cm]{figs/FrLoops}
\includegraphics[width=6cm]{figs/FtLoops}
\end{center}
\caption{
Radial (left) and time (right) components of the self-force through one radial
cycle, for three different geodesic orbits in Schwarzschild spacetime. Solid lines
indicate the full self-force and dashed lines indicate the conservative-only (left)
or dissipative-only (right) pieces. Arrows denote the direction of the geodesic
motion. Reproduced from Ref. \cite{Vega:2013wxa}.
}
\label{fig:loops}
\end{figure}
To improve on the flux-based dissipative model, one can instead make use of a
half-retarded minus half-advanced scheme \cite{Mino:2003yg,Gralla:2005et,Hinderer:2008dm,Flanagan:2012kg} to
compute a local, instantaneous dissipative self-force (see Fig.~\ref{fig:loops}).
This approach captures all dissipative effects responsible for driving the
inspiral, but still neglects small but potentially important
conservative corrections to the orbital phase of the system \cite{Warburton:2011fk}.
\subsection{Osculating self-forced geodesics}
In order to account for both conservative and dissipative effects from the first order self-force, it is essential to build an
inspiral model around a full calculation of the local self-force. This, however, still leaves some
flexibility in the choice of method for computing this local self-force. One option, based on the
osculating geodesics framework \cite{Pound:2007th} is particularly compelling as it allows for dramatic
improvements in computational efficiency by separating the coupled problem of simultaneously
solving for the orbit and for the regularized retarded field into two independent, largely
uncoupled computational problems. The basic idea is to perturbatively expand the worldline about a
geodesic of the background spacetime, $z(\tau) = z_0(\tau) + \cdots$. Then, the self-force at first order is a
functional not of the evolving orbit $z(\tau)$, but of the geodesic orbit, $z_0(\tau)$, which is
instantaneously tangent to the worldline. This expansion is valid in the adiabatic regime, where
the orbit is evolving slowly and the difference between the geodesic and evolving orbits is small;
the error introduced by the approximation appears at the same order in the equations of motion
as a second-order perturbative correction to the field equations.
The improvements in computational efficiency brought about by the osculating geodesics
approximation are dramatic. The problem of self-consistently computing the regularized self-force
coupled to an arbitrarily evolving orbit is reduced to the much simpler problem of determining the
self-force for geodesic worldlines. Since orbits of black hole spacetimes are parametrised by at
most three conserved quantities (energy and angular momentum in the Schwarzschild case,
supplemented by the Carter constant in the Kerr case), it is computationally tractable to span the
entire parameter space of moderately-eccentric geodesic orbits. Even better, the
highly-accurate frequency domain mode-sum method can be used because bound geodesic orbits are
efficiently represented by a small frequency spectrum. In Ref.~\cite{Warburton:2011fk} this approach was explored
over an entire radiation-reaction timescale for moderately-eccentric inspirals in the Schwarzschild
spacetime\footnote{See also Ref.~\cite{Lackeos:2012de} for an approximate version which assumes a sequence of
quasi-circular orbits.}. This was achieved by tabulating the values for the self-force in a relevant portion of
the energy-angular momentum phase space of geodesic orbits, and using an interpolated model of the
tabulated results to drive an orbital inspiral.
\subsection{Self-consistent evolution}
Unfortunately, the adiabatic approximation responsible for the dramatic improvements in
computational efficiency brought by the osculating geodesics framework also introduces errors in the equation of motion at
second perturbative order, making the method inadequate for the purposes of precision EMRI astrophysics.
One possible solution, implemented in \cite{Diener:2011cc} for the
scalar field case, is to avoid expanding out the worldline and instead evolve the self-consistent
coupled system of equations, Eqs.~\eqref{eq:field-gravity} and \eqref{eq:accel-gravity};
\eqref{eq:field-scalar}, \eqref{eq:accel-scalar} and \eqref{eq:mdot-scalar}; or
\eqref{eq:field-em} and \eqref{eq:accel-em}. This is a computationally much more difficult problem as the field
equations must be solved in the time domain and one cannot rely on an efficient off-line tabulation
of self-force values. This self-consistent evolution can in principal be implemented using a
time-domain mode-sum scheme, but in practice implementations have instead used the effective source
approach since that will be necessary at second perturbative order (see Sec.~\ref{sec:msr} for an explanation of this issue).
While a self-consistent evolution incorporates all effects contributing to the first order self-force, and only
neglects second order contributions from the second order field equations, this comes at the cost
of computational efficiency. Whereas an osculating geodesics framework can evolve a large number
($\sim 10,000$) of orbits with ease once the initial off-line tabulation phase is complete
\cite{Warburton:2011fk}, the
self-consistent scheme requires a long calculation of the solutions of the first-order field
equations for each new orbit. In reality, this limits the practicality of the scheme to tens of
orbits using current methods. The self-consistent evolution scheme is therefore most valuable as
a benchmark against which other, less accurate but more efficient methods can be validated. In fact,
comparisons between the self-consistent and osculating geodesics scheme for the scalar case indicate
that the osculating geodesics scheme performs remarkably well \cite{Warburton:Capra17}.
\subsection{Two-timescale expansions}
The osculating geodesics scheme is fast, but inaccurate, while the self-consistent evolution scheme
is accurate, but slow. This begs the question of whether there is a middle ground which
incorporates most of the accuracy of a self-consistent evolution while maintaining the
computational efficiency of the osculating geodesics scheme. One promising possibility is that the
use of a two-timescale expansion could give just such a scheme. In the two-timescale scheme, rather
than using an expansion about a background geodesic (which is valid over short timescales
characterised by the orbital motion), one introduces an additional radiation reaction timescale
into the problem and incorporates the relevant effects over this radiation reaction timescale.
The relevant two-timescale expansion of the worldline equations of motion was completed in
Ref.~\cite{Hinderer:2008dm}, and follow-up work has improved our understanding of
important resonance effects for inspiral orbits
\cite{Flanagan:2010cd,Hirata:2010xn,Grossman:2011ps,Grossman:2011im,Gair:2011mr,Flanagan:2012kg,Isoyama:2013yor,Brink:2013nna,Ruangsri:2013hra,vandeMeent:2013sza,vandeMeent:2014raa}.
In order to consistently incorporate a two-timescale expansion into an orbital evolution scheme,
it will also be necessary to have a two-time expansion of the field equations. Promising progress
towards such an expansion was recently reported in \cite{Moxon:Capra17}, indicating that it is likely
that a self-consistent orbital evolution using a two-timescale expansion is indeed feasible.
\section{Discussion}
The techniques described in this review represent the current state of the art of self-force
calculations. The three primary approaches: worldline convolutions, mode-sum regularization and the
effective source approach can be considered complimentary, with each having regimes where they are
most appropriate:
\begin{itemize}
\item The mode-sum scheme gives unparalleled accuracy (particularly in the frequency domain) for
orbits which have a small frequency spectrum.
\item The worldline convolution method provides valuable physical insight and can be easily
applied to arbitrary orbital configurations, including those inaccessible by other means.
\item The effective source approach can be used in arbitrary spacetimes without relying on any
symmetries, and also stands out as the most applicable to a second order calculation.
\end{itemize}
There still remain important developments to be made in each case. For example, the mode sum and
effective source schemes
are still under development for the Kerr gravitational case, and worldline convolution
approaches have yet to be fully applied to the gravitational case for any spacetime.
Finally, it should be pointed out that this review is not an exhaustive exposition of all self-force
computation strategies. Many other calculations have not been described, including alternative
regularization strategies \cite{Rosenthal:2003qr,Rosenthal:2004wp,Kol:2013tfa},
near-horizon waveform calculations \cite{d'Ambrosi:2014iga,Hadar:2009ip,Hadar:2011vj,Hadar:2014dpa}, methods
based on effective field theory \cite{Birnholtz:2013nta,Birnholtz:2013ffa,Birnholtz:2014fwa,Birnholtz:2014gna}, analytic
calculations in particularly simple cases \cite{Wiseman:2000rm,Cho:2007jj,Ottewill:2012aj}, black holes in higher
dimensions \cite{Beach:2014aba}, and calculations in
non-black hole spacetimes such as wormholes \cite{Taylor:2012mv,Taylor:2014vsa} and cosmological
models \cite{Burko:2002ge,Haas:2004kw}. More details can be found in the reviews
\cite{Poisson:2011nh,Detweiler:2005kq}, and references therein.
\section*{Acknowledgements}
I am grateful to Scott Field and David Nichols for helpful discussions during the preparation of
this article, to Peter Diener for providing the data used in Fig.~\ref{fig:src-smoothness},
and to Niels Warburton, Scott Field and Eanna Flanagan for helpful comments on an early draft of this paper.
This material is based upon work supported by the National Science Foundation
under Grant Number 1417132. B.W. was supported by Science Foundation Ireland
under Grant No. 10/RFP/PHY2847, by the John Templeton Foundation New Frontiers
Program under Grant No. 37426 (University of Chicago) - FP050136-B (Cornell
University), and by the Irish Research Council, which is funded under the
National Development Plan for Ireland.
\bibliographystyle{unsrt}
|
\section{Preliminary estimates}\label{Walsh}
We denote by $\mathcal {P}_{\Pi}$ the probability measure on $\widehat \Omega =\{\pm 1\}^{\mathbb{Z}^{d}\times\mathbb{Z}_{+}}$ generated by the initial distribution $\Pi$, and by $\frak {M}_{t_0}^{t_1}$, $0\leq t_{0} \leq t_{1}$ the $\sigma$-algebra of subsets of $\widehat \Omega$ generated by $\{\eta_{t}\}_{t=t_{0}}^{t_{1}}$. As $\Pi$ is invariant, $\mathcal {P}_{\Pi}$ is invariant under the time shift.\par\smallskip
We consider functionals $f$ which depend only on
the values of the field at the origin, i.e., on the sequence of random variables $\{\eta_{t}(0)\}_{t=0}^{\infty}$.
We set for brevity $\zeta_t=\eta_t(0)$ and $\widehat \zeta = \{ \zeta_t: t\in \mathbb{Z}_+\}\in \Omega_{+} = \{\pm 1\}^{\mathbb{Z}_+}$. $\mathcal {M}_{t_0}^{t_1}$, $0\leq t_0 < t_1$ will denote the $\sigma$-algebra generated by the variables $\{\eta_t(0)\}_{t=t_0}^{t_1}$, which is a subalgebra of $\frak {M}_{t_0}^{t_1}$.\par\smallskip
By
abuse of notation, $f(\widehat \zeta)$ may denote a function on $\widehat \Omega$ or on $\Omega_{+}$, according to the circumstances, and similarly for the $\sigma$-algebras $\mathcal {M}_{t_0}^{t_1}$, $0\leq t_0 < t_1$. We also write $\mathcal{M}_{t}$ and $\frak M_{t}$ for $\mathcal{M}_{t}^{t}$ and $\frak M_{t}^{t}$, respectively.
\par\smallskip
In what follows if $f$ is a function on $\widehat \Omega$ we introduce the notation $\langle f(\cdot)\vert \frak M_{0}\rangle(\eta) = G^{(f)}(\eta)$, $\eta \in \Omega$. The following lemma is a simple consequence of Theorem \ref{base}.
\begin{lemma}\label{lemma1} Let $f(\widehat \zeta)$ be a cylinder function on $\Omega_{+}$, depending only on the variables $\zeta_{0},\ldots, \zeta_{m-1}$, $m\geq 1$. Then $ G^{(f)}(\eta) \in \mathcal{H}_{M}$ and
\begin{equation} \label{agg} \left \| G^{(f)} \right \|_{M} \leq \; C \; \max_{\gamma \in \{0,\ldots, m-1\}} |f_{\gamma}| (1+\mu_{*})^{m},
\end{equation}
where $\mu_{*} = M \sqrt {\bar \mu (1+ 2 \bar \mu)}$ and $C>0$ is a constant.
\end{lemma}
\begin{proof} \label{Proof.} As $f$ depends only on $\zeta_{0},\ldots, \zeta_{m-1}$ it can be written in the form
\begin{equation}\label{expansion} f(\widehat \zeta) = \sum_{\gamma\subset \{0, \ldots, m-1\}} f_{\gamma} \Psi_{\gamma}(\widehat \zeta) \end{equation} where the sum runs over the subsets of $\{0, \ldots, m-1\}$, and the functions
\begin{equation}\label{basis2} \Psi_\gamma (\widehat \zeta) = \prod_{t\in \gamma} \zeta_t , \quad \gamma \neq \emptyset, \qquad \Psi_\emptyset(\widehat \zeta) =1 \end{equation}
are called ``Walsh functions''. The first assertion follows from the fact that for any subset $\gamma = \{t_{0}, t_{1}, \ldots, t_{k}\} \subset \mathbb{Z}_{+}$, $t_{0}< t_{1}< \ldots < t_{k}$,
we have
\begin{equation} \label{condprob} G_\gamma(\bar \eta) := \left \langle \Psi_\gamma | \frak M_{t_0} \right \rangle \in \mathcal{H}_{M}, \qquad \bar \eta \in \Omega. \end{equation}
In fact, if $r_{j}= t_{k-1-j}-t_{k-j}$, $j=1, \ldots, k $, $G_\gamma$ can be written as
\begin{equation}\label{A2} G_{\gamma}(\bar \eta) = \Phi_{\{0\}}(\bar \eta) \left [ \mathcal{T}^{r_{k}} \Phi_{\{0\}}\ldots \mathcal{T}^{r_{1}} \Phi_{\{0\}}\right ](\bar \eta), \quad \bar \eta \in \Omega, \end{equation}
i.e., $G_\gamma$ is obtained by successive applications of $\mathcal{T}$ and of the multiplication operator by $\Phi_{\{0\}}$. As both operations leave $\mathcal{H}_M$ invariant, $G_{\gamma}\in \mathcal H_{M}$.
\par\smallskip
Moreover the following inequality is proved in the Appendix
\begin{equation} \label{lemmadue} \ \|G_\gamma\|_M \; \leq \; M^{|\gamma|}\; \bar \mu^{ [{|\gamma|\over 2}]} (1+ 2 \bar \mu)^{[{|\gamma|-1\over 2}]} \leq C \; \mu_{*}^{|\gamma|},\end{equation} where $[\cdot]$ denotes the integer part, and $C>0$ is a constant which is easily worked out. \par\smallskip
The proof of the lemma follows by observing that the inequality \eqref{lemmadue} implies
\begin{equation} \label{agg1} \left \| \langle f(\cdot)|\frak M_{0}\rangle \right \|_{M} \leq \; C \; \max_{\gamma \in \{0,\ldots, m-1\}} |f_{\gamma}| \sum_{\gamma\subset \{0, \ldots, m-1\}} \mu_{*}^{|\gamma|}.
\end{equation}
\end{proof}
We denote by $\wp$ the probability measure induced by $\mathcal {P}_{\Pi}$ on $\Omega_{+}$. $\wp$ is stationary with respect to the time shift on $\Omega_+$: $ S\widehat \zeta = \{\zeta_1, \zeta_2, \ldots \}$.\par\smallskip
The following assertion is a simple consequence of the previous lemma.
\begin{lemma}\label{lemma1b} Under the assumptions of the previous lemma, if $\bar \mu$ is so small that $\mu_{*} <1$, then the probability measure $\wp$ on $\Omega_+$ is continuous.
\end{lemma}
\begin{proof} \label{Proof.} We need to prove that any point $\widehat \zeta^{(0)} =\{\bar \zeta_{k}\}_{k=0}^{\infty}\in \Omega_{+}$ has zero $\wp$-measure. Consider the cylinders $Z_n(\widehat \zeta^{(0)}) = \{\zeta_{j} = \bar \zeta_{j}:\; j=0,1,\ldots, n-1\}$, which are decreasing $Z_{n+1}(\widehat \zeta^{(0)})\subset Z_n(\widehat \zeta^{(0)})$ and such that $\cap_{n} Z_{n}(\widehat \zeta^{(0)}) =\{\widehat \zeta^{(0)}\}$. The probabilities \begin{equation}\label{probz} \wp \left ( Z_n(\widehat \zeta^{(0)}) \right ) = {1\over 2^n} \left \langle \prod_{j=0}^{n-1} \left ( 1 + \bar \zeta_j \zeta_j \right ) \right \rangle_\wp \end{equation}
are computed by expanding the internal product in terms of the functions $\Psi_\gamma$:
$$ \prod_{j=0}^{n-1} \left ( 1 + \bar \zeta_j \zeta_j \right ) = \sum_{\gamma \subset \{0, \ldots, n-1\}} \Psi_{\gamma}(\widehat {\bar \zeta} ) \Psi_{\gamma}(\widehat \zeta), \qquad \widehat {\bar \zeta}= \{\bar \zeta_{j}\}_{j=0}^{n-1}.$$
Recalling that $|\Psi_{\gamma}(\widehat \zeta)|=1$, we have
$$\left | \left \langle \sum_{\gamma \subset \{0, \ldots, n-1\}} \Psi_{\gamma}(\widehat {\bar \zeta} ) \Psi_{\gamma}(\widehat \zeta) \right \rangle_{\wp}\right |\leq \sum_{\gamma \subset \{0, \ldots, n-1\}} \left |\left \langle \Psi_{\gamma}(\widehat \zeta) \right \rangle_{\wp} \right | =$$
$$ = \sum_{\gamma \subset \{0, \ldots, n-1\}} \left | \left \langle \left \langle \Psi_{\gamma}|\frak M_{0}\right \rangle(\cdot) \right \rangle_{\Pi}\right | = \sum_{\gamma \subset \{0, \ldots, n-1\}} \left | \left \langle G_{\gamma}(\cdot)\right \rangle_{\Pi} \right | . $$
Therefore by the inequality \eqref{lemmadue} the right side is bounded by
$${C\over 2^{n}} \sum_{\gamma \subset \{0, \ldots, n-1\}} \mu_{*}^{|\gamma|} = C \left ({1+\mu_{*}\over 2 }\right )^{n}.$$
Hence if $\mu_{*} <1$, the right side tends to $0$ as $n\to\infty$, which proves the lemma.
\end{proof}
From now on we assume that $\mu_{*}<1$.
\par\smallskip
We pass to consider functions for which the expansion \eqref{expansion} is infinite, i.e., $\gamma$ runs over the collection
$\frak g$ of the finite subsets $\mathbb{Z}_{+}$. The functions $\{\Psi_{\gamma}: \gamma \in \frak g\}$, are an orthonormal basis in $L_2(\Omega_+, \wp_0)$, where $\wp_0 = \pi^{\mathbb{Z}_+} $ is the probability measure on $\Omega_+$ corresponding to the random variables $\{\zeta_k\}_{{k=0}}^{\infty}$ being i.i.d. with distribution $\pi(\pm 1) ={1\over 2}$. The corresponding series is called ``Fourier-Walsh expansion'' \cite{Fine}. \par
\smallskip
A map $\mathcal{F}: \Omega_+\to T^1$, where $T^1 =[0, 1) \mod 1$ is the one-dimensional torus, is defined by associating to a point $\widehat \zeta\in \Omega_+$ the binary expansion $x =0,a_0 a_1\ldots\in [0, 1]$, with $a_t = {1-\zeta_t\over 2} $, $t\in \mathbb{Z}_+$.
$\mathcal{F}$ is not invertible because the dyadic points of $T^{1}$ have two binary expansions, but it becomes invertible if we exclude the sequences such that $\zeta_{t}= -1$ for all $t$ large enough. Such sequences are a countable set, which has zero $\wp_0$-measure, and also, by Lemma \ref{lemma1b}, zero $\wp$-measure.
\par\smallskip
Under the map $\mathcal{F}$ the basis functions $\Psi_{\gamma}$ go into the functions $$\psi_{\gamma}(x) = \prod_{t\in \gamma} \phi_{t}(x), \qquad \gamma \in \frak g .$$
where $\phi_{t}(x)$ is the image of $\zeta_{t}$, $t\in \mathbb{Z}_{+}$, i.e., $$\phi_{0}(x) =\left\{
\begin{array}{ll}
1, & 0\leq x < {1\over 2}\\
-1, & {1\over 2}\leq x < 1
\end{array}\right. $$
and for $t\geq 0$, $\phi_{t}(x) = \phi_{0}(2^{t} x)$, where $2^{t} x$ is understood $\mod 1$. \par\smallskip
If $f\in L_2(\Omega_{+}, \wp_{0})$ then $\tilde f(x) = f (\mathcal{F}^{-1} x) \in L^{2}(T^1, dx)$ and can be expanded in the orthonormal basis $\{\psi_\gamma: \gamma \in \frak g\}$, with coefficients
\begin{equation} \label{coeff} f_{\gamma} = \int_{\Omega_{+}} f(\widehat \zeta) d\wp_{0}(\widehat \zeta) = \int_{0}^{1} \tilde f(x) \psi_{\gamma}(x) dx .\end{equation}
\par\smallskip
A natural way of ordering the collection $\frak g$ of the finite subsets of $\mathbb{Z}_{+}$, which plays an important role in the theory, is obtained by setting $\gamma_0=\emptyset$ and $\gamma_n = \{ t_1, t_2, \ldots, t_r\}$, where $r$ and $0 \leq t_1< t_2 < \ldots < t_r$ are uniquely defined by the relation $ n = 2^{t_1} + \ldots + 2^{t_r} $.
We call Walsh series both the expansion
\begin{equation} \label{f1} f(\widehat \zeta) =\sum_{\gamma \in \frak g} f_{\gamma} \Psi_{\gamma}(\widehat \zeta) = \sum_{n=0}^\infty f_{\gamma_n} \Psi_{\gamma_n}(\widehat \zeta), \end{equation}and the corresponding expansions of $\tilde f(x)$.
For the latter, an important role is played by a particular set of partial sums
\begin{equation} \label{ridotta} \Sigma_{2^{k}}(\tilde f; x) = \sum_{\gamma \subset \{0, 1, \ldots, k-1\}} f_{\gamma} \psi_{\gamma}(x) = \sum_{n=0}^{2^{k}-1} f_{\gamma_{n}} \psi_{\gamma_{n}}(x) \end{equation}
for which it can be seen \cite{Fine} that
\begin{equation}\label{ridotta} \Sigma_{2^{k}}(\tilde f; x) = 2^{k} \int_{\alpha_{k}}^{\beta_{k}} \tilde f(y) dy, \qquad \alpha_k = m\; 2^{-k}, \; \beta_k = (m+1) 2^{-k} \end{equation}
where the integer $m$ is such that $\alpha_{k} \leq x < \beta_{k}$.
\par\smallskip
The following result is proved in \cite{Fine}. We repeat it here, with a shorter proof based on conditional probabilities.\par
\begin{lemma}\label{lemmaf} Let $\tilde f(x)$ be a bounded function. Then its Walsh-Fourier coefficients $f_{\gamma}$, given by \eqref{coeff}, satisfy the following inequality
\begin{equation}\label{stimacoeff} \left | f_{\gamma} \right | \; \leq \; {\omega(\tilde f; 2^{-n-1})\over 2^{n+2}}, \qquad n = max\{ t: t\in \gamma \}, \end{equation}
where $\omega(f;\delta)$ is the modulus of continuity of $\tilde f$:
\begin{equation}\label{coeffcont} \omega(f;\delta) = \sup_{x, x\prime \in T^1\atop |x-x^{\prime}|=\delta} {|f(x) - f(x^{\prime})|\over \delta}. \end{equation} \end{lemma}
\par\noindent
\begin{proof}[Proof.] We have
$$ f_\gamma = \left \langle f(\widehat \zeta) \prod_{t\in \gamma} \zeta_t \right \rangle_{\wp_0} = \left \langle \prod_{t\in \gamma\setminus \{n\}} \zeta_t \; \; \left \langle f(\widehat \zeta) \zeta_n \vert \mathcal{M}_{0}^{n-1}\right \rangle \right \rangle_{\wp_0} .$$
Going back to $T^{1}$, and setting $x_{n} = {a_{0}\over 2}+ \ldots + {a_{n-1}\over 2^{n}}$, $a_{j}={1-\zeta_{j}\over 2}$, we have
$$\left | \left \langle f(\widehat \zeta) \zeta_n \vert \mathcal{M}_{0}^{n}\right \rangle \right | = 2^{n} \int_{x_{n}}^{x_{n}+ 2^{-n}} \tilde f(x) \left ( 1 - 2\phi_{n}(x) \right ) dx= $$
\begin{equation}\label{decay} = 2^{n} \int_{x_{n}}^{x_{n}+ 2^{-n-1}} \left [ \tilde f(x) - \tilde f(x + 2^{-n-1}) \right ] dx , \end{equation}
from which, taking into account \eqref{coeffcont}, the inequality \eqref{stimacoeff} follows immediately. \end{proof}
\par\smallskip
The results above allow us to prove the analogue of Lemma \ref{lemma1} for functions $f$ such that $\tilde f (x) = f(\mathcal{F}^{-1}x)$ is H\"older continuous: $\tilde f \in \mathcal{C}^{\alpha}(T^{1})$, $\alpha\in (0,1)$. In what follows if $g\in \mathcal{C}^\alpha(T^1)$ we denote by $\|g\|_{\mathcal{C}^{\alpha}}$ the norm and by $\|g \|_\alpha$ the semi--norm
\begin{equation*}
\|g\|_{\alpha}=\sup_{x,y\in T^1}\frac{|g(x)-g(y)|}{|x-y|^\alpha}.
\end{equation*}
\begin{lemma}\label{lemma:hoelder}
Let $f$ be a function on $\Omega_+$, such that $\tilde f\in \mathcal C^\alpha(T^1)$, $\alpha \in (0, 1)$. If $\bar \mu$ is so small that $\kappa := 2^{-\alpha}(1+\mu_*) < 1$, then
$G^{(f)} \in \mathcal{H}_M$ and the following inequality holds
\begin{equation}\label{norma}\left \| G^{(f)} \right \|_M \leq {C_{\a}\over 1- \kappa} \|\tilde f\|_{\mathcal{C}^{\alpha}} , \end{equation} where $C_{\a}>0$ is a positive constant.
\end{lemma} \par\smallskip
\begin{proof}\label {Proof.} If $2^k \leq n < 2^{k+1}$ the Fourier coefficient $\gamma_{n}$ in the Walsh series \eqref{f1} is such that $\max \{t\in \gamma_n\} = k$. Hence, as $\delta \; \omega(\tilde f; \delta) \leq \delta^\alpha \|\tilde f\|_\alpha $, the inequality \eqref{stimacoeff} gives
\begin{equation}\label{correz} \left | f_{\gamma_n} \right | \leq { \|\tilde f\|_\alpha\over 2^{1+\a}} 2^{- k \alpha}, \qquad 2^k \leq n < 2^{k+1}.\end{equation}
Therefore we have
$$ \left \| \sum_{n= 2^k}^{2^{k+1}-1} f_{\gamma_n} \left \langle \Psi_{\gamma_n}|\frak M_0\right \rangle \right \|_M \leq \frac{||\tilde f||_{\a}}{2^{1+\a}} 2^{- k \alpha} \sum_{n= 2^k}^{2^{k+1}-1} \left \| \left \langle \Psi_{\gamma_n}|\frak M_0\right \rangle \right \|_M .$$
Observe moreover that the number of elements of $\gamma_n$ is $r_n = |\gamma_n| = u_n -1$ where $u_n$ is the number of "$1$" in the binary expansion of $n$. Hence, by the inequality \eqref{lemmadue} we find
$$ \sum_{n= 2^k}^{2^{k+1}-1} \left \| \left \langle \Psi_{\gamma_n}|\frak M_0\right \rangle \right \|_M \leq C \sum_{s=0}^{k-1} { k-1 \choose s} \mu_*^s = C (1+ \mu_*)^{k-1} ,$$
which, as $|f_{\emptyset}|\leq \|f\|_{\infty}$, together with \eqref{correz}, implies \eqref{norma}. \end{proof}
\par\smallskip
\section{Weak dependence and the Central Limit Theorem}\label{WDCLT}
In the present paragraph we prove our main results for sums of sequences of the type
$f(S^{t}\widehat \zeta)$, $t=0, 1,\ldots$. As $\mathcal P_{\Pi}$ and the measure $\wp$ induced by it on $\Omega_{+}$ are invariant under time shift, the sequence is stationary in distribution.\par\smallskip
In what follows $\langle \cdot \rangle$ denotes an average with respect to $\wp$, $\mathcal P_{\Pi}$ or $\Pi$, according to the context. Moreover we denote by $c_{i}, i=1,2,\ldots$, and sometimes by $\rm const$, different constants which depend on the parameters of the model.
\par\smallskip
Let $f$ be a bounded measurable function on $\Omega_{+}$ with $\langle f\rangle_{\wp}=0$, and
\begin{equation}\label {somma0} S_{n}(\widehat\zeta\vert f) = \sum_{t=0}^{n-1} f(S^{t} \widehat \zeta), \qquad n=1,2,\ldots . \end{equation}
If $f$ admits a Walsh expansion \eqref{f1} then $ \sum_{\gamma \in \frak g} f_{\gamma}\langle \Psi_{\gamma} \rangle_{\wp} = \langle f \rangle_{\wp}=0$, so that
\begin{equation}\label{somma1} f(\widehat \zeta) = \sum_{\gamma\in \frak g,\; \gamma \neq \emptyset} f_{\gamma} \widehat \Psi_{\gamma}(\widehat \zeta), \qquad \widehat \Psi_{\gamma}(\widehat \zeta) = \Psi_{\gamma}(\widehat \zeta) - \left \langle \Psi_{\gamma}( \cdot)\right \rangle_{\wp}. \end{equation}
\par\smallskip
In what follows we make repeated use of the fact that if $f$ is a function on $\widehat \Omega$ and $G^{(f)}:= \langle f(\cdot)\vert \frak M_{0}\rangle\in \mathcal{H}_{M}$, then, by Theorem \ref{base}, $ \langle f(S^{t+h}\cdot)\vert \frak M_{h}\rangle = \mathcal{T}^{t}G^{(f)}\in \mathcal{H}_{M}$. \par\smallskip
\begin{theorem}\label{theorem:clt} Let $f$ be a function on $\Omega_{+}$, depending only on $\zeta_{0}, \ldots,\zeta_{m-1}$, $m\geq 1$, and such that $\langle f \rangle_{\wp} =0$. Then the dispersion of normalized sums
$ S_{n}(\widehat \zeta\vert f)\over \sqrt n$
tends, as $n\to \infty$, to a finite non-negative limit \begin{equation}\label {dispersione} \sigma^{2}_{f} = \left \langle f^{2}(\cdot) \right \rangle_{\wp} + 2 \sum_{t=1}^{\infty} \left \langle f(\cdot) f(S^{t}\cdot) \right \rangle_{\wp} \end{equation}
and the series is absolutely convergent. Moreover, if $\sigma^{2}_{f}>0$, the sequence $S_{n}(\widehat\zeta\vert f)\over \sqrt n$ tends weakly to the centered gaussian distribution with dispersion $\sigma^{2}_{f}$.
\end{theorem}
\begin{proof}\label {Proof.}
The proof of the theorem is based on two basic inequalities.
\begin{equation}\label{ineq1} \left \| \left \langle f(\cdot) f(S^{t}\cdot)\vert \frak M_{0} \right \rangle \right \|_{M} \leq c_{1}\|f\|^{2}_{\infty} \bar \mu^{\max \{0, t-m+1\}} (1+ \mu_{*})^{2m},\end{equation}
\begin{equation} \label{giesse} \left \| G^{(S_{n})}\right \|_{M } \leq c_{2}\; \|f\|_{\infty} \; (1+ \mu_{*})^{m}, \qquad \left \| G^{(\widehat S^{2}_{n})}\right \|_{M } \leq c_{3}\; \|f\|^{2}_{\infty} \;m (1+ \mu_{*})^{m}, \end{equation}
where $\widehat S^{2}_{n}(\widehat \zeta\vert f) = S^{2}_{n}(\widehat \zeta\vert f) - \langle S^{2}_{n}(\cdot\vert f) \rangle $ and $c_{1}, c_{2}, c_{3}$ are constants independent of $m$. \par\smallskip
For the proof of \eqref{ineq1}, observe that if $t\leq m-1$, then $ f^{(2)}_{t}(\widehat \zeta):=f(\widehat \zeta) f(S^{t}(\widehat \zeta))$ is a cylinder function $\mathcal{M}_{0}^{t+m-1}$- measurable and
bounded by $\|f\|^{2}_{\infty}$. Hence, by Lemma \ref{lemma1} and \eqref{coeff}, $\|f^{(2)}_{t}\|_{M}\leq {\rm const} \|f\|^{2}_{\infty}(1+\mu_{*})^{m+t}$, and \eqref{ineq1} holds for $t\leq m-1$.\par
If $t\geq m$ taking the expectation with respect to $\frak M_{0}^{m-1}$ we have
$$ \left \langle f(\cdot) f(S^{t}\cdot)\vert \frak M_{0} \right \rangle = \left \langle f(\widehat \zeta) [\mathcal{T}^{t-m+1} G^{(f)}](\eta_{m-1})\vert \frak M_{0} \right \rangle .$$
As $ G^{(f)} \in \widehat \mathcal{H}_{M}$, expanding $f$ in Walsh series and using Proposition \ref{proposition} in the Appendix and Lemma \ref{lemma1} we see that Inequality \eqref{ineq1} also holds for $t\geq m-1$.
\par\smallskip
The first inequality in \eqref{giesse} is a simple consequence of the inequality $\|\langle f(S^{t}\cdot)\vert \frak M_{0}\|_{M} =\| \mathcal{T}^{t}G^{(f)}\|_{M}\leq \bar \mu^{t} \|G^{(f)}\|_{M}$, of Lemma \ref{lemma1} and of the inequality $|f_{\gamma}| \leq \|f\|_{\infty}$ (see \eqref{coeff}). Moreover, setting $\widehat f^{2}(\widehat \zeta) = f^{2}(\widehat \zeta) - \langle f^{2}(\cdot) \rangle $ and $\widehat f^{(2)}_{t}(\widehat \zeta) = f^{(2)}_{t}(\widehat \zeta) - \langle f^{(2)}_{t}(\cdot) \rangle$, we have
\begin{equation}\label{ineq2} \left \|G^{(\widehat S^{2}_{n})}\right \|_{M} \leq \sum_{j=0}^{n-1} \left \| \mathcal{T}^{j} G^{(\widehat f^{2})}\right \|_{M} + 2 \sum_{j=0}^{n-2} \sum_{t=1}^{n-j-1} \left \| \mathcal{T}^{j} G^{(\widehat f_{t}^{2})}\right \|_{M}, \end{equation}
and the second inequality in \eqref{giesse} follows by observing that $\widehat f^{2}$ is a cylinder function with zero average and $\|\widehat f^{2}\|_{\infty}\leq \|f\|_{\infty}^{2}$, and using the estimate \eqref{ineq1} for $\|G^{(\widehat f_{t}^{2})} \|_{M}$.
\par\smallskip
Passing to the assertions of the theorem, observe first that, by the property \eqref{normsup} of $\mathcal{H}_{M}$, the absolute convergence of the series in \eqref{dispersione} follows
from Inequality \eqref{ineq1}.
\par\smallskip
Assuming that $\sigma_{f}^{2}>0$, for the proof of the CLT we adopt a Bernstein scheme. Let $p_{n} = [n^{\beta}]$, $q_{n} = [n^{\delta}]$, $k_{n} =[{n\over p(n) + q(n)}]$, with $0< \delta < \beta < 1/4$.
The interval of integers $[0, n-1]$ is divided into subintervals of length $p_{n}$ and $q_{n}$: $$I_{\ell} = [(\ell-1)(p_{n}+q_{n}), \; \ell p_{n} +(\ell-1) q_{n}-1] ,\quad J_{\ell}=[ \ell p_{n}+ (\ell-1) q_{n}, \ell (p_{n}+q_{n})-1],$$ $\ell = 1, \ldots k_{n}$, and the rest $J_{*}= [0, n-1]\setminus \cup_{j=1}^{k_{n}} [I_{j}\cup J_{j}]$. \par\smallskip
The sum \eqref{somma0} is then written as $S_{n}(\widehat\zeta\vert f) = S^{(M)}_{n}(\widehat\zeta\vert f) + S^{(R)}_{n}(\widehat\zeta\vert f)$ where
\begin{equation}\label{splits} S^{(M)}_{n}(\widehat\zeta\vert f) = \sum_{\ell=1}^{k_{n}} S_{I_{\ell}}(\widehat\zeta\vert f), \qquad S^{(R)}_{n}(\widehat\zeta\vert f) =
\sum_{\ell=1}^{k_{n}} S_{J_{\ell}}(\widehat\zeta\vert f) + S_{J_{*}}(\widehat\zeta\vert f) \end{equation}
and $S_{I_{\ell}}, S_{J_{\ell}}, S_{J_{*}}$ denote the sums over the corresponding subinterval. \par\smallskip
We first prove that the $L_{2}$-norm of $ S^{(R)}_{n}/\sqrt n$ vanishes as $n\to \infty$, i.e.,
\begin{equation}\label{estim1} \left \langle \left ( \sum_{\ell=1}^{k_{n}} S_{J_{\ell}}(\cdot\vert f) \right )^{2} \right \rangle = k_{n} \left \langle S^{2}_{J_{1}}(\cdot\vert f) \right\rangle + 2 \sum_{1\leq s < t \leq k_{n}} \left \langle S_{J_{s}}(\cdot|f) S_{J_{t}}(\cdot|f) \right \rangle = \mathcal{O} (n^{1-\beta +\delta}).\end{equation}
For the proof, observe that Inequality \eqref{ineq1} implies that $\langle S^{2}_{J_{1}}(\cdot|f) \rangle = \mathcal{O}(q_{n})$, so that the first term on the right of \eqref{estim1} is of the order $k_{n} q_{n} \sim n^{1-\beta +\delta}$. \par
For the second term, observe that, by translation invariance, recalling that $S_{q_{n}}(\widehat \zeta\vert f)$ is $\frak M_{0}^{q_{n}+m-2}$-measurable and taking the corresponding conditional probability, \begin{equation}\label{elle}\langle S_{J_{s}}(\cdot|f) S_{J_{t}}(\cdot|f)\vert \frak M_{0} \rangle = \left \langle S_{q_{n}}(\cdot|f) \left [ \mathcal{T}^{(t-s)\ell_{n} + p_{n}-m +2} \; G^{(S_{q_{n}})}\right ](\eta_{q_{n}-m +2})\vert \frak M_{0} \right \rangle ,\end{equation}
where $\ell_{n}= p_{n}+q_{n}$.
Therefore, recalling the inequalities \eqref{normsup}, we get the estimate
\begin{equation}\label{ineq3} \left | \left \langle S_{J_{s}}(\cdot|f) S_{J_{t}}(\cdot|f) \right \rangle \right | \leq {\rm const} \; (1+\mu_{*})^{m} \; \|f\|^{2}_{\infty} \; q_{n} \bar \mu^{(t-s) \ell_{n}+ p_{n}- m}.\end{equation}
As $k_{n} q_{n} \bar \mu^{p_{n}} \leq {\rm const} \; \bar \mu ^{p_{n}\over 2}$, the double sum on the right of \eqref{estim1} is of the order $\mathcal{O}(\bar \mu^{p_{n}\over 2})$, so that \eqref{estim1} is proved.\par \smallskip
As for $S_{J_{*}}$, \eqref{ineq1} implies $\langle S^{2}_{J_{*}}(\cdot|f)\rangle\leq \langle S^{2}_{p_{n}+q_{n}}(\cdot|f)\rangle= \mathcal{O}(n^{-1+\beta})$. This fact, together with \eqref{estim1}, proves that $\langle (S^{(R)}_{n}(\widehat\zeta\vert f))^{2}\rangle /n = \mathcal{O}(n^{-\beta+\delta})$, and, as $\beta >\delta$, $S^{(R)}_{n}$ does not contribute to the limiting distribution.
\par\smallskip
We now show that the random variables $\{S_{I_{\ell}}\}_{\ell=1}^{k_{n}}$ are almost independent for large $n$, i.e., for the characteristic functions $\phi^{(\ell)}_{n}(\lambda|\widehat \zeta)= \exp\{i{\lambda\over \sqrt n}S_{I_{\ell}}(\widehat \zeta|f)\}$ we have
\begin{equation}\label{indep} \left \langle \prod_{\ell=1}^{k_{n}} \phi^{(\ell)}_{n}(\lambda|\widehat \zeta) \right \rangle - \prod_{\ell=1}^{k_{n}} \left \langle \phi^{(\ell)}_{n}(\lambda|\widehat \zeta)
\right \rangle \to 0, \qquad n\to \infty.\end{equation}
We proceed by iteration. As a first step we consider the difference
$$ \left \langle \prod_{\ell=1}^{k_{n}} \phi^{(\ell)}_{n}(\lambda|\widehat \zeta) \right \rangle - \left \langle \prod_{\ell=1}^{k_{n}-1} \phi^{(\ell)}_{n}(\lambda|\widehat \zeta) \right \rangle \left \langle \phi^{(k_{n})}_{n}(\lambda|\widehat \zeta) \right \rangle = $$
\begin{equation}\label{quasind} =\left \langle \prod_{\ell=1}^{k_{n}-1} \phi^{(\ell)}_{n}(\lambda|\widehat \zeta) \; \widehat \phi^{(k_{n})}_{n}(\lambda|\widehat \zeta) \right \rangle , \qquad \widehat \phi^{(\ell)}_{n}(\lambda|\widehat \zeta) = \phi^{(\ell)}_{n}(\lambda|\widehat \zeta) - \left \langle \phi^{(\ell)}_{n}(\lambda|\widehat \zeta) \right \rangle. \end{equation}
We expand $\widehat \phi^{(k_{n})}_{n}(\lambda|\widehat \zeta)$ in Taylor series at $\lambda=0$, we have, for some $\lambda_{*}$, $|\lambda_{*}| \leq |\lambda|$,
\begin{equation}\label{3exp} \widehat \phi^{(k_{n})}_{n}(\lambda|\widehat \zeta)= i{\lambda\over \sqrt n} S_{I_{k_{n}}}(\widehat \zeta|f) - {\lambda^{2}\over 2n} \left (S^{2}_{I_{k_{n}}}(\widehat \zeta|f) -\left \langle (S^{2}_{I_{k_{n}}}(\cdot|f) \right \rangle \right ) + i^{3}{\lambda^{3}\over n^{3\over 2} 3!} R_{n}(\lambda_{*}, \widehat \zeta),\end{equation}
\begin{equation} \label{expanding} R_{n}(\lambda_{*}, \widehat \zeta) = S^{3}_{I_{\ell}}(\widehat\zeta\vert f) \exp \{i {\lambda_{*}\over \sqrt n} S_{I_{\ell}}(\widehat\zeta\vert f)\} - \left \langle S^{3}_{I_{\ell}}(\widehat\zeta\vert f) \exp \{i {\lambda_{*}\over \sqrt n} S_{I_{\ell}}(\widehat\zeta\vert f)\} \right \rangle , \end{equation}
Clearly $|R_{n}(\lambda_{*}, \widehat \zeta)|\leq 2
p^{3}_{n} |\lambda|^{3}\|f\|^{3}_{\infty} = \mathcal{O}(n^{3 \beta})$, so that,
as $\beta < 1/4$, we need only consider the first two terms of the expansion \eqref{3exp}.\par\smallskip
The product of the first $k_{n}-1$ factors in the expectation in \eqref{quasind} is measurable with respect to $\frak M^{t_{n}}_{0}$, where $t_{n} = (k_{n}-1) p_{n} + (k_{n}-2) q_{n} +m-2$. Taking the corresponding conditional expectation, by Inequality \eqref{giesse} we get for
the first order term the estimate
\begin{equation}\label{ineq4} \left | \left \langle \prod_{\ell=1}^{k_{n}-1} \phi^{(\ell)}_{n}(\lambda|\widehat \zeta) \left [\mathcal{T}^{q_{n}-m+2} G^{(S_{p_{n}})} \right ](\eta_{t_{n}}) \right \rangle \right | \leq c_{4}\; \bar \mu^{q_{n}-m}\; (1+\mu_{*})^{m} \|f\|_{\infty}. \end{equation}
\par\smallskip
For the second order term, proceeding in the same way, and taking into account the second inequality \eqref{ineq2} we come to the estimate
\begin{equation}\label{ineq4bis} \left | \left \langle \prod_{\ell=1}^{k_{n}-1} \phi^{(\ell)}_{n}(\lambda|\widehat \zeta) \left [\mathcal{T}^{q_{n}-m+2} G^{(\widehat S^{2}_{p_n})} \right ](\eta_{t_{n}}) \right \rangle \right | \leq c_{5}\; \bar \mu^{q_{n}-m}\; m (1+\mu_{*})^{m} \|f\|_{\infty}^{2} . \end{equation}
\par\smallskip
Iterating the procedure for the remaining product $ \langle \prod_{\ell=1}^{k_{n}-1} \phi^{(\ell)}_{n}(\lambda|\widehat \zeta) \rangle$, we see that the quantity on the left of \eqref{indep} is of the order $\mathcal{O}(k_{n} n^{-3({1\over 2}-\beta)}) = \mathcal{O}(n^{-{1\over 2} + 2\beta})$, so that , as $\beta < 1/4$, it vanishes as $n\to \infty$.
\par\smallskip
We are left with a sum $\tilde S^{(M)}_{n}$ of $k(n)$ independent variables distributed as $S_{p_{n}}(\widehat \zeta\vert f)$. The $\log$ of the characteristic function of the corresponding normalized sum is
\begin{equation}\label{charact} k_{n}\; \psi_{n}({\lambda\over \sqrt n} \vert f), \qquad \quad \psi_{n}(\lambda\vert f) = \log \left \langle e^{i {\lambda } S_{q_{n}}^{(f)}(\cdot)}\right \rangle . \end{equation}
Expanding $\psi_{n}$ in Taylor series at $\lambda=0$, we see, in analogy to the proof above, that the third order remainder is of order $\mathcal{O}(n^{-3({1\over2}-\beta)})$, so that it does not contribute to the limit. The first order term vanishes, and we see that
$$\lim_{n\to\infty} k_{n} \psi_{n}({\lambda\over \sqrt n} \vert f) =- {\lambda^{2}\over 2 } \; \lim_{n\to \infty} {k_{n}\over n} \left [ p_{n} \left \langle f^{2}(\cdot) \right \rangle + 2 \sum_{j=0}^{p_{n}-1} \sum_{k=1}^{p_{n}-j-1} \left \langle f(\cdot) f(S^{k})\cdot \right \rangle \right ] .$$
As ${k_{n}p_{n}\over n} \to 1$, the expression on the right tends to $-{\lambda^{2}\over 2} \sigma^{2}_{f}.$
The theorem is proved.
\end{proof}
\begin{theorem}\label{theorem:cltalpha} Let $f$ be a function on $\Omega_{+}$, satisfying the assumptions of Lemma \ref{lemma:hoelder} with $\alpha > 1/2$ and such that $\langle f \rangle_{\wp} = 0$. Then, if $\bar \mu$ is small enough, the dispersion of the normalized sums $ S_{n}(\widehat \zeta\vert f)\over \sqrt n$
tends, as $n\to \infty$, to a finite non-negative limit \begin{equation}\label {dispersione1} \sigma^{2}_{f} = \left \langle f^{2}(\cdot) \right \rangle_{\wp} + 2 \sum_{t=1}^{\infty} \left \langle f(\cdot) f(S^{t}\cdot) \right \rangle_{\wp} \end{equation}
were the series on the right is absolutely convergent. Moreover, if $\sigma^{2}_{f}>0$, the sequence $S^{(f)}_{n}(\widehat \zeta)$ tends weakly to the centered gaussian distribution with dispersion $\sigma^{2}_{f}$.
\end{theorem}
\begin{proof} \label {Proof} The proof repeats the pattern of the previous proof, to which we refer. Inequalities \eqref{ineq1} and \eqref{giesse} are replaced by
\begin{equation} \label{nnjeravv} \left \| \left \langle f(\cdot) f(S^{t}\cdot) |\frak M_{0}\right \rangle\right |_{M} \leq c_{6} \|\tilde f\|_{\mathcal{C}^{\alpha}}^{2} \kappa^{t}, \end{equation}
\begin{equation} \label{giesseb} \left \| G^{(S_{n})}\right \|_{M } \leq c_{7}\; \|\tilde f\|_{\mathcal{C}^{\alpha}}, \qquad \left \| G^{(\widehat S^{2}_{n})}\right \|_{M } \leq c_{8}\; \|\tilde f\|_{\mathcal{C}_{\alpha}}^{2}.\end{equation}
\par\smallskip
The proof of the estimate \eqref{nnjeravv} is deferred to the Appendix. The first inequality \eqref{giesseb} is proved as in the previous theorem, recalling Lemma \ref{lemma:hoelder}. \par\smallskip
The second inequality \eqref{giesseb} follows from Inequality \eqref{ineq2}, observing that $\widehat f^{2}(\mathcal{F}^{-1}\cdot) \in \mathcal{C}^{\alpha}$ and using Inequality \eqref{nnjeravv}.
\par\smallskip
For the estimate \eqref{estim1} observe that \eqref{nnjeravv} again implies that $\langle S^{2}_{J_{1}}(\cdot|f) \rangle = \mathcal{O}(q_{n})$. For the second term on the right of \eqref{estim1}
we need, as in \cite{Ibr71}, that the functions are well approximated by their conditional probabilities on finite $\sigma$-algebras.
This property is provided by the representation \eqref{ridotta} for the partial sums, which gives
\begin{equation} \label{njerav5} \left | f(\widehat \zeta) - \Sigma_{2^{n}}(f; \widehat \zeta ) \right | \leq \|\tilde f \|_{\alpha} 2^{- \alpha n}. \end{equation}
Let $m_{n} =[ {4\over \alpha} \log_{2} n]$, where $[\cdot]$ denotes the integer part. In the expression \eqref{elle}, in the sum $S_{J_{s}}(\widehat \zeta\vert f)$,
we replace the function $f$ by its partial sum $\Sigma_{2^{m_{n}}}$. The corresponding sum is denoted $\tilde S_{J_{s}}$.
By Inequality \eqref{njerav5} we have
$$\left \langle S_{J_{s}}(\cdot|f) S_{J_{t}}(\cdot|f)\vert \frak M_{0} \right \rangle = \left \langle \tilde S_{J_{s}}(\cdot|f) S_{J_{t}}(\cdot|f)\vert \frak M_{0} \right \rangle + \mathcal{O}\left ({q_{n}^{2} / n^{4}}\right ).$$
$\tilde S_{J_{s}}(\cdot|f)$ can be treated as $ S_{J_{s}}(\cdot|f)$ in the previous proof, so that the corresponding conditional expectation is written, if $n$ is so large that $p_{n}> m_{n}$, as
\begin{equation}\label{elleb} \left \langle \tilde S_{q_{n}}(\cdot|f) \left [ \mathcal{T}^{(t-s)\ell_{n} + p_{n}-m_{n} +2} \; G^{(\tilde S_{q_{n}})}\right ](\eta_{q_{n}-m_{n} +2})\vert \frak M_{0} \right \rangle\end{equation} (the tilder in $\tilde S_{q_{n}}$ again denotes that $f$ is replaced by $\Sigma_{2^{m_{n}}}$). By the first inequality \eqref{giesseb}
$$\left | \left \langle \tilde S_{J_{s}}(\cdot|f) S_{J_{t}}(\cdot|f)\vert \frak M_{0} \right \rangle \right | \leq {\rm const} \; \|\tilde f \|_{\mathcal{C}^{\alpha}}^{2} \bar \mu^{(t-s)\ell_{n}+ p_{n}-m_{n}},$$
and, as $k_{n}q_{n} \bar \mu^{p_{n}-m_{n}}\leq {\rm const} \; \bar \mu^{p_{n}\over 2}$, we see that the same \eqref{estim1} holds in this case. The estimate for $S_{I_{*}}$ is obvious, so that the negligibility of $S^{(R)}_{n}$ is proved.
\par \smallskip
Further, we pass to the variables $\tilde S_{I_{\ell}}$, $\ell =1,\ldots, k_{n}$, obtained, as before, by replacing $f$ with the partial sum $\Sigma_{2^{m_{n}}}$. The correction is of order $\mathcal{O}(n^{-3})$, so that it can be neglected. The rest of the proof repeats the previous steps, with the only changes that $m$ is replaced by $m_{n}$ and we use the estimates \eqref{giesseb}.
We omit the obvious details.
\end{proof}
\section{Appendix}
\par\noindent
{\bf Proof of inequality \eqref{lemmadue}.} Observe that, by symmetry with respect to the change of sign $\bar \eta(x)\rightarrow-\bar \eta(x)$, $x\in\mathbb{Z}^d$, the density $v(\bar \eta)$ is even. Moreover any finite trajectory of the Markov chain has the same probability of the trajectory obtained by sign exchange.\par\smallskip
The functions $\Phi_{\Gamma}$ defined by \eqref{besis} are even (odd) for $|\Gamma|$ even (odd). Therefore for $|\Gamma|$ odd we have $\langle \Phi_{\Gamma}\rangle_{\Pi}=0$, and also $\langle \mathcal T^r\Phi_{\Gamma}\rangle_{\Pi}=0$, $r>0$. The functions $\Psi_{\gamma}$ are also even (odd) for $|\gamma|$ even (odd), and for $|\gamma|$ odd $\langle \Psi_{\gamma}\rangle_{\wp} =\langle G_{\gamma}\rangle_{\Pi}=0$.
\par\smallskip
For $|\gamma|$ even we set
\begin{equation}\label{decompo} G_{\gamma} = \langle G_{\gamma} \rangle_{\Pi} + \widehat G_{\gamma}, \qquad \widehat G_{\gamma}\in \widehat \mathcal{H}_{M}. \end{equation}
If $\gamma = \{t_{0}, \ldots, t_{k}\} $, $k\geq 1$, we have, by \eqref{A2},
$ G_{\gamma}(\bar \eta) = \Phi_{\{0\}}(\bar \eta) [ \mathcal{T}^{r_{k}} G_{\gamma\setminus \{t_{0}\}} ](\bar \eta)$.
Therefore, if $|\gamma|\geq 2$ is even we have
\begin{equation}\label{even} \| G_{\gamma}\|_{M} \leq M \bar \mu^{r_{k}} \|G_{\gamma\setminus \{t_{0}\}}\|_{M}, \end{equation}
and if $|\gamma| > 1$ is odd
\begin{equation}\label{odd} \| G_{\gamma}\|_{M} \leq M \left ( | \langle G_{\gamma\setminus \{t_{0}\}}\rangle | + \bar \mu^{r_{k}} \|\widehat G_{\gamma\setminus \{t_{0}\}}\|_{M} \right) \leq M (1 + 2\bar \mu^{r_{k}}) \| G_{\gamma\setminus \{t_{0}\}}\|_{M}, \end{equation}
where in the second inequality we take into account that $| \langle G_{\gamma} \rangle | \leq \| G_{\gamma}\|_{\infty} \leq \|G_{\gamma}\|_{M}$.
\par
\smallskip
For $|\gamma|= 1$, $G_{\{t_{0}\}}(\bar \eta) = \Phi_{\{0\}}(\bar \eta)$ so that $\|G_{\{t_{0}\}}\|_{M} = M$, and for $|\gamma|=2$ we have
$ \|G_\gamma \|_M \leq M \| \mathcal{T}^{r_1} \Phi_{\{0\}} \|_M \leq M^2 \bar \mu^{r_1}$.
Inequalities \eqref{even} and \eqref{odd} imply that
\begin{equation}\label{iterazione} \|G_\gamma\|_M \leq M^{|\gamma|} \prod_{j\; \rm{odd}}\bar \mu^{r_{j}} \prod_{j\; \rm{even}} (2\bar \mu^{r_{j}} +1) \end{equation}
which implies \eqref{lemmadue}. \hfill \hbox{$\sqcap\kern-7pt\sqcup$}
\par\smallskip
The following proposition is a simple consequence of the previous proof. \par
\begin{proposition} \label{proposition} Under the assumptions of Lemma \ref{lemma1}, if $\gamma = \{t_0, \ldots, t_k\}$, $G\in \widehat \mathcal{H}_{M}$, and $t \geq t_{k}$, the following inequality holds, for some positive constant $C_{*}$.
\begin {equation} \label{oddeven} \left \| \left \langle \Psi_{\gamma}(\zeta)\; G(\eta_{t}) \vert \frak M_{0}\right \rangle \right \|_{M} \leq C_{*}\; \| G\|_{M} \; \bar \mu^{t-t_{k}} \; \mu_{*}^{|\gamma|}. \end{equation}
\end{proposition}
\begin{proof} \label {Proof.} Proceeding as in the previous proof, we see that if $G$ is odd and $|\gamma|>1$, we get, in analogy with \eqref{iterazione},
\begin {equation} \label{odd1} \left \| \left \langle \Psi_{\gamma}(\zeta)\; G(\eta_{t}) \vert \frak M_{0}\right \rangle \right \|_{M} \leq M^{|\gamma|} \| G\|_{M}\; \bar \mu^{t-t_{k}} \prod_{j \;even} \bar \mu^{r_{j}} \prod_{j\; odd} (1+ 2\bar \mu^{r_{j}}), \end{equation}
and if $G$ is even, by an obvious modification of the proof,
\begin {equation} \label{even1} \left \| \left \langle \Psi_{\gamma}(\zeta)\; G(\eta_{t}) \vert \frak M_{0}\right \rangle \right \|_{M} \leq M^{|\gamma|} \| G\|_{M}\; \bar \mu^{t-t_{k}} \prod_{j \;odd} \bar \mu^{r_{j}} \prod_{j\; even} (1+ 2\bar \mu^{r_{j}}). \end{equation}
Writing $G(\eta) = G^{(+)}(\eta) + G^{(-)}(\eta)$, where $G^{(\pm)}(\eta) = {G(\eta) \pm G(-\eta)\over 2} \in \widehat \mathcal{H}_{M}$, and observing that
$\|G\|_{M} = \|G^{(+)}\|_{M} + \|G^{(-)}\|_{M}$, we get the result \eqref{oddeven}.
\end{proof}. \par \smallskip
\noindent {\bf Proof of Inequality \ref{nnjeravv}.}
We denote by $m_{\gamma}, M_{\gamma}$ the minimum and maximum of the set $\gamma\in \frak g$, $\gamma\neq \emptyset$, and if $\gamma = \{t_{0}, \ldots, t_{k}\}$, then $\gamma +t =\{t_{0}+t, \ldots, t_{k}+t\}$, $t\geq - t_{0}$. \par\smallskip
Using the Walsh expansion \eqref{somma1} we have
$f(S^{t}\widehat \zeta) = \sum_{\gamma} f_{\gamma} \widehat \Psi_{\gamma+ t}(\widehat \zeta)$, and we write
\begin{equation}\label{split} f(\widehat \zeta) f(S^{t}\widehat \zeta) = f_{\emptyset} f(S^{t}\widehat \zeta) + C^{(1)}_{t}(\widehat \zeta) + C^{(2)}_{t}(\widehat \zeta) - R_{t}(\widehat \zeta)\end{equation}
$$ C^{(1)}_{t}(\widehat \zeta) = \sum_{\gamma, \gamma\prime\neq \emptyset\atop M_{\gamma} < m_{\gamma\prime}+t} f_{\gamma} f_{\gamma\prime} \Psi_{\gamma}(\widehat \zeta) \widehat \Psi_{\gamma\prime +t}(\widehat \zeta), \qquad C^{(2)}_{t}(\widehat \zeta) = \sum_{\gamma, \gamma\prime\neq \emptyset\atop M_{\gamma} \geq m_{\gamma\prime+t} }f_{\gamma} f_{\gamma\prime} \Psi_{\gamma}(\widehat \zeta) \Psi_{\gamma\prime +t} (\widehat \zeta) ,$$
and $R_{t}(\widehat \zeta) = \sum_{\gamma, \gamma\prime\neq \emptyset } f_{\gamma}\Psi_{\gamma}(\widehat \zeta) f_{\gamma\prime} \langle \Psi_{\gamma\prime}\rangle \chi(M_{\gamma} \geq m_{\gamma\prime+t})$, where $\chi$ is the indicator function. \par\smallskip
As $|\langle \Psi_{\gamma\prime}\rangle | \leq \|\langle \Psi_{\gamma\prime}\vert \frak M_{0}\rangle \|_{M}$ we see, by \eqref{lemmadue} and \eqref{correz}, that
\begin{equation}\label{erre} \left \| \left \langle R_{t}(\cdot) \vert \frak M_{0}\right \rangle \right \|_{M} \leq \sum_{\gamma\prime\neq \emptyset} \left |f_{\gamma\prime} \langle \Psi_{\gamma\prime}\rangle \right | \sum_{\gamma: M_{\gamma}\geq t} |f_{\gamma} \left \| \left \langle \Psi_{\gamma}\vert \frak M_{0}\right \rangle \right \|_{M} \; \leq \; {\rm const} \; \|\tilde f\|^{2}_{\mathcal{C}^{\alpha}} \; \kappa^{t} .\end{equation}
Passing to $C^{(1)}_{t}$,
let $\gamma, \gamma\prime\in \frak g$ be such that $M_{\gamma} = r$, $m_{\gamma\prime} = m$ and $r< m+t$. Taking the conditional expectation with respect to $\frak M_{0}^{r}$ we get by Proposition \ref{proposition},
$$ \left | \left \langle \Psi_{\gamma} \widehat \Psi_{\gamma\prime +t} \vert \frak M_{0}\right \rangle \right | = \left | \left \langle \Psi_{\gamma}\left [\mathcal{T}^{t+m-r} \widehat G_{\gamma\prime}\right ](\eta_{r}) \vert \frak M_{0}\right \rangle \right | \leq C_{*} \; \mu_{*}^{|\gamma|} \bar \mu^{t+m-r} \|\widehat G_{\gamma\prime} \|_{M}$$
where $\widehat G_{\gamma\prime}$ is defined in \eqref{decompo}. Therefore, again by Inequalities \eqref{lemmadue} and \eqref{correz}, we see that
$$ \left \| \left \langle C^{(1)}_{t}(\cdot) \vert \frak M_{0}\right \rangle \right \|_{M} \leq {\rm const}\; \|\tilde f\|_{\alpha}^{2}
\sum_{m=0}^{\infty} \sum_{r=0}^{t+m-1} \sum_{k=0}^{\infty}\;\bar \mu^{t+m-r} 2^{-\alpha (r+m+k)} A_{0}^{r} A_{m}^{m+k} $$
where, for $0\leq j\leq k\in \mathbb{Z}_{+}$ we set $ A_{j}^{k} : = \sum_{\gamma\in \frak g} \mu_{*}^{|\gamma|} \chi(m_{\gamma}=j, M_{\gamma} =k) \; < (1+\mu_{*})^{k-j+1} $.
As $\bar \mu < \kappa = 2^{-\alpha}(1+\mu_{*}) <1$, we get the estimate
\begin{equation}\label{pezzo1} \left \| \left \langle C^{(1)}_{t}(\cdot) \vert \frak M_{0}\right \rangle \right \|_{M} \leq {\rm const}\; \|f\|_{\alpha}^{2} \sum_{m=0}^{\infty} 2^{-\alpha m} \sum_{r=0}^{t+m-1} \kappa^{r} \bar \mu^{t+m-r} \leq {\rm const}\; \|f\|_{\alpha}^{2} \;\kappa^{t}. \end{equation}
\par\smallskip
Turning to $C^{(2)}_{t}$, observe that $\Psi_{\gamma} \Psi_{\gamma\prime} = \Psi_{\gamma\Delta \gamma\prime}$, where $\gamma\Delta\gamma\prime = \gamma\setminus \gamma\prime \; \cup \; \gamma\prime \setminus \gamma$, so that
\begin{equation} \label{quarta} \left \| \left \langle C^{(2)}_{t}(\cdot) \vert \frak M_{0}\right \rangle \right \|_{M} \leq \sum_{m, s, k=0}^{\infty} \; \sum_{\gamma: M_{\gamma} =t+m+s} |f_{\gamma}| \; \sum_{\gamma\prime: m_{\gamma\prime} = m \atop M_{{\gamma\prime}} = m+k} | f_{\gamma\prime}| \left | \left \langle \Psi_{\gamma \Delta \{\gamma\prime+t\}} \vert \frak M_{0} \right \rangle \right |_{M} . \end{equation}
Let $n= \min \{s, k\}$, $N=\max\{s,k\}$, $\Omega_{n} = \{m, \ldots, m+n\}$ and $\gamma_{1}=\gamma \cap \{0, \ldots, m-1\}$, $\gamma_{11}= \gamma\cap \Omega_{n}$, $\gamma_{12}= \gamma\prime\cap \Omega_{n}$, $\gamma_{2} = \gamma \cup \{\gamma\prime +t\} \cap \{t+n+1,\ldots, t+N\}$. If $\bar \gamma = \gamma_{11} \Delta \gamma_{12}$ we have
$\gamma \Delta \{\gamma\prime +t\} = \gamma_{1}\cup \bar \gamma \cup \gamma_{2}$, and the sets $\gamma_{1}, \bar \gamma, \gamma_{2}$ have no common elements.\par\smallskip
It is not hard to see by induction that
$$ \sum_{\gamma_{11}, \gamma_{12}\subseteq \Omega_{n}} \mu_{*}^{|\gamma_{11}\Delta \gamma_{12}|} = (2(1+\mu_{*}))^{n+1},$$
so that the sum on the right of \eqref{quarta}, for fixed $m,s,k$, is bounded by
$$ {\rm const}\; \|\tilde f\|_{\alpha}^{2} \; \kappa^{t+m} 2^{-\alpha m} \kappa^{N-n} (2^{-2\alpha+1 } (1+\mu_{*}))^{n} .$$
If $\alpha > 1/2$ and $\bar \mu$ is so small that $2^{-2\alpha+1 } (1+\mu_{*})< 1$, all series converge and we get
\begin{equation}\label{pezzo2} \left \| \left \langle C^{(2)}_{t}(\cdot) \vert \frak M_{0}\right \rangle \|_{M} \right \| \leq {\rm const}\; \|\tilde f\|_{\alpha}^{2} \;\kappa^{t}. \end{equation}
\par\smallskip Finally, the inequality $|f_{\emptyset}| \| \langle f(S^{t}\cdot
\vert \frak M_{0}\rangle \|_{M} \leq {\rm const}\; \bar \mu^{t} \|\tilde f\|_{\mathcal{C}^{\alpha}}^{2}$ is an immediate consequence of Theorem \ref{base} and Lemma \ref{lemma:hoelder}. The proof of \eqref{nnjeravv} follows from this estimate, together with the previous estimates \eqref{erre}, \eqref{pezzo1} and \eqref{pezzo2}.
\hfill \hbox{$\sqcap\kern-7pt\sqcup$}
\subsection*{Acknowledgements}
A.M. is supported by the European social fund within the framework of realizing the project ``Support of inter-sectoral mobility and quality enhancement of research teams at Czech Technical University in Prague'' (CZ.1.07/2.3.00/30.0034).\\
C.S. is supported by ERC Grant MAQD 240518.
\vskip 1.5 cm
|
\subsection*{Question #1}}
\newcommand{\qpart}[1]{\paragraph{(#1)}}
\newcommand{\bigparen}[1]{\left( #1 \right)}
\renewcommand\arraystretch{1.0}%
\begin{document}
\begin{titlepage}
\begin{center}
{\LARGE \bf Who's good this year? Comparing the Information Content of Games in the Four Major US Sports} \\
\ \\
{\large Julian Wolfson and Joseph S. Koopmeiners \\
\ \\
Division of Biostatistics, University of Minnesota, Minneapolis, Minnesota }\\
\ \\
{\large \it Corresponding author's email address: <EMAIL>} \\
\ \\
{\large \today}
\end{center}
\begin{abstract}
In the four major North American professional sports (baseball, basketball, football, and hockey), the primary purpose of the regular season is to determine which teams most deserve to advance to the playoffs. Interestingly, while the ultimate goal of identifying the best teams is the same, the number of regular season games played differs dramatically between the sports, ranging from 16 (football) to 82 (basketball and hockey) to 162 (baseball). Though length of season is partially determined by many factors including travel logistics, rest requirements, playoff structure and television contracts, it is hard to reconcile the 10-fold difference in the number of games between, for example, the NFL and MLB unless football games are somehow more ``informative'' than baseball games. In this paper, we aim to quantify the amount of information games yield about the relative strength of the teams involved. Our strategy is to assess how well simple paired comparison models fitted from $X$\% of the games within a season predict the outcomes of the remaining ($100-X$)\% of games, for multiple values of $X$. We compare the resulting predictive accuracy curves between seasons within the same sport and across all four sports, and find dramatic differences in the amount of information yielded by individual game results in the four major U.S. sports.
\end{abstract}
\end{titlepage}
\newpage
\section{Introduction}
In week 14 of the 2012 NFL season, the 9-3 New England Patriots squared off on Monday Night Football against the 11-1 Houston Texans in a game with major implications for both teams. At the time, the Texans had the best record in the AFC and were in line to earn home-field advantage throughout the playoffs, while the New England Patriots had the best record in their division and were hoping to solidify their playoff position and establish themselves as the favorites in the AFC. The Patriots ultimately defeated the Texans 42-14, which led some commentators to conclude that the Patriots were the favorites to win the Super Bowl \citep{Walker2012,MacMullan2012} and that Tom Brady was the favorite for the MVP award \citep{Reiss2012}, while others opined that the Texans were closer to pretenders than the contenders they appeared to be for the first 13 weeks of the season \citep{Kuharsky2012}. These are strong conclusions to reach based on the results of a single game, but the power of such ``statement games'' is accepted wisdom in the NFL. In contrast, it is rare for the outcome of a single regular-season game to create or change the narrative about a team in the NBA, NHL, or MLB. While one might argue that the shorter NFL season simply drives commentators to imbue each game with greater metaphysical meaning, an alternative explanation is that the outcome of a single NFL contest actually does carry more information about the relative strengths of the teams involved than a single game result in the other major North American professional sports. In this paper, we ask and attempt to answer the basic question: how much does the outcome of a single game tell us about the relative strength of the two teams involved?
In the four major North American professional sports (baseball, basketball, football, and hockey), the primary purpose of the regular season is to determine which teams most deserve to advance to the playoffs. Interestingly, while the ultimate goal of identifying the best teams is the same, the number of regular season games played differs dramatically between the sports, ranging from 16 (football) to 82 (basketball and hockey) to 162 (baseball). Though length of season is partially determined by many factors including travel logistics, rest requirements, playoff structure and television contracts, it is hard to reconcile the 10-fold difference in the number of games in the NFL and MLB seasons unless games in the former are somehow more informative about team abilities than games in the latter. Indeed, while it would be near-heresy to determine playoff eligibility based on 16 games of an MLB season (even if each of the 16 games was against a different opponent), this number of games is considered adequate for the same purpose in the NFL.
There is a well-developed literature on the topic of competitive balance and parity in sports leagues \citep{Owen2010,Horowitz1997,Mizak2005,Lee2010,Hamlen2007,Cain2006,Larsen2006,Ben-Naim2006a,Kesenne2000,Vrooman1995,Koopmeiners2012}. However, most papers focus on quantifying the degree of team parity over consecutive years along with the effects of measures taken to increase or decrease it. In papers which compare multiple sports, season length is often viewed as a nuisance parameter to be adjusted for rather than a focus of inquiry. Little attention has been directed at the question of how information on relative team strength accrues over the course of a single season.
In this paper, we aim to quantify the amount of information each games yields about the relative strength of the teams involved. We estimate team strength via paired-comparison \citep{Bradley1952} and margin-of-victory models which have been applied to ranking teams in a variety of sports \citep{McHale2011, Koehler1982, Sire2009, Martin1999}. The growth in information about the relative strength of teams is quantified by considering how these comparison models fitted from $X$\% of the games in a season predict the outcomes of the remaining ($100-X$)\% of games, for multiple values of $X$ (games are partitioned into training and test sets at random to reduce the impact of longitudinal trends over the course of a season). We begin by describing the data and analysis methods we used in Section 2. Section 3 presents results from recent seasons of the four major North American sports, and compares the ``information content'' of games across the four sports. In Section 4 we discuss the strengths and limitations of our analysis.
\section{Methods}
\subsection{Data}
We consider game results (home and away score) for the 2004-2012 seasons for the NFL, the 2003-2004 to 2012-2013 seasons of the NBA, the 2005-2006 to 2012-2013 seasons of the NHL, and the 2006-2012 seasons of MLB. Game results for the NFL, NBA and NHL were downloaded from Sports-Reference.com \citep{pfr} and game results for MLB were downloaded from Retrosheet \citep{retrosheet}. Only regular season games were considered in our analysis. The NFL plays a 16-game regular season, the NBA and NHL play 82 regular season games and MLB plays 162 regular season games.
\subsection{Methods}
Let $\mathcal{G}$ represent all the games within a single season of a particular sport. Our goal is to quantify the amount of information on relative team strength contained in the outcomes of a set of games $G \subset \mathcal{G}$, as the number of games contained in $G$ varies. We consider how well the results of games in the ``training set'' $G$ allow us to predict the outcomes of games in a ``test set'' $G' = \mathcal{G} \setminus G$. Specifically, given $G$ and $G'$, our information metric (which we formally define later) is the percentage of games in $G'$ which are correctly predicted using a paired comparison model applied to $G$.
We consider two types of paired comparison models in our work. Each game $g \in G$ provides information on the home team ($H_g = i$), away team ($A_g = j$) and the game result as viewed from the home team's perspective. When only the binary win/loss game result $W_g$ is considered, we fit a standard Bradley-Terry model \citep{Bradley1952, Agresti02},
\begin{equation}
logit \left(\pi_{i,j}\right) = \beta_{i} - \beta_{j} + \alpha,
\label{eq:BT}
\end{equation}
where $\pi_{i,j} = P(W_g = 1)$ is the probability that the home team, team $i$, defeats the visiting team, team $j$. $\beta_{i}$ and $\beta_{j}$ are the team strength parameters for teams $i$ and $j$, respectively, and $\alpha$ is a home-field advantage parameter.
We fit a similar model when the actual game scores are considered. In this context, home team margin of victory (MOV) $\Delta_g$ is recorded for each game; $\Delta_g$ is positive for a home team win, negative for a home team loss, and zero for a tie. The paired comparison model incorporating margin of victory is:
\begin{equation}
\mu_{i,j} = \delta_{i} - \delta_{j} + \lambda,
\label{eq:MOV}
\end{equation}
where $\mu_{i,j} = E(\Delta_g)$ is the expected margin of victory for the home team, team $i$, over the visiting team, team $j$. $\delta_{i}$ and $\delta_{j}$ are team strengths on the margin-of-victory scale for teams $i$ and $j$, respectively, and $\lambda$ is a home-field advantage on the margin-of-victory scale.
Both models \eqref{eq:BT} and \eqref{eq:MOV} can be fit using standard statistical software, such as R \citep{rsoftware}. Given estimates $\hat \beta_i$, $\hat \beta_j$, and $\hat \alpha$ derived by fitting model \eqref{eq:BT} to a set of games $G$, a predicted home team win probability $\hat \pi_g$ can be derived for every game $g \in G'$ based on which teams $i$ and $j$ are involved. A binary win/loss prediction for the home team is obtained according to whether $\hat \pi_g$ is greater/less than 0.5. Given estimates $\hat \delta_i$, $\hat \delta_j$, and $\hat \lambda$ from fitting model \eqref{eq:MOV}, home team margin of victory $\hat \mu_g$ can similarly be predicted for every game in $g \in G'$. A binary win/loss prediction for the home team is obtained according to whether $\hat \mu_g$ is positive, negative, or zero.
Our metrics for summarizing the amount of information on relative team strength available from a set of game results $G$ for predicting the outcomes of a set of games in $G'$ are simply the fraction of games that are correctly predicted by the paired comparison models:
\begin{align*}
\mathcal{I}^{BT}(G,G') &= \frac{ \sum_{g \in G'} W_g \indicatorBig{\hat \pi_g > 0.5}}{ | G' | } \ \ \text{for the Bradley-Terry model \eqref{eq:BT}}\\
\mathcal{I}^{MOV}(G,G') &= \frac{ \sum_{g \in G'} W_g \indicatorBig{\hat \mu_g > 0}}{ | G' | } \ \ \text{for the margin-of-victory model \eqref{eq:MOV}}
\end{align*}
where $\hat \pi_g$ and $\hat \mu_g$ are estimates derived from game results in $G$, and $|G'|$ denotes the number of games in $G'$.
For a given season, training data sets $G_1, G_2, \dots, G_K$ were formed by randomly sampling games corresponding to X\% of that season. Test data sets $G'_1, G'_2, \dots, G'_K$ were created as the within-season complements of the training sets, i.e., if $G_k$ consists of a number of games corresponding to X\% of the season, then $G'_k$ contains the remaining (100-X)\% of games in that season. Training (and corresponding test) data sets were created for X\% = 12.5\%, 25.0\%, 37.5\%, 50.0\%, 62.5\%, 75.0\% and 87.5\% of the games in each available season. Games were sampled at random so as to reduce the influence of temporal trends over the course of a season, for example, baseball teams who are out of playoff contention trading away valuable players and giving playing time to minor league prospects in August and September.
Information on relative team strength over a single season was computed and summarized as follows:
\begin{enumerate}
\item For X = 12.5, 25, 37.5, 50, 62.5, 75, and 87.5:
\begin{enumerate}
\item Generate 100 training sets $G_1, G_2, \dots, G_{100}$ (and complementary test sets $G'_1, G'_2, \dots, G'_{100}$) by randomly sampling X\% of games without replacement from $\mathcal{G}$.
\item For each training set $G_k$:
\begin{enumerate}
\item Fit models \eqref{eq:BT} and \eqref{eq:MOV} to the games in $G_k$.
\item Obtain binary win/loss predictions for the games in the test set $G'_k$.
\item Evaluate the information metrics $\mathcal{I}^{BT}(G_k, G'_k)$ and $\mathcal{I}^{MOV}(G_k,G'_k)$
\end{enumerate}
\item Average the computed information metrics to estimate the predictive accuracy of paired comparison models fitted to data from X\% of the entire season ($\mathcal{I}^{BT}$ and $\mathcal{I}^{MOV}$).
\end{enumerate}
\item Tabulate and plot $\mathcal{I}^{BT}$ and $\mathcal{I}^{MOV}$ across different values of X.
\end{enumerate}
The natural comparison value for our information metrics is the predictive accuracy of a naive model which chooses the home team to win every game. As shown in the plots below the average win probability for the home team (as determined by the parameters $\alpha$ and $\lambda$ in models \eqref{eq:BT} and \eqref{eq:MOV} respectively) varies from approximately 53\% to 61\% across the four sports we consider.
\section{Results}
\subsection{National Football League}
\begin{figure}[!h]
\centering
\includegraphics*[width = 6in]{NFL04-12.pdf}
\caption{Percent of games correctly predicted on test set vs. average number of games per team in training set, NFL seasons 2004-2012.\label{nfl_results}}
\end{figure}
Figure~\ref{nfl_results} plots the percent of games correctly predicted on the test set versus the average number of games per team in the training set for the 2004-2012 National Football League seasons. Both paired comparison models (i.e., those which incorporate and ignore margin of victory) outperform simply picking the home team to win every game. The margin of victory model appears to perform slightly better than the paired comparison model, though the differences are modest and in some seasons (e.g., 2004 and 2008) are non-existent. The prediction accuracy of both models improves throughout the season in most seasons (years 2008 and 2009 being notable exceptions), indicating that we are gaining information about the relative strengths of teams even in the final weeks of the season.
\subsection{National Basketball Association}
\begin{figure}[!h]
\centering
\includegraphics*[width = 6in]{NBA03-13.pdf}
\caption{Percent of games correctly predicted on test set vs. average number of games per team in training set, NBA seasons 2003-2013. \label{nba_results}}
\end{figure}
Results for the National Basketball Association can be found in Figure~\ref{nba_results}. The NBA was the most predictable of the four major North American professional sports leagues. Using 87.5\% of games as a training set, our model was able to accurately predict up to 70\% across seasons. The NBA also had the largest home court advantage with home teams winning approximately 60\% of games. There was virtually no advantage in including margin of victory in our model; indeed, it led to slightly worse predictions during the 05-06 season. The only major difference between the NFL and NBA was the growth of information over the season. While the accuracy of our predictions for the NFL continued to improve as more games were added to the training set, model accuracy for the NBA was no better when 75\% of games were included in the training set than when 25\% of games were included. Analyses using the \texttt{segmented} package in R for fitting piecewise linear models \citep{Muggeo2003, Muggeo2008} confirmed an inflection point in the prediction accuracy curve approximately 25-30 games into the season.
\subsection{Major League Baseball and the National Hockey League}
\begin{figure}[!h]
\centering
\includegraphics[width = 6in]{MLB06-12.pdf}
\caption{Percent of games correctly predicted on test set vs. average number of games per team in training set, MLB seasons 2006-2012. \label{mlb_results}}
\end{figure}
\begin{figure}[!h]
\centering
\includegraphics[width = 6in]{NHL05-13.pdf}
\caption{Percent of games correctly predicted on test set vs. average number of games per team in training set, NHL seasons 2005-2013. \label{nhl_results}}
\end{figure}
Results from Major League Baseball and the National Hockey League are found in Figures~\ref{mlb_results} and ~\ref{nhl_results}, respectively. The results for MLB and the NHL were quite similar, in that both leagues were substantially less predictable than the NFL and NBA. The percentage of games correctly predicted for MLB never exceeded 58\% even when 140 games (7/8 of a season) were included in the training set. The NHL was slightly better but our model was never able to predict more than 60\% of games correctly (and this was only achieved in the 2005-2006 season when the home team win probability was relatively high at 58\%). More importantly, prediction accuracy was rarely more than 2-3 percentage points better than the simple strategy of picking the home team in every game for either league. In fact, during the 2007-2008 and 2011-2012 seasons picking the home team performed better than paired comparison models constructed using a half-season's worth of game results.
It is perhaps not surprising that the outcome of a randomly chosen baseball game is hard to predict based on previous game results given the significant role that the starting pitcher plays in determining the likelihood of winning. In a sense, the ``effective season length'' of MLB is far less than 162 games because each team-pitcher pair carries a different win probability. In additional analyses (results not shown), we fit paired comparison models including a starting pitcher effect, but this did not substantially affect our results.
\subsection{Comparing the sports}
Figure \ref{allsports} displays curves of summarizing predictive accuracy of the MOV model for the four major sports, aggregated across the years of available data (results from the win-loss model were similar). We see that, even after only 1/8th of the games in a season have been played, substantial information on relative team strength has already accrued in the NBA, while much less can be said at this point about the NFL, NHL, and MLB. Predictive accuracy increases most rapidly with additional games in the NFL, so that predictive accuracy approaches that of the NBA when a substantial fraction of games are used for prediction. As seen above, the overall predictive accuracies for the MLB and NHL remain low, and do not increase markedly with the fraction of games in the training set.
\begin{figure}[!h]
\centering
\includegraphics[width=\textwidth]{allsports_MOV.pdf}
\caption{Percent of games correctly predicted by margin of victory model on test set vs. percent of season in training set, for four major U.S. sports leagues. For each sport, connected plotting symbols represent the average predictive accuracy and shaded regions enclose the range of predictive accuracies across the seasons of available data. \label{allsports}}
\end{figure}
Table \ref{EOS.OR} gives one way of summarizing the informativeness of games in the four major sports, via an odds ratio comparing the predictive accuracy of two models: 1) the MOV paired comparison model using game data from 87.5\% of the season, and 2) a prediction ``model'' which always picks the home team. There is a clear separation between the NFL/NBA, where games played during the season improve the odds of making correct predictions by about 40\% over a ``home-field advantage only'' model, and the NHL/MLB, where the improvement is under 10\%.
\begin{table}[!h]
\centering
\caption{Odds ratio comparing the predictive accuracy of a MOV paired comparison model using data from 87.5\% of a season to the accuracy of a model which always picks the home team. \label{EOS.OR}}
\begin{tabular}{l|c}
& OR\\
\hline
\textbf{NBA} & 1.41\\
\textbf{NFL} & 1.46\\
\textbf{NHL} & 1.09\\
\textbf{MLB} & 1.06
\end{tabular}
\end{table}
Table \ref{infopergame} summarizes the per-game rate of increase in predictive model accuracy for the four sports. The estimates are obtained by fitting least-squares regression lines to the data displayed in Figure \ref{allsports}. The lines for each sport are constrained to a an intercept of 0.5, representing the predictive accuracy of a ``no-information'' model before any games have been played. In developing prediction models for actual use, one might want to incorporate prior information on the home-field advantage based on previous seasons, but in our simple paired comparison models both team strengths and the home-field advantage are estimated purely from current-season data. Hence, prior to any games being played these models can perform no better than flipping a fair coin. The columns of Table \ref{infopergame} correspond to the estimated rate of increase in predictive accuracy, on a percentage point per game scale, over 25\%, 37.5\%, 50\% and 87.5\% of the season.
\begin{table}[!h]
\centering
\caption{Estimated per-game percentage point increase in predictive accuracy of a margin-of-victory model for the four U.S. sports leagues, by percentage games used to train the model. \label{infopergame}}
\begin{tabular}{l|cccc}
& 25\% of games & 37.5\% of games & 50\% of games & 87.5\% of games\\
\hline
\textbf{NBA} & 0.91 & 0.69 & 0.55 & 0.34\\
\textbf{NFL} & 2.6 & 2.3 & 2 & 1.4\\
\textbf{NHL} & 0.29 & 0.23 & 0.19 & 0.13\\
\textbf{MLB} & 0.12 & 0.094 & 0.079 & 0.053\\
\end{tabular}
\end{table}
The results in Table \ref{infopergame} allow us to compute a ``per-game informativeness ratio'' between pairs of sports. For example, considering the last column allows us to estimate that, over the course the season, NFL games are approximately 4 times more informative than NBA games, which are in turn about 2-3 times more informative than NHL games, which are themselves approximately 2-3 times more informative than MLB games. The ``informativeness ratio'' of NFL to MLB games is on the order of 65, or about 6 times larger than the inverse ratio of their respective season lengths (162/16 $\approx$ 10). In contrast, the ratio comparing NFL to NBA games ($\approx$ 4) is slightly smaller than the inverse ratio of their respective season lengths (82/16 $\approx$ 5).
\section{Conclusions and discussion}
Our results reveal substantial differences between the major North American sports according to how well one is able to discern team strengths using game results from a single season. NBA games are most easily predicted, with paired comparison models having good predictive accuracy even early in the season; indeed, since our information metric for the NBA appears to plateau around game 30, an argument could be made that the latter half of the NBA season could be eliminated without substantially affecting the ability to identify the teams most deserving of a playoff spot. NFL game results also give useful information for determining relative team strength. On a per-game basis, NFL contests contain the largest amount of information. With the exception of the 2008 season, there was no obvious ``information plateau'' in the NFL, though the rate of increase in information did appear to slow somewhat after the first 5 games. These results suggest that games in the latter part of the NFL season contribute useful information in determining who the best teams are.
The predictive ability of paired comparison models constructed from MLB and NHL game data remains limited even when results from a large number of games are used. One interpretation of this finding is that, in comparison to the NBA and NFL, games in MLB and the NHL carry little information about relative team strength. Our results may also reflect smaller variance in team strengths (i.e., greater parity) in hockey and baseball: Because our information metric considers the predictive accuracy averaged across all games in the test set, if most games are played between opposing teams of roughly the same strength then most predictive models will fare poorly. Indeed, the inter-quartile range for winning percentage in these sports is typically on the order of $\sim$20\%, while in football and basketball it is closer to 30\%. Our observation that the hockey and baseball regular seasons do relatively little to distinguish between teams' abilities is reflected in playoff results, where ``upsets'' of top-seeded teams by teams who barely qualified for the postseason happen much more regularly in the NHL and MLB than in the NFL and NBA. One possible extension of this work would be to quantify this effect more formally.
Indeed, given the relative inability of predictive models to distinguish between MLB teams upon completion of the regular season, a compelling argument could be made for increasing the number of teams that qualify for the MLB playoffs since the current 10-team format is likely to exclude teams of equal or greater ability than ones that make it. Using similar logic, one might also argue that if the goal of the playoffs is to identify the best team (admittedly an oversimplification), then perhaps the NBA playoffs are \emph{overly} inclusive as there is ample information contained in regular season game outcomes to distinguish between the best teams and those that are merely average.
More surprising to us was the enormous discrepancy in the informativeness of game results between hockey and basketball, which both currently play seasons of the same length but perhaps ought not to. One possible explanation for why basketball game results more reliably reflect team strength is that a large number of baskets are scored, and the Law of Large Numbers dictates that each team approaches their ``true'' ability level more closely. In contrast, NHL games are typically low-scoring affairs, further compounded by the fact that a large fraction of goals are scored on broken plays and deflections which seem to be strongly influenced by chance. We have not analyzed data from soccer, but it would be interesting to explore whether the ``uninformativeness'' of hockey and baseball game results extends to other low-scoring sports.
Our analysis has several limitations. First, we chose to quantify information via the predictive accuracy of simple paired comparison models. It is possible that using more sophisticated models for prediction might change our conclusions, though we doubt it would erase the sizable between-sport differences that we observed. Indeed, as mentioned above, accounting for starting pitcher effects in our MLB prediction model did not substantially affect the results. Second, it could be argued that team win probabilities change over the course of a season due to roster turnover, injuries, and other effects. By randomly assigning games to our training and test set without regard to their temporal ordering, we are implicitly estimating ``average'' team strengths over the season, and applying these to predict the outcome of an ``average'' game. We chose a random sampling approach over one which would simply split the season because we wanted to eliminate time trends in team strengths when describing how information accrued as more game results were observed. While our approach does not directly describe how predictive accuracy improves as games are played in their scheduled order, we anticipate that the patterns would be similar to what we observed.
\bibliographystyle{plainnat}
\begingroup
\sloppy
|
\section{Introduction}
\subsection{Logic, automata and coalgebra}
The aim of this paper is to strengthen the link between the areas of logic,
automata and coalgebra.
More in particular, we provide a coalgebraic generalization of the
automata-theoretic approach towards monadic second-order logic ($\mathlang{MSO}$), and
we address the question whether the Janin-Walukiewicz Theorem can be
generalized from Kripke structures
to the setting of arbitrary coalgebras.
The connection between \textit{monadic second-order logic} and \textit{automata}
is classic, going back to the seminal work of B\"uchi, Rabin, and others.
For instance, Rabin's decidability result for the monadic second-order theory
of binary trees, or $S2 S$, makes use of a translation of
monadic second-order logic into a class of automata, thus reducing the
satisfiability problem for $S 2 S$ to the non-emptiness problem for the
corresponding automata \cite{rabi:deci69}.
The link between $\mathlang{MSO}$ and automata over trees with arbitrary branching
was further explored by Walukiewicz~\cite{walu:mona96}.
Janin and Walukiewicz considered monadic second-order logic interpreted over
Kripke structures, and used automata-theoretic techniques to obtain a van
Benthem-like characterization theorem for monadic second-order logic, identifying
the modal $\mu$-calculus as the bisimulation invariant fragment of
$\mathlang{MSO}$ \cite{jani:expr96}.
Given the fact that in many applications bisimilar models are considered
to represent the \emph{same} process, one has little interest in properties of
models that are \emph{not} bisimulation invariant.
Thus the Janin-Walukiewicz Theorem can be seen as an expressive completeness
result, stating that all \emph{relevant} properties in monadic second-order
logic can be expressed in the modal $\mu$-calculus.
Coalgebra enters naturally into this picture.
Recall that Universal Coalgebra~\cite{rutt:univ00} provides the notion of a
\emph{coalgebra} as the natural mathematical generalization of state-based
evolving systems such as streams, (infinite) trees, Kripke models,
(probabilistic) transition systems, and many others.
This approach combines simplicity with generality and wide applicability: many
features, including input, output, nondeterminism, probability, and interaction,
can easily be encoded in the coalgebra type $\mathsf{T}$ (formally an endofunctor on
the category $\mathbf{Set}$ of sets as objects with functions as arrows).
Starting with Moss' seminal paper~\cite{moss:coal99}, coalgebraic logics have
been developed for the purpose of specifying and reasoning about \emph{behavior},
one of the most fundamental concepts that allows for a natural coalgebraic
formalization.
And with Kripke structures constituting key examples of coalgebras, it should
come as no surprise that most coalgebraic logics are some kind of modification
or generalization of \emph{modal logic}~\cite{cirs:moda11}.
The coalgebraic modal logics that we consider here originate with
Pattinson~\cite{patt:coal03}; they are characterized by a completely standard
syntax, in which the semantics of each modality is determined by a so-called
\emph{predicate lifting} (see Definition~\ref{d:pl} below).
Many well-known variations of modal logic in fact arise as the coalgebraic logic
$\mathlang{ML}_{\Lambda}$ associated with a set $\Lambda$ of such predicate liftings; examples
include both standard and (monotone) neighborhood modal logic, graded and
probabilistic modal logic, coalition logic, and conditional logic.
Extensions of coalgebraic modal logics with fixpoint operators, needed
for describing \emph{ongoing} behavior, were developed
in~\cite{vene:auto06,cirs:expt09}.
The link between coalgebra and automata theory is by now well-established.
For instance, finite state automata operating on finite words have been
recognized as key examples of coalgebra from the
outset~\cite{rutt:univ00}.
More relevant for the purpose of this paper is the link with precisely the
kind of automata mentioned earlier, since the (potentially infinite) objects
on which these devices operate, such as streams, trees and Kripke frames,
usually are coalgebras.
Thus, the automata-theoretic perspective on modal fixpoint logic could be
lifted to the abstraction level of coalgebra~\cite{vene:auto06,font:auto10}.
In fact, many key results in the theory of automata operating on infinite
objects, such as Muller \& Schupp's Simulation Theorem~\cite{mull:simu95} can
in fact be seen as instances of more general theorems in Universal
Coalgebra~\cite{kupk:coal08}.
\subsection{Coalgebraic monadic second-order logic}
Missing from this picture is, to start with, a coalgebraic version of
\emph{(monadic) second-order logic}.
Filling this gap is the first aim of the current paper, which introduces a notion
of \emph{monadic second-order logic} $\MSO_{\fun}$ for coalgebras of type $\mathsf{T}$.
Our formalism combines two ideas from the literature.
First of all, we looked for inspiration to the coalgebraic versions of
\emph{first-order logic} of Litak \& alii~\cite{lita:coal12}.
These authors introduced Coalgebraic Predicate Logic as a common generalisation
of first-order logic and coalgebraic modal logic, combining first-order
quantification with coalgebraic syntax based on predicate liftings.
Our formalism $\MSO_{\fun}$ will combine a similar syntactic feature with second-order
quantification.
Second, following the tradition in automata-theoretic approaches towards monadic
second-order logic, our formalism will be \emph{one-sorted}.
That is, we \emph{only} allow second-order quantification in our language,
relying on the fact that individual quantification, when called for, can be
encoded as second-order quantification relativized to singleton sets.
Since predicate liftings are defined as families of maps on powerset
algebras, these two ideas fit together very well, to the effect that our
second-order logic is in some sense simpler than the first-order
formalism of~\cite{lita:coal12}.
In section~\ref{sec:mso} we will define, for any set $\Lambda$ of
monotone\footnote{%
In the most general case, restricting to monotone predicate liftings is not
needed, one could define $\MSO_{\fun}$ as the logic obtained by taking for $\Lambda$
the set of \emph{all} predicate liftings.
However, in the context of this paper, where we take an automata-theoretic
perspective on $\mathlang{MSO}$, this restriction makes sense.
}
predicate liftings, a formalism $\MSO_{\Lambda}$, and we let $\MSO_{\fun}$ denote the logic
obtained by taking for $\Lambda$ the set of \emph{all} monotone predicate
liftings.
Clearly we will make sure that this definition generalizes the standard case,
in the sense that the standard version of $\mathlang{MSO}$ for Kripke structures
instantiates the logic $\mathlang{MSO}_{\{\Diamond\}}$ and is equivalent to the coalgebraic
logic $\mathlang{MSO}_{\mathcal{P}}$ (where $\mathcal{P}$ denotes the power set functor).
The introduction of a monadic second-order logic $\MSO_{\fun}$ for $\mathsf{T}$-coalgebras
naturally raises the question, for which $\mathsf{T}$ the coalgebraic modal
$\mu$-calculus for $\mathsf{T}$ corresponds to the bisimulation-invariant fragment of
$\MSO_{\fun}$.
\begin{question}
\label{q:Q}
\text{Which functors $\mathsf{T}$ satisfy $\muML_{\fun} \equiv \MSO_{\fun}/{\simeq}$?}
\end{question}
\subsection{Automata for coalgebraic monadic second-order logic}
In order to address Question~\ref{q:Q}, we take an automata-theoretic
perspective on the logics $\MSO_{\fun}$ and $\muML_{\fun}$, and as the second contribution
of this paper we introduce a class of parity automata for $\MSO_{\fun}$.
As usual, the operational semantics of our automata is given in terms
of a two-player acceptance game, which proceeds in \emph{rounds} moving from
one basic position to another, where a basic position is a pair consisting of
a state of the automaton and a point in the coalgebra structure under
consideration.
In each round, the two players, $\exists$ and $\forall$, focus on a certain
local `window' on the coalgebra structure.
This `window' takes the shape of a \emph{one-step $\mathsf{T}$-model}, that is, a
triple $(X,\alpha,V)$ consisting of a set $X$, a \emph{chosen object} $\alpha
\in \mathsf{T} X$, and
a valuation $V$ interpreting the states of the automaton as subsets of $X$.
More specifically, during each round of the game it is the task of $\exists$
to come up with a valuation $V$ that creates a one-step model in which a
certain \emph{one-step formula} $\delta$ (determined by the current basic
position in the game) is true.
Generally, our automata will have the shape $\mathbb{A} = (A,\Delta,\Omega,a_{I})$
where $A$ is a finite carrier set with initial state $a_{I} \in A$, and $\Omega$
and $\Delta$ are the parity and transition map of $\mathbb{A}$, respectively.
The flavour of such an automaton is largely determined by the co-domain of its
transition map $\Delta$, the so-called \emph{one-step language} which consists of
the one-step formulas that feature in the acceptance game as
described.
Each one-step language $\mathlang{L}$ induces its own class of automata
$\mathit{Aut}(\mathlang{L})$.
For instance, the class of automata corresponding to the coalgebraic fixpoint
logic $\muML_{\Lambda}$ can be given as $\mathit{Aut}(\mathlang{ML}_{\Lambda})$, where $\mathlang{ML}_{\Lambda}$ is the set
of positive modal formulas of depth one that use modalities from
$\Lambda$~\cite{font:auto10}.
Basically then, the problem of finding the right class of automata for the
coalgebraic monadic second-order logic $\MSO_{\Lambda}$ consists in the identification
of an appropriate one-step language.
Our proposal comprises a one-step \emph{second-order logic} which uses predicate
liftings to describe the chosen object of the one-step model.
Finally, note that similar to the case of standard $\mathlang{MSO}$, the equivalence
between formulas in $\MSO_{\fun}$ and automata in $\mathit{Aut}(\mathlang{SO})$ is only guaranteed to
hold for coalgebras that are `tree-like' in some sense (to be defined further
on).
\begin{theo}[Automata for coalgebraic $\mathlang{MSO}$]
\label{t:automatachar}
For any set $\Lambda$ of monotone predicate liftings for $\mathsf{T}$ there is an
effective construction mapping any formula $\varphi\in \MSO_{\Lambda}$ into an
automaton $\mathbb{A}_\varphi \in \mathit{Aut}(\mathlang{SO}_{\Lambda})$,
which is
equivalent to $\varphi$ over $\mathsf{T}$-tree models.
\end{theo}
The proof of Theorem~\ref{t:automatachar} proceeds by induction on the
complexity of $\MSO_{\fun}$-formulas, and thus involves various \emph{closure
properties} of automata, such as closure under complementation, union and
projection.
In order to establish these results, it will be convenient to take an
\emph{abstract} perspective, revealing how closure properties of a class of
automata are completely determined at the level of the one-step language.
\subsection{Bisimulation Invariance}
With automata-theoretic characterizations in place for both coalgebraic $\mathlang{MSO}$
and the coalgebraic $\mu$-calculus $\mathlang{\mu ML}$, we can address Question~\ref{q:Q}
by considering the following question:
\begin{question}
\label{q:Qa}
\text{Which functors $\mathsf{T}$ satisfy $\mathit{Aut}(\mathlang{ML})\equiv \mathit{Aut}(\mathlang{SO})/{\simeq}$?}
\end{question}
Continuing the program of the third author~\cite{vene:expr14}, we will approach
this question \emph{at the level of the one-step languages}, $\mathlang{SO}$ and $\mathlang{ML}$.
To start with, observe that any translation (from one-step formulas in) $\mathlang{SO}$ to
(one-step formulas in) $\mathlang{ML}$ naturally induces a translation from $\mathlang{SO}$-automata
to $\mathlang{ML}$-automata.
A new observation we make here is that any so-called \emph{uniform construction}
on the class of one-step models for the functor $\mathsf{T}$ that satisfies certain
\emph{adequacy} conditions, provides
(1) a translation $(\cdot)^{*}: \mathlang{SO} \to \mathlang{ML}$, together with
(2) a construction $(\cdot)_{*}$ transforming a pointed $\mathsf{T}$-model $(\mathbb{S},s)$
into a tree model $(\mathbb{S}_{*},s_{*})$ which is a coalgebraic pre-image of
$(\mathbb{S},s)$ satisfying
\[
\mathbb{A} \text{ accepts } (\mathbb{S}_{*},s_{*}) \text{ iff }
\mathbb{A}^{*} \text{ accepts } (\mathbb{S},s).
\]
From this it easily follows that an $\mathlang{SO}$-automaton $\mathbb{A}$ is bisimulation
invariant iff it is equivalent to the $\mathlang{ML}$-automaton $\mathbb{A}^{*}$.
On the basis of these observations we can prove the following generalisation of
the Janin-Walukiewicz Theorem.
\begin{theo}[Coalgebraic Bisimulation Invariance]
\label{t:main2}
If the set functor $\mathsf{T}$ admits an adequate uniform construction, then
\[
\muML_{\fun} \equiv \MSO_{\fun}/{\simeq}.
\]
\end{theo}
In our eyes, the significance of Theorem~\ref{t:main2} is twofold.
First of all, the proof separates the `clean', abstract part of
bisimulation-invariance results from the more functor-specific parts.
As a consequence, Theorem~\ref{t:main2} can be used to obtain immediate results
in particular cases.
Examples include the power set functor (standard Kripke structures),
where the adequate uniform construction roughly consists of taking $\omega$-fold
products (see Example~\ref{ex:psf}), the bag functor (Example~\ref{ex:bag}),
and all exponential polynomial functors (Corollary~\ref{c:epf}).
Second, in case the functor does \emph{not} admit an adequate uniform
construction, Theorem~\ref{t:main2} may still be of use in proving alternative
characterization results for the functor.
Instantiating the latter phenomenon is the \emph{monotone neighborhood functor}
$\mathcal{M}$ (see the next section for its definition).
The importance of this functor lies, among other things, in it providing a
coalgebraic semantics for monotone modal logic~\cite{hans:coal04}.
The coalgebraic monadic second-order language $\mathlang{MSO}_\mathcal{M}$ is equivalent to a
natural second-order language for reasoning about monotone neighborhood
structures that we shall denote by $\mathtt{MMSO}$, and $\mathlang{\mu ML}_\mathcal{M}$ is
equivalent to the fixpoint-extension of the monotone $\mu$-calculus, denoted
$\mu \mathtt{MML}$.
As we shall see in Proposition~\ref{p:no-adc-mon} below, $\mathcal{M}$ does \emph{not}
admit an adequate uniform construction.\footnote{%
This does not necessarily mean that the monotone $\mu$-calculus $\mathlang{\mu ML}_{\mathcal{M}}$
does \emph{not} correspond to the bisimulation-invariant fragment of
$\mathlang{MSO}_{\mathcal{M}}$, but it does mean that a proof of such a result, if provable at
all, will be significantly more involved than for those functors where
Theorem~\ref{t:main2} does apply.}
This, however, is not the end of the story.
It turns out that we \emph{can} find an adequate uniform construction for
a \emph{variant} $\mathcal{M}^{\star}$ of the functor $\mathcal{M}$ (see
Proposition~\ref{p:ad-monstar}).
As a corollary, we obtain a characterization of the fragment of
$\mathtt{MMSO}$ that is invariant under \emph{global} bisimulations
(bisimulations that are full on both domain and codomain).
This fragment turns out to be exactly the extension of the monotone $\mu$-calculus
with the global modalities (for precise definitions we refer to
section~\ref{sec:mon}), which we shall denote $\mu \mathtt{MML}_g$.
In this notation, our final contribution is the following characterization result:
\begin{theo}
\label{t:JWmonglob}
A formula in $\mathtt{MMSO}$ is invariant for global neighborhood
bisimulations if, and only if, it is equivalent to a formula of the logic
$\mu \mathtt{MML}_g$.
\end{theo}
\section{Some technical background}
In this paper we assume familiarity with the basic theory of modal (fixpoint)
logic, monadic second-order logic, coalgebra, coalgebraic modal (fixpoint)
logic, and parity games.
Here we fix some notation and terminology.
\subsection{Kripke models and their logics}
We restrict to the theory of modal logic with one modality (and hence, one
accessibility relation).
Let $\mathit{Var}$ be a fixed infinite supply of variables.
A \textit{Kripke model} is a structure $\mathbb{S} = (S,R,V)$ where $S$ is a set,
$R \subseteq S \times S$ and $V : \mathit{Var} \rightarrow \mathcal{P}(S)$ is a
$\mathit{Var}$-valuation.
Associated with such a valuation $V$, we define the \emph{conjugate coloring}
$V^{\dagger}: S \to \mathcal{P}(\mathit{Var})$ by $V^{\dagger}(s) \mathrel{:=} \{ p \in \mathit{Var} \mid
s \in V(p)\}$.
Given a subset $T \subseteq S$, the valuation $V[p \mapsto T]$ is as $V$ except
that it maps the variable $p$ to $T$.
A \textit{pointed} Kripke model is a structure $(\mathbb{S},u)$ where $\mathbb{S}$ is a
Kripke model and $u$ is a point in $\mathbb{S}$.
Turning to syntax, we define the formulas of monadic second-order logic $\mathlang{MSO}$
through the following grammar:
$$
\varphi \mathrel{::=} \mathtt{sr}(p) \mathrel{|} p \subseteq q \mathrel{|} R(p,q)
\mathrel{|} \neg \varphi \mathrel{|}\varphi \vee\varphi \mathrel{|} \exists p. \varphi,
$$
with $p,q \in \mathit{Var}$.
Formulas are evaluated over pointed Kripke models by the following induction:
\begin{itemize}
\item $(S,R,V,u) \vDash \mathtt{sr}(p) $ iff $V(p) = \{u\}$
\item $(S,R,V,u)\vDash p \subseteq q $ iff $V(p) \subseteq V(q)$
\item $(S,R,V,u) \vDash R(p,q)$ iff for all $v \in V(p)$ there is $w \in V(q)$
with $v R w$
\item standard clauses for the boolean connectives
\item $(S,R,V,u)\vDash \exists p. \varphi$ iff $(S,R,V[p \mapsto T],u)\vDash
\varphi$ for some $T \subseteq S$.
\end{itemize}
We present the language of the modal $\mu$-calculus $\mathlang{\mu ML}$ in
negation normal form, by the following grammar:
$$ \varphi \mathrel{::=} p \mathrel{|} \neg p \mathrel{|} \bot \mathrel{|} \top
\mathrel{|} \phi \lor \phi \mathrel{|} \phi \land \phi
\mathrel{|} \Box \varphi \mathrel{|} \Diamond \varphi
\mathrel{|} \eta p. \varphi
$$
where $p \in \mathit{Var}$, $\eta \in \{ \mu, \nu \}$, and in the formula $\eta p. \varphi$
no free occurrence of the variable $p$ may be in the scope of a negation.
The satisfaction relation between pointed Kripke models and formulas in
$\mathlang{\mu ML}$ is defined by the usual induction, with, e.g.
\begin{itemize}
\item
$(S,R,V,u)\vDash \mu p. \varphi$ iff $u \in \bigcap \{Z \subseteq S
\mid \varphi_{p}(Z) \subseteq Z\}$
where $\varphi_{p}(Z)$ denotes the truth set of the formula $\varphi$ in the
model $(S,R,V[p \mapsto Z])$.
\end{itemize}
We assume familiarity with the notion of bisimilarity between two (pointed)
Kripke models, and say that a formula of $\mathlang{MSO}$ is \textit{bisimulation
invariant} if it has the same truth value in any pair of bisimilar pointed
Kripke models.
\begin{fact}\cite{jani:expr96}
A formula $\varphi$ of $\mathlang{MSO}$ is equivalent to a formula of $\mathlang{\mu ML}$ iff
$\varphi$ is invariant for bisimulations.
\end{fact}
\subsection{Coalgebras and models}
Our basic semantic structures consist of coalgebras together with valuations.
We only consider coalgebras over the base category $\mathbf{Set}$ with sets as
objects and functions as arrows.
The co- and contravariant power set functors will be denoted by $\mathcal{P}$ and
$\mathcal{Q} : \mathbf{Set} \to \mathbf{Set}^{op}$, respectively.
Covariant endofunctors on $\mathbf{Set}$ will be called \textit{set
functors}.
\begin{defi}
Let $\mathsf{T}$ be a set functor.
A \textit{$\mathsf{T}$-coalgebra} is a pair $(S,\sigma)$ consisting of a set $S$,
together with a map $\sigma : S \to \mathsf{T} S$.
A $\mathsf{T}$-\textit{model} is a structure $\mathbb{S} = (S,\sigma,V)$ where
$(S,\sigma)$ is a $\mathsf{T}$-coalgebra and $V : \mathit{Var} \rightarrow \mathcal{P} S$.
A \textit{pointed} $\mathsf{T}$-model is a structure $(\mathbb{S},s)$ where
$\mathbb{S}$ is a $\mathsf{T}$-model and $s \in S$.
\end{defi}
The usual notion of a $p$-morphism between Kripke models can be generalized as
follows:
Let $\mathbb{S}_1 = (S_1,\sigma_1,V_1)$ and $\mathbb{S}_2 = (S_2,\sigma_2,V_2)$
be two $\mathsf{T}$-models and let $f : S_1 \rightarrow S_2$ be any map.
Then $f$ is said to be a $\mathsf{T}$-\textit{model homomorphism} if:
\begin{enumerate}
\item
for each variable $p$ and each $u \in S_1$, we have $u \in V_1(p)$ iff
$f(u) \in V_2(p)$;
\item
the map $f$ is a \textit{coalgebra morphism}, i.e. we have
$$\sigma_2 \circ f = \mathsf{T} f \circ \sigma_1.$$
\end{enumerate}
Two pointed coalgebras $(\mathbb{S},s)$ and $(\mathbb{S}',s')$ are \emph{behaviorally
equivalent} if $s$ and $s'$ can be identified by coalgebra morphisms $f:
\mathbb{S} \to \mathbb{T}$ and $f': \mathbb{S}' \to \mathbb{T}$ such that $f(s) = f'(s')$.
A \emph{coalgebraic logic} consists of a set $\mathlang{L}$ of \emph{formulas}
together with, for each coalgebra
$(S,\sigma)$, a truth or satisfaction relation ${\Vdash} \subseteq S \times
\mathlang{L}$.
A formula $\phi$ is called \emph{bisimulation invariant}\footnote{Strictly speaking, behavioral equivalence and bisimilarity are distinct concepts. However, in many concrete cases, behavioural equivalence and bisimilarity coincide, so we shall be content to use the more common parlance of ``bisimulation invariance'' rather than ``invariance for behavioural equivalence.}
if $\mathbb{S}, s
\Vdash \phi \iff \mathbb{S}',s' \Vdash \phi$ whenever $\mathbb{S},s \simeq \mathbb{S}', s'$.
Kripke frames are coalgebras for the (covariant) power set functor $\mathcal{P}$.
A functor of particular interest in this paper is the \textit{monotone
neighborhood} functor $\mathcal{M}$, usually defined as the subfunctor of $\mathcal{Q} \circ
\mathcal{Q}$ given by
$$
\mathcal{M} X = \{N \in \mathcal{Q} \mathcal{Q} X \mid \forall Z,Z': Z \in N
\mathrel{\&}
Z \subseteq Z' \Rightarrow Z' \in N\}.$$
This functor comes equipped with the following notion of bisimilarity.
A \emph{neighborhood bisimulation} between $\mathcal{M}$-models $\mathbb{S}_1$ and
$\mathbb{S}_2$ is a relation $R \subseteq S_{1} \times S_{2}$ such that, if
$s_1 R s_2$
then:
\begin{itemize}
\item $V_1^\dagger(s_1) = V_2^\dagger(s_2)$;
\item for all $Z_1$ in $\sigma_1(s_1)$ there is $Z_2 $ in $\sigma_2(s_2)$ such
that, for all $t_2 \in Z_2$ there is $t_1 \in Z_1$ with $t_1 R t_2$;
\item for all $Z_2$ in $\sigma_2(s_2)$ there is $Z_1 $ in $\sigma_1(s_1)$ such
that, for all $t_1 \in Z_1$ there is $t_2 \in Z_2$ with $t_1 R t_2$.
\end{itemize}
\subsection{Coalgebraic $\mu$-calculus \& coalgebra automata}
The modal $\mu$-calculus is just one in a family of logical systems that may
collectively be referred to as the \textit{coalgebraic
$\mu$-calculus}~\cite{cirs:expt09}.
These logics essentially make use of \textit{predicate liftings}.
\begin{defi}
\label{d:pl}
Given a set functor $\mathsf{T}$, an \textit{$n$-place predicate lifting} for $\mathsf{T}$
is a natural transformation
$$\lambda : \mathcal{Q}(-)^n \rightarrow \mathcal{Q} \circ \mathsf{T},
$$
where $\mathcal{Q}(-)^n$ denotes the $n$-fold product of $\mathcal{Q}$ with itself.
A predicate lifting $\lambda$ is said to be \textit{monotone} if
$$
\lambda_X(Y_1,...,Y_n) \subseteq \lambda_X(Z_1,...,Z_n),
$$
whenever $Y_i \subseteq Z_i$ for each $i$.
The \textit{Boolean dual} $\lambda^d$ of $\lambda$ is defined by
$$(Z_1,...,Z_n) \mapsto \mathsf{T} X \setminus
(\lambda_X(X\setminus Z_1,...,X \setminus Z_n)).
$$
\end{defi}
Given a set functor $\mathsf{T}$, the language $\muML_{\fun}$ of the coalgebraic
$\mu$-calculus for $\mathsf{T}$ is defined thus:
$$
\varphi \mathrel{::=} p \mathrel{|} \neg p \mathrel{|} \bot \mathrel{|} \top
\mathrel{|} \lambda(\varphi_1,...,\varphi_n)
\mathrel{|} \varphi \vee \varphi \mathrel{|} \varphi \wedge \varphi
\mathrel{|} \eta p. \varphi
$$
where $p \in \mathit{Var}$, $\lambda$ is any monotone $n$-place predicate lifting for
$\mathsf{T}$, $\eta \in \{\mu,\nu\}$, and, in $\eta p. \varphi$, no free occurrence of
the variable $p$ is in the scope of a negation.
If we restrict the formulas $\lambda(\varphi_1,...,\varphi_n)$ so that $\lambda$
must come frome some distinguished set of liftings $\Lambda$, then we denote the
corresponding sublanguage of $\muML_{\fun}$ by $\muML_{\Lambda}$.
The semantics of formulas in a pointed $\mathsf{T}$-model is defined as follows:
\begin{itemize}
\item
$(\mathbb{S},s)\vDash p $ iff $s \in V(p)$ and $(\mathbb{S},s)\vDash \neg p$ iff
$s \notin V(p)$
\item $(\mathbb{S},s)\vDash \lambda(\varphi_1,...,\varphi_n)$ iff $\sigma(s)\in
\lambda_S(\Vert \varphi_1 \Vert,...,\Vert \varphi_n\Vert)$, where
$\Vert \varphi_i \Vert = \{t \in S \mid (\mathbb{S},t)\vDash \varphi_i\}$
denotes the ``truth set'' of $\varphi_i$ in $\mathbb{S}$
\item standard clauses for the boolean connectives
\item $(\mathbb{S},s)\vDash \mu p. \varphi$ iff
$s \in \bigcap \{X \subseteq S \mid \varphi_{p}(X) \subseteq X\}$,
where $\varphi_{p}(Z)$ denotes the truth set of the formula $\varphi$ in the
$\mathsf{T}$-model $(S,\sigma,V[p \mapsto Z])$.
\end{itemize}
It is routine to prove that all formulas in $\muML_{\fun}$ are bisimulation
invariant.
Turning to the parity automata corresponding to the language $\muML_{\Lambda}$, we
first define the \textit{modal one-step language} $\mathlang{ML}_{\Lambda}^{1}$.
Its set $\mathlang{ML}_{\Lambda}^{1}(A)$ of \emph{modal one-step formulas} over a set $A$ of
variables is given by the following grammar:
$$
\varphi \mathrel{::=}
\bot \mathrel{|} \top \mathrel{|} \lambda(\psi_1,...,\psi_n)
\mathrel{|} \varphi \vee \varphi \mathrel{|} \varphi \wedge \varphi
$$
where $\psi_1,...,\psi_n$ are formulas built up from variables in $A$ using
disjunctions and conjunctions.
\begin{defi}
Given a functor $\mathsf{T}$ and a set of variables $A$, a \textit{one-step model}
over $A$ is a triple $(X,\alpha,V)$ where $X$ is any set, $\alpha \in \mathsf{T} X$
and $V : A \rightarrow \mathcal{P}(X)$ is a valuation.
\end{defi}
The semantics of formulas in the modal one-step language in a one-step model is
given as follows:
\begin{itemize}
\item standard clauses for the boolean connectives,
\item $(X,\alpha,V) \vDash_1 \lambda(\psi_1,...,\psi_n)$ iff
$\alpha \in \lambda_X(\Vert \psi_1\Vert,...,\Vert \psi_n \Vert)$, where
$\Vert \psi_i \Vert \subseteq X$ is the (classical) truth set of the formula
$\psi_i$ under the valuation $V$.
\end{itemize}
We can now define the class of automata used to characterize the coalgebraic
$\mu$-calculus.
\begin{defi}
Let $P$ be a finite set of variables and $\Lambda$ a set of predicate liftings.
Then a \textit{($P$-chromatic) modal $\Lambda$-automaton} is a tuple
$(A,\Delta,\Omega,a_I)$ where $A$ is a finite set of states with $a_I \in A$,
\[
\Delta : A \times \mathcal{P}(P) \rightarrow \mathlang{ML}_{\Lambda}^{1}(A)
\]
is the transition map of the automaton, and $\Omega : A \rightarrow \omega$ is
the parity map.
The class of these automata is denoted as $\mathit{Aut}(\mathlang{ML}_{\Lambda})$.
\end{defi}
The acceptance game for an automaton $\mathbb{A} = (A,\Delta,\Omega,a_I)$ and
a $\mathsf{T}$-model $(S,\sigma,V)$ is given by the following table:
\begin{table}[h]
\centering
\begin{tabular}{|l|c|l|}
\hline
Position & Pl'r & Admissible moves
\\ \hline
$(a,s)\in A\times S$
& $\exists$
& $\{U:A\rightarrow {\mathcal{P}}S \mid
(S,\sigma(s),U) \vDash_{1}\Delta(a, V^{\dagger}(s))$
\\
$U:A\rightarrow{\mathcal{P}}S$
& $\forall$
& $\{(b,t)\mid t\in U(b)\}$ \\
\hline
\end{tabular}
\end{table}
The loser of a finite match is the player who got stuck, and the
winner of an infinite match is $\exists$ if the greatest parity that
appears infinitely often in the match is even, and the winner is
$\forall$ if this parity is odd.
The automaton $\mathbb{A}$ \emph{accepts} the pointed model $(\mathbb{S},s)$
if $\exists$ has a winning strategy in the acceptance game from the starting
position $(a_I,s)$.
We say that and automaton $\mathbb{A}$ is \emph{equivalent} to a formula
$\varphi \in \muML_{\Lambda}$ if, for every pointed $\mathsf{T}$-model
$(\mathbb{S},s)$, we have that $\mathbb{A}$ accepts $(\mathbb{S},s)$ iff
$(\mathbb{S},s)\vDash \varphi$.
\begin{fact}\cite{font:auto10}
\label{coalgebraicmu}
Let $\mathsf{T}$ be a set functor, and $\Lambda$ a set of monotone predicate
liftings for $\mathsf{T}$, closed under Boolean duals.
Then
\[
\muML_{\Lambda} \equiv \mathit{Aut}(\mathlang{ML}_{\Lambda}).
\]
That is, there are effective transformations of formulas in $\muML_{\Lambda}$ into
equivalent automata in $\mathit{Aut}(\mathlang{ML}_{\Lambda})$, and vice versa.
\end{fact}
\section{Coalgebraic $\mathlang{MSO}$}
\label{sec:mso}
We now introduce coalgebraic monadic second-order logic for a set functor $\mathsf{T}$
and a set of liftings $\Lambda$ and show how $\mathlang{MSO}$ can be recovered as a special
case.
We define the syntax of the monadic second-order logic $\mathlang{MSO}_\mathsf{T}$ by the
following grammar:
\[
\varphi \mathrel{::=}
\bot \mathrel{|} \mathtt{sr}(p) \mathrel{|} p \subseteq q
\mathrel{|} \lambda(p,q_1,..,q_n) \mathrel{|}
\varphi \vee \varphi \mathrel{|} \neg \varphi \mathrel{|} \exists p. \varphi
\]
where $\lambda$ is any $n$-place monotone predicate lifting and $p,q,q_1,...,q_n \in Var$.
More generally, restricting to a set $\Lambda$ of monotone liftings for $\mathsf{T}$,
we define the sublanguage $\MSO_{\Lambda} \subseteq \MSO_{\fun}$ by the same grammar
except that we require the liftings to be in $\Lambda$.
For the semantics, let $(\mathbb{S},s)$ be a pointed $\mathsf{T}$-model.
We define the satisfaction relation ${\vDash} \subseteq S \times
\mathlang{MSO}_\mathsf{T}$ as follows:
\begin{itemize}
\item $(\mathbb{S},u) \vDash \mathtt{sr}(p)$ iff $V(p) = \{u\}$,
\item $(\mathbb{S},u)\vDash p \subseteq q$ iff $V(p) \subseteq V(q)$,
\item $(\mathbb{S},u) \vDash \lambda(p, q_1,...,q_n)$ iff $\sigma (v)\in
\lambda_S(V(q_1),..,V(q_n))$ for all $ v \in V(p)$,
\item standard clauses for the Boolean connectives
\item $(\mathbb{S},u) \vDash \exists p .\varphi$ iff
$(S,\sigma,V[p \mapsto Z],u)\vDash \varphi$, some $Z \subseteq S$.
\end{itemize}
\noindent
We introduce the following abbreviations:
\begin{itemize}
\item $p = q$ for $p \subseteq q \wedge q \subseteq p$,
\item $\mathtt{Em}(p)$ for $\forall q. (q \subseteq p \rightarrow q = p)$,
\item $\mathtt{Sing}(p)$ for $\neg \mathtt{Em} (p) \wedge
\forall q (q \subseteq p \rightarrow (em(q) \vee q = p))$
\end{itemize}
expressing, respectively, that $p$ and $q$ are equal, that $p$ denotes the
empty set, and that $p$ denotes a singleton.
Clearly, standard $\mathlang{MSO}$ is the logic $\mathlang{MSO}_{\{\Diamond\}}$, where $\Diamond$
is the predicate lifting corresponding to the usual diamond modality over
Kripke models.
Obviously then, $\mathlang{MSO}_\mathcal{P}$ contains $\mathlang{MSO}$.
In order to see that the languages are in fact equivalent in expressive power,
we need the notion of \emph{expressive completeness}, which plays an important
role in this paper.
\begin{defi}
A set of monotone liftings $\Lambda$ for a set functor $\mathsf{T}$ is said to be
\textit{expressively complete} if, for every finite set of variables $A$ and
every monotone predicate lifting $\lambda : \mathcal{Q}(-)^A \to \mathcal{Q} \circ \mathsf{T} $,
there exists a formula $\varphi \in \mathlang{ML}_{\Lambda}^1(A)$ such that, for every one-step
model $(X,\alpha,V)$ with $V : A \to \mathcal{Q} (X)$, we have
$$
(X,\alpha,V)\vDash_1 \varphi \text{ iff } \alpha \in \lambda_X(V).
$$
\end{defi}
If $\Lambda$ is expressively complete, then clearly $\muML_{\Lambda}$ is equivalent in
expressive power to the full language $\muML_{\fun}$.
It is not much harder to show that, under the same conditions, $\MSO_{\Lambda}$ is
equivalent in expressive power to the full language $\mathlang{MSO}_\mathsf{T}$.
Furthermore, expressive completeness can often be obtained fairly easily if
we make use of an application of the Yoneda lemma to represent $n$-place
predicate liftings as subsets of $\mathsf{T} (2^n)$, a method developed
in~\cite{schr:expr08}.
In particular, since the liftings $ \{\Box,\Diamond\}$ for $\mathcal{P}$ are
expressively complete and $\Box$ is clearly definable in $\mathlang{MSO}_{\{\Diamond\}}$,
one can show that $\mathlang{MSO} = \mathlang{MSO}_{\{\Diamond\}}$ is equivalent in expressive
power to the full coalgebraic logic $\mathlang{MSO}_{\mathcal{P}}$.
Furthermore, $\mathlang{\mu ML}_{\mathcal{P}}$ is equivalent to $\mathlang{\mu ML}_{\{\Box,\Diamond\}}$.
As a second example, involving the monotone neighborhood functor $\mathcal{M}$, let
$\Box$ here be the predicate lifting defined by $\alpha \in \Box_X(Z)$ iff
$Z \in \alpha$, and let $\Diamond$ be its dual.
Then the language $\mathlang{MSO}_\mathcal{M}$ is equivalent to $\mathlang{MSO}_{\{\Box,\Diamond\}}$, and
also $\mathlang{\mu ML}_{\mathcal{M}}$ is equivalent to $\mu \mathlang{ML}_{\{\Box,\Diamond\}}$.
Finally, as mentioned in the introduction, the key question in this paper will
be to compare the expressive power of coalgebraic monadic second-order logic
to that of the coalgebraic $\mu$-calculus.
The following observation, of which the (routine) proof is omitted, provides
the easy part of the link.
\begin{prop}
\label{p:mu-to-mso}
Let $\Lambda$ be a set of monotone predicate lfitings for the set functor $\mathsf{T}$.
There is an inductively defined translation $(\cdot)^{\diamond}$ mapping any
formula $\phi \in \muML_{\Lambda}$ to an equivalent formula $\phi^{\diamond}\in \MSO_{\Lambda}$.
\end{prop}
\section{Automata for coalgebraic $\mathlang{MSO}$}
\label{sec:aut}
In this section, we introduce automata for coalgebraic monadic second-order
logic.
\subsection{A general perspective on parity automata}
Standard monadic second-order formulas can be translated to equivalent automata
over \textit{trees}, but this equivalence is not guaranteed to extend to
arbitrary Kripke models.
In the case of general coalgebra, we should expect having to introduce a
coalgebraic concept of ``tree-like'' models.
\begin{defi}
Given a set $S$ and $\alpha \in \mathsf{T} S$, a subset $X \subseteq S$ is said to
be a \textit{support} for $\alpha$ if there is some $\beta \in \mathsf{T} X$ with
$\mathsf{T} \iota_{X,S}(\beta) = \alpha$. A \textit{supporting Kripke frame} for a
$\mathsf{T}$-coalgebra $(S,\sigma)$ is a binary relation $R \subseteq S \times S$
such that, for all $u \in S$, $R(u) = \{v \mid u R v\}$ is a support for
$\sigma(u)$.
\end{defi}
\begin{defi}
A $\mathsf{T}$-\textit{tree model} is a structure $(\mathbb{S},R,u)$ where
$\mathbb{S} = (S,\sigma,V)$ is a $\mathsf{T}$-model and $u \in S$, such that $R$ is
a supporting Kripke frame for the coalgebra $(S,\sigma)$, and furthermore
$(S,R)$ is a tree rooted at $u$, so that there is a unique $R$-path from $u$
to $w$ for each $w \in S$.
\end{defi}
Our goal is to translate formulas in $\MSO_{\fun}$ to equivalent automata over
$\mathsf{T}$-tree models.
We start by introducing a very general type of automaton, originating
with~\cite{vene:expr14}.
\begin{defi}
Given a finite set $A$, a \textit{generalized predicate lifting} over $A$
comprises an assignment of a map
$$\varphi_X : (\mathcal{Q} X)^A \rightarrow \mathcal{Q} \mathsf{T} X.
$$
to every set $X$.
Concepts like \emph{Boolean dual} and \emph{monotonicity} apply to these
liftings in the obvious way.
\end{defi}
The difference with respect to standard predicate liftings is that the
components of a generalized predicate lifting do not need to form a natural
transformation.\footnote{%
In the style of abstract logic, it would make sense to require a
general predicate lifting to be natural with respect to certain
maps, in particular, bijections.
For the purpose of this paper such a restriction is not needed,
however.
}
\begin{defi}
A \textit{one-step language} $\mathlang{L}$ consists of a collection
$\mathlang{L}(A)$ of generalized predicate liftings for every finite set $A$.
The semantics of a generalized predicate lifting $\varphi$ in a one-step model
$(X,\alpha,V)$ is given by
$$
(X,\alpha,V)\vDash_1 \varphi \textit{ iff } \alpha \in \varphi_X(V).
$$
\end{defi}
Our automata will be indexed by a (finite) set of variables involved,
corresponding to the set of free variables of the $\MSO_{\fun}$-formula.
\begin{defi}
Let $P \subseteq \mathit{Var}$ be a finite set of variables and let $\mathlang{L}$ be a
one-step language for functor $\mathsf{T}$.
A \textit{($P$-chromatic) $\mathlang{L}$-automaton} is a structure
$(A,\Delta,\Omega,a_I)$ where
\begin{itemize}
\item $A$ is a finite set, with $a_I \in A$,
\item $\Omega : A \rightarrow \omega$ is a parity map, and
\item $\Delta : A \times \mathcal{P}(P) \rightarrow \mathlang{L}(A) $ is the
transition map of $\mathbb{A}$.
\end{itemize}
\end{defi}
The \textit{acceptance game} of $\mathbb{A}$ with respect to a $\mathsf{T}$-tree model
$(T,R,\sigma,V,u)$ is given by Table~\ref{table:accgame}.
We say that the automaton $\mathbb{A}$ accepts the model $(T,R,\sigma,V,u)$
if $\exists$ has a winning strategy in this game (initialized at position
$(a_{I},u)$).
\begin{table*}[ht]
{\normalsize
\centering
\begin{tabular}{|l|c|l|c|}
\hline
Position & Player & Admissible moves & Parity \\
\hline
$(a,s) \in A \times T$
& $\exists$
& $\{U : A \to \mathcal{P}(R(s)) \mid $ & \\
& & $(R(s),\sigma(s), U)
\vDash_{1} \Delta(a,V^{\dag}(s)) \}$
& $\Omega(a)$
\\
$U : A \rightarrow \mathcal{P}(T)$
& $\forall$
& $\{(b,t) \mid t \in U(b) \}$
& $0$
\\ \hline
\end{tabular}
\caption{\small Acceptance game for parity automata.}
\label{table:accgame}
}
\end{table*}
\subsection{Closure properties}
This abstract level is useful for establishing some simple closure properties
of automata, based on properties of the one-step language.
The first, easy, results establish sufficient conditions for closure under
union and complementation.
\begin{prop}
\label{closureunion}
If the one-step language $\mathlang{L}$ is closed under disjunction,
then the class of $\mathlang{L}$-automata is closed under union.
\end{prop}
\begin{prop}
\label{closurecomplementation}
If the monotone fragment of the one-step language $\mathlang{L}$ is closed under
Boolean duals, then the class of $\mathlang{L}$-automata is closed under
complementation.
\end{prop}
The most interesting property concerns closure under existential projection.
The following terminology is taken from~\cite{jani:expr96}, but instead
of relying on a particular syntactic shape of one-step formulas, we define
the concepts in purely semantic terms.
\begin{defi}
A predicate lifting $\varphi$ over $A$ is said to be \textit{special basic} if,
for every one-step model $(X,\alpha,V)$ such that
$$(X,\alpha,V)\vDash_1 \varphi$$
there is a valuation $V^* : A \rightarrow \mathcal{Q}(X)$ such that
\begin{itemize}
\item $V^\ast(a) \subseteq V(a)$ for each $a \in A$,
\item $V^\ast(a) \cap V^\ast(b) = \emptyset$ whenever $a \neq b$, and
\item $(X,\alpha,V^*) \vDash_1 \varphi$.
\end{itemize}
Call an $\mathlang{L}$-automaton \textit{non-deterministic} if every lifting
$\Delta(a,c)$ is special basic.
\end{defi}
It is easy to see that if the language $\mathlang{L}$ is closed under
disjunctions, then so is its fragment of special basic liftings.
From this we obtain the following.
\begin{prop}
\label{existentialclosure}
If the one-step language $\mathlang{L}$ is closed under disjunction, then the
class of non-deterministic $\mathlang{L}$-automata is closed under existential
projection over $\mathsf{T}$-tree models.
\end{prop}
\begin{proof}
Suppose $\mathbb{A} = (A,\Delta,a_I,\Omega)$ is a non-deterministic
$\mathlang{L}$-automaton for the variable set $P$.
Define the $P\setminus q$-chromatic automaton
$\exists q.\mathbb{\mathbb{A}} = (A,\Delta^*,a_I,\Omega)$ by setting
$$
\Delta^*(a,c) = \Delta(a,c) \vee \Delta(a,c\cup\{q\}).
$$
It is easy to see that every $\mathsf{T}$-tree model accepted by $\mathbb{A}$ is
also accepted by $\exists p. \mathbb{A}$.
Conversely, suppose $\exists p.\mathbb{A}$ accepts some $\mathsf{T}$-tree
model $(S,R,\sigma,V,s_I)$. For each winning position $(a,s)$ in the
acceptance game, let $V_{(a,s)}$ be the valuation chosen by $\exists$
according to some given winning strategy $\chi$. Note that we can assume that $\chi$ is a \textit{positional} winning strategy, since $\exists p. \mathbb{A}$ is a parity automaton.
It is not difficult to see that the
automaton $\exists p .\mathbb{A}$ is a non-deterministic automaton,
and so for each winning position $(a,s)$ there is a valuation $V_{(a,s)}^* :
A \rightarrow \mathcal{P}(R(s))$, which is an admissible move for $\exists$, such that
$V_{(a,s)}^*(b) \subseteq V_{(a,s)}(b)$ and such that for all $b_1 \neq b_2
\in A$ we have $V_{(a,s)}^*(b_1) \cap V_{(a,s)}^*(b_2) = \emptyset$.
Define the strategy $\chi^*$ by letting $\exists$ choose the valuation
$V_{(a,s)}^*$ at each winning position $(a,s)$ - this is still a winning
strategy, since the valuations chosen by $\exists$ are smaller and so no new
choices for $\forall$ are introduced. Furthermore, $\chi^*$ is clearly still a positional winning strategy.
From these facts follow by a simple induction on the height of the nodes in the supporting
tree that the strategy $\chi^*$ is \textit{scattered}, i.e. that for every $s \in S$
there is at most one automaton state $a$ such that $(a,s)$ appears in a
$\chi^*$-guided match of the acceptance game.
So we can define a valuation $V^\prime$ like $V$ except we evaluate $q$ to be
true at all and only the states $s$ such that
$$
(R(s),\sigma(s), V^*_{(a_s,s)})\vDash_1 \Delta(a_s,c \cup \{q\}),
$$
where $a_s$ is a necessarily \emph{unique} automaton state such that $(a,s)$
appears in some $\chi^*$-guided match, and $c$ is the color consisting of the
variables true under $V$ at $s$.
It is not hard to show that $\mathbb{A}$ accepts $(S,R,\sigma,V',s_I)$.
\end{proof}
\subsection{Second-order automata}
We now introduce a more concrete one-step language for a given set functor
$\mathsf{T}$ and a given set of (natural) liftings $\Lambda$, and show that
$\MSO_{\Lambda}$ can be translated into the corresponding class of automata.
Let $\Lambda$ be a set of monotone predicate liftings for $\mathsf{T}$.
The set of \textit{second-order one-step formulas} over any
set of variables $A$ and relative to the set of liftings $\Lambda$
is defined by the grammar:
$$\varphi \mathrel{::=} a \subseteq b \mathrel{|} \lambda(a_1,...,a_n) \mathrel{|} \neg \varphi
\mathrel{|} \varphi \lor \varphi
\mathrel{|} \exists a.\varphi,
$$
where $a,b,a_1,...,a_n \in A$ and $\lambda$ is any predicate lifting in
$\Lambda$.
Fixing an infinite set of ``one-step variables'' $Var_1$, and given a finite set
$A$, the set of \textit{second-order one-step sentences} over $A$,
denoted $\mathlang{SO}_{\Lambda}^{1}(A)$, is the set of one-step formulas over $A \cup Var_1$,
with all free variables belonging to $A$.
We write $\mathlang{SO}_{\fun}^{1}(A)$ when $\Lambda$ comprises all monotone liftings for $\mathsf{T}$.
The semantics of a one-step second-order $A$-formula in a one-step model
$(X,\alpha,V)$ (with $V : A \to \mathcal{P}(X)$ is defined by the following clauses:
\begin{itemize}
\item $(X,\alpha,V)\vDash_1 p \subseteq q$ iff $V(p) \subseteq V(q)$,
\item $(X,\alpha,V)\vDash_1 \lambda (p_1,...,p_n)$ iff $\alpha \in
\lambda_X(V(p_1),...,V(p_n))$,
\item standard clauses for the Boolean connectives,
\item $(X,\alpha,V)\vDash_1 \exists p. \varphi$ iff
$(X,\alpha,V[p\mapsto S])\vDash_1 \varphi$ for some $S \subseteq X$.
\end{itemize}
Any one-step second-order $A$-sentence $\phi$ can be regarded as a generalized
predicate lifting over $A$, with
$$
\varphi_X(V) = \{\alpha \in \mathsf{T} X \mid (X,\alpha,V) \vDash_1 \varphi\}.
$$
Note that the syntax of $\mathlang{SO}_{\fun}^{1}$ allows negations, implying that not all
these predicate liftings are monotone.
\begin{defi}
Let $\Lambda$ be a set of monotone predicate liftings for $\mathsf{T}$.
A \textit{second-order $\Lambda$-automaton} is an $\mathlang{L}$-automaton for
$\mathlang{L}$ being the assignment of the one-step second-order $A$-sentences
$\mathlang{SO}_{\Lambda}^{1}(A)$ to every set of variables $A$.
We write $\mathit{Aut}(\mathlang{SO}_{\Lambda})$ to denote this class, and $\mathit{Aut}(\mathlang{SO}_{\fun})$ in case $\Lambda$ is
the set of \emph{all} monotone predicate liftings for $\mathsf{T}$.
\end{defi}
Our aim is to prove that every formula of $\MSO_{\Lambda}$ can be
translated into an equivalent second-order $\Lambda$-automaton (over rooted
$\mathsf{T}$-tree models), and the main problem here is to obtain closure under
existential projection.
The key to this step is a simulation theorem.
First, a useful trick due to Walukiewicz \cite{walu:mona96} allows us to
transform any second-order automaton into one in which all the one-step formulas are monotone, when regarded as generalized predicate liftings. We call such an automaton a \textit{monotone} automaton.
\begin{prop}
\label{p:mon-aut}
Let $\Lambda$ be any set of monotone predicate liftings.
Every automaton $\mathbb{A} \in \mathit{Aut}(\mathlang{SO}_{\Lambda})$ is equivalent to a monotone second-order
$\mathbb{A} \in \mathit{Aut}(\mathlang{SO}_{\Lambda})$.
\end{prop}
\begin{proof}
Enumerate $A$ as $\{a_1,...,a_k\}$, and just replace each formula $\Delta(a,c)$
by
$$
\exists Z_1 ... \exists Z_k. Z_1 \subseteq a_1 \wedge ... \wedge
Z_k \subseteq a_k \wedge \Delta(a,c)[Z_i / a_i]
$$
where $\Delta(a,c)[Z_i/a_i]$ is the result of substituting the variable
$Z_i$ for each open variable $a_i$ in $\Delta(a,c)$.
This new formula is monotone in the variables $A$ and the resulting automaton
is equivalent to $\mathbb{A}$.
\end{proof}
The intuition behind the simulation theorem is the same as that behind the
standard ``powerset construction'' for word automata: the states of the new
non-deterministic automaton $\mathbb{A}_n$ are ``macro-states'' representing
several possible states of $\mathbb{A}$ at once.
Formally, the states of $\mathbb{A}_n$ will be binary relations over $A$, and
given a macro-state $R$, its range gives an exact description of the states in
$\mathbb{A}$ that are currently being visited simultaneously. It is safe to
think of the macro-states as subsets of $A$, however: the only reason that we
have binary relations over $A$ as states rather than just subsets is to have
a memory device so that we can keep track of traces in infinite matches.
For each macro-state $R$ and each colour $c$ we want to be able to say that the
one-step formulas corresponding to each state in the range of $R$ hold, so we
want to translate the one-step formulas over $A$ into one-step formulas over
the set of macro-states. In order to translate a formula $\Delta(a,c)$ to a
new one-step formula with macro-states as variables, we have to replace the
variable $b$ in $\Delta(a,c)$ with a new variable that acts as a stand-in for
$b$. For this purpose we introduce a new, existentally quantified variable
$Z_b$, together with a formula stating explicitly that $Z_b$ is to represent
the union of the values of all those macro states that contain $b$.
Furthermore we want all the one-step formulas to be special basic, and for
this purpose we simply add a conjunct ``$\mathsf{disj}$'' to each one-step
formula, stating that the values of any pair of distinct variables appearing
in the formula are to be disjoint. Finally, in order to turn $\mathbb{A}_n$
into a parity automaton, we use a stream automaton to detect bad traces (see
for instance \cite{vene:lect12} for the details in a more specific case).
\begin{theo}[Simulation]
\label{simulationtheorem}
Let $\Lambda$ be a set of monotone predicate liftings for $\mathsf{T}$.
For any monotone automaton $\mathbb{A} \in \mathit{Aut}({\mathlang{SO}_{\Lambda}})$ there exists an equivalent
non-deterministic $\mathbb{A}'\in \mathit{Aut}({\mathlang{SO}_{\Lambda}})$.
\end{theo}
Given a set $A$, we consider the set $\mathcal{P}(A \times A)$ as a set of variables.
Let
$$\mathsf{disj} : = \bigwedge_{B \neq B' \subseteq A\times A} \forall X . (X \subseteq B \wedge X \subseteq B') \rightarrow \mathtt{Em}(X)$$
Pick a fresh variable $Z_a$ for each $a \in A$. Given a 1-step formula $\varphi$, let
$$\varphi[Z_a/ a]$$ be the result of substituting $Z_a$ for each free variable $a \in A$ in $\varphi$.
If we enumerate the elements of $A$ as $a_1,...,a_k$, we now define the formula $\varphi^{\uparrow b}$ for $b \in A$ to be
\begin{displaymath}
\begin{array}{lcl}
& \exists Z_{a_1}...\exists Z_{a_k}. \bigwedge_{1 \leq i \leq k} ( \mathsf{Eq}(Z_{a_i} \bigcup\{B^\prime \mid (b,a_i) \in B^\prime \})\; \wedge \\
& \varphi[Z_a/ a] )
\end{array}
\end{displaymath}
where $\mathsf{Eq}(Z_{a_i} \bigcup\{B^\prime \mid (b,a_i) \in B^\prime \})$ is a formula asserting that the value of the variable $Z_{a_i}$ is the union of the values of all variables $B^\prime$ with $(b,a_i)\in B^\prime$.
Let $\mathbb{A} = (A,\Delta,a_I,\Omega)$ be any monadic $\Lambda$-automaton. We can assume w.l.o.g. that $\mathbb{A}$ is monotone. We first construct the automaton $\mathbb{A}_n = (A_n,\Delta_n,a_I^*,F)$ with a non-parity acceptance condition $F \subseteq (A_n)^\omega$ as follows:
\begin{itemize}
\item $A_n = \mathcal{P}(A \times A)$
\item $\Delta_n(B,c) = \mathsf{disj} \wedge \bigwedge_{b \in \pi_2[B]}\Delta(b,c)^{\uparrow b}$
\item $a_I^\ast = \{(a_I,a_I)\}$
\item $F$ is the set of streams over $\mathcal{P}(A \times A)$ with no bad traces.
\end{itemize}
Here, $\pi_2$ is the second projection of a relation $B$ so that $\pi_2[B]$ denotes the range of $B$. A \textit{trace} in a stream $(B_1,B_2,B_3,...)$ over $\mathcal{P}(A \times A)$ is a stream $(a_1,a_2,a_3,...)$ over $A$ with $a_1 \in \pi_2[B_1]$ and $(a_{j-1},a_j) \in B_j$ for $j > 1$. A trace is \textit{bad} if the greatest number $n$ with $\Omega(a_i) = n$ for infinitely many $a_i$ is odd.
The new automaton $\mathbb{A}_n$ is clearly special basic, because of the $\mathsf{disj}$-formulas.
\begin{lemma}
$\mathbb{A}_n$ is equivalent to $\mathbb{A}$, provided that $\mathbb{A}$ is monotone
\end{lemma}
\begin{proof}
Fix a pointed $\mathsf{T}$-tree model $(\mathbb{S},R,s_I)$ where $\mathbb{S} = (S,\sigma,V)$. We want to show that $\mathbb{A}$ accepts $(\mathbb{S},R,s_I)$ if and only if $\mathbb{A}_n$ does. That is, we want to show that the languages $L(\mathbb{A})$ and $L(\mathbb{A}_n)$ defined by these two automata are the same.
\subsubsection*{First part: $L(\mathbb{A}) \subseteq L(\mathbb{A}_n)$}
Suppose first that $\mathbb{A}$ accepts $(\mathbb{S},R,s_I)$. Let $\chi$ be a positional winning strategy for $\exists$ in the acceptance game, mapping each winning position $(a,s)$ to a valuation $U : A \rightarrow \mathcal{Q}(R(s))$ such that
$$(R(s),\sigma(s),U) \vDash_1 \Delta(a,V^\dagger(s))$$
Such a strategy exists since $\mathbb{A}$ is a parity automaton, and so the acceptance game is a parity game.
We define the winning strategy $\chi^*$ for $\exists$ in the acceptance game for $\mathbb{A}_n$ as follows: given a position $(B,s)$, define the function $f_{B,s} : R(s) \rightarrow \mathcal{P}(A \times A)$ by setting
$$f_{B,s} (s^\prime) = \{(a,b) \mid a \in \pi_2[B] \;~\&~\;s^\prime \in \chi(a,s)(b)\}$$
At the position $(B,s)$, let $\exists$ choose the following valuation $\chi^*(B,s)$, defined by:
$$\chi^*(B,s)(B^\prime) = \{s^\prime \in R(s) \mid f_{B,s}(s^\prime) = B^\prime\}$$
Our first claim is that, for each position of the form $(B,s)$ where each $(b,s)$ for $b \in \pi_2[B]$ appears in some $\chi$-coherent match of the acceptance game for $\mathbb{A}$ with start position $(a_I,s_I)$, the move for $\exists$ given by the strategy $\chi^*$ is legal. To prove this claim we need to check that, for each position $(B,s)$, we have
$$(R(s),\sigma(s),\chi^*(B,s))\vDash_1 \Delta_n(B,V^\dagger(s))$$
provided each $(b,s)$ for $b\in \pi_2[B]$ appears in some $\chi$-coherent match. First, the formula $\mathsf{disj}$ is true since the marking $\chi^*(B,s)$ is the inverse of a mapping from $R(s)$ to $\mathcal{P}(A \times A)$. We now have to check that,
for each $a^\prime \in \pi_2[B]$ we have
$$(R(s),\sigma(s),\chi^*(B,s)) \vDash_1 \Delta(a^\prime,V^\dagger(s))^{\uparrow a^\prime} $$
We need to find sets $S_{a_1},...,S_{a_k} \subseteq R(s)$ such that $(R(s),\sigma(s),\chi^*(B,s))$ with the assignment $Z_{a_i} \mapsto S_{a_i}$ satisfies the formula
\begin{displaymath}
\begin{array}{lcl}
\bigwedge_{1 \leq i \leq k} \mathsf{Eq} (Z_{a_i}, \bigcup\{B' \mid (a^\prime,a_i) \in B' \} )\; & \wedge & \\
\Delta(a^\prime,V^\dagger(s))[Z_a/a] & &
\end{array}
\end{displaymath}
Since $\chi$ gives a legal move at the position $(a^\prime,s)$ for each $a' \in \pi_2[B]$, each one-step model of the form $(R(s),\sigma(s),\chi(a^\prime,s))$ satisfies the formula $\Delta(a^\prime,V^\dagger(s))$.
Hence, if we assign to each variable $Z_{a_i}$ the set $\chi(a^\prime,s)(a_i)$, then this variable assignment satisfies the formula
$$\Delta(a^\prime,V^\dagger(s))[a \mapsto Z_a \mid a \in A]$$
Since the formula $\Delta(a^\prime,s)$ is monotone in all the variables $A$, the same is true for any larger assignment. So
it now suffices to prove that
$$\chi(a^\prime,s)(a_i) \subseteq \bigcup \{\chi^*(B,s)(B^\prime) \mid (a^\prime,a_i) \in B^\prime\}$$
since we can then safely take
$$S_{a_i} = \bigcup \{\chi^*(B,s)(B^\prime) \mid (a^\prime,a_i) \in B^\prime\}$$
To prove this inclusion, suppose
$s^\prime \in \chi(a^\prime,s)(a_i)$. Let $B^\prime$ be the relation defined by
$$ (d,d^\prime) \in B^\prime \Leftrightarrow d \in \pi_2[B] \;\& \; s^\prime \in \chi(d,s)(d^\prime)$$
Clearly, $(a^\prime,a_i) \in B^\prime$. Moreover, $f_{B,s}(s^\prime) = B^\prime$ by definition, and so $s^\prime \in \chi^*(B,s)(B^\prime)$ as required.
We now show that any $\chi^*$-coherent match with start position $(a_I^*,s_I)$ is winning for $\exists$. We have to prove two things: first, that $\exists$ never gets stuck in a $\chi^*$-coherent match, and second, that $\exists$ wins every infinite $\chi^*$-coherent match, i.e. no infinite $\chi^*$-coherent match contains a bad trace.
First we show that $\exists$ never gets stuck. For this to be the case, all we need to show that if $(B,s)$ is the last position of some $\chi^*$-coherent partial match, then all the positions $(a,s)$ for $a \in \pi_2[B]$ are winning positions for the strategy $\chi$ - by our previous claim this guarantees the move $\chi^*(B,s)$ to be legal. We prove by induction on the length of a finite partial match that this holds for the last position of the match: it holds for $(\{(a_I,a_I)\},s_I)$, clearly, since $\chi$ is a winning strategy at $(a_I,s_I)$. Suppose that the induction hypothesis holds for a finite match with last position $(B,s)$. Let $(B',s')$ be any position such that $s' \in \chi^*(B,s)(B')$. Then
$$B' = f_{B,s}(s^\prime) = \{(a,b) \mid a \in \pi_2[B] \;~\text{and}~ \;s^\prime \in \chi(a,s)(b)\}$$
So suppose $b\in \pi_2[B']$. Then there is some $a$ with $(a,b) \in B^\prime$, and we must have $a \in \pi_2[B]$ and $s^\prime \in \chi(a,s)(b)$. But since the position $(a,s)$ is winning by the inductive hypothesis, this means that $(b,s^\prime)$ is a winning position for $\chi$, and we are done.
We now show that $\exists$ wins every infinite $\chi^*$-coherent match. For this, it suffices to show that every trace $(a_1,a_2,a_3,...)$ in a $\chi^*$-coherent infinite match
$$(B_1,s_1),(B_2,s_2),(B_3,s_3),...$$
corresponds to a $\chi$-coherent match
$$(a_I,s_I) = (a_1,s_1),(a_2,s_2),(a_3,s_3),...$$
So fix a trace $(a_1,a_2,s_3,...)$ meaning that for each $a_i$ we have $(a_i,a_{i + 1}) \in B_{i + 1}$. We have to show that $s_{i + 1} \in \chi(a_i,s_i)(a_{i + 1})$ for each $i$. We have
$$s^{i + 1} \in \chi^*(B_{i},s_i)(B_{i + 1})$$ meaning that $B_{i + 1}$ is equal to
$$ f_{B_i,s_i}(s_{i + 1}) = \{(a,b) \mid a \in \pi_2[B_i] \;\& \;s_{i + 1} \in \chi(a,s_i)(b)\}$$
In particular, since $(a_i,a_{i + 1}) \in B_{i + 1}$, this means we must have $s_{i +1} \in \chi(a_i,s_i)(a_{i + 1})$ as required.
\subsubsection*{Second part: $L(\mathbb{A}_n) \subseteq L(\mathbb{A})$}
Conversely, suppose $\mathbb{A}_n$ accepts $(\mathbb{S},s_I)$ with winning strategy $\chi$. We construct a winning strategy $\chi^*$ for $\exists$ w.r.t $\mathbb{A}$. Note that the strategy $\chi$ is not necessarily positional, since the acceptance game is not a parity game.
By induction on the length of a $\chi^\ast$-coherent partial match
$$M = (a_1,s_1),(a_2,s_2),(a_3,s_3)...(a_{k},s_{k})$$
with $(a_1,s_1) = (a_I,s_I)$,
we are going to define a next legal move $\chi^*(M)$ for $\exists$, and by a simultaneous induction we construct a $\chi$-coherent partial match
$$N = (B_1,s_1),(B_2,s_2),(B_3,s_3)...(B_{k},s_{k})$$
with $(B_1,s_1) = (a_I^*,s_I)$, $a_j \in \pi_2[B_j]$ for each $j$ and $(a_{j- 1},a_{j}) \in B_j$ for each $k \geq j > 1$.
Furthermore we will make sure that whenever a $\chi^*$-coherent match $M$ is an initial segment of a match $M'$, the $\chi$-coherent match associated with $M$ is an initial segment of the $\chi$-coherent match associated with $M'$. It will follow at once that $\chi^*$ is a winning strategy, since $\exists$ never gets stuck in any $\chi^*$-coherent partial match and, furthermore, every infinite $\chi^*$-coherent match corresponds to a trace in some $\chi$-coherent infinite match.
The base case of the induction is the unique match of length 1 with the single position $(a_I,s_I)$, and we take the corresponding position to be $(a_I^* ,s_I)$. Now, suppose $\chi^* $ has been defined on all matches of length $< k$, and let $M$ be a $\chi^*$-coherent partial match of length $k$ of the form
$$(a_1,s_1),(a_2,s_2),(a_3,s_3)...(a_{k},s_{k})$$
By the inductive hypothesis we have a corresponding $\chi$-coherent match $N$ which we write as
$$(B_1,s_1),(B_2,s_2),(B_3,s_3)...(B_{k},s_{k})$$
with $a_k \in \pi_2[B_k]$. Now we define the next legal move $\chi^*(M)$ for $\exists$, and we show that for every position $(a',s')$ such that $s'\in \chi^*(M)(a')$, we can find a relation $B'$ such that $(a_k,a') \in B'$ and
$$ (B_1,s_1),(B_2,s_2),(B_3,s_3)...(B_{k},s_{k}),(B',s')$$
is a $\chi$-coherent match.
Since $N$ is a $\chi$-coherent partial match and $\chi$ is a winning strategy for $\exists$, we have that
$$(R(s_k),\sigma(s_k),\chi(N)) \vDash_1 \Delta_n(B_k,V^\dagger(s_k))$$
Since $a_k \in \pi_2[B_k]$,
this means that there exist sets $S_b \subseteq R(s_k)$ for each $b \in A$ such that the 1-step model $(R(s_k),\sigma(s_k),\chi(N))$ satisfies the formula
\begin{displaymath}
\begin{array}{lcl}
\bigwedge_{b \in A} \mathsf{Eq}(Z_{b}, \bigcup\{B' \mid (a_k,b) \in B' \} )\; & \wedge & \\
\Delta(a_k,V^\dagger(s_k))[Z_b / b] & &
\end{array}
\end{displaymath}
under the assignment $Z_b \mapsto S_b$. Hence, the valuation $U$ defined by
$$U : b \mapsto S_b$$
will be such that
$$(R(s_k),\sigma(s_k),U) \vDash_1 \Delta(a_k,V^\dagger(s_k))$$
So we set $\chi^*(M) = U$, a legal move. Note that we have
$$U(b) = \bigcup\{\chi(N)(B') \mid (a_k,b) \in B'\}$$
Now, let $(a',s')$ be such that $s' \in U(a')$. This means that there is some $B'$ with $(a_k,a')\in B'$ and $s' \in \chi(N)(B')$. Hence, $(B',s')$ satisfies the required conditions, and we are done.
\end{proof}
The only thing left to do at this point is to transform this automaton into one that has its acceptance condition given by a parity map. The set of streams over $\mathcal{P}(A \times A)$ that contain no bad traces w.r.t. the parity map $\Omega$ is an $\omega$-regular stream language, so let
$$\mathbb{Z} = (Z,\delta,z_I,\Omega_z)$$
be a parity stream automaton, $\delta : Z \times \mathcal{P}(A \times A) \rightarrow Z$, that recognizes this language. We now construct the automaton
$$\mathbb{A}_n \odot \mathbb{Z} = (A_n',\Delta_n',a_I',\Omega_n')$$
as follows:
\begin{itemize}
\item $A_n' = A_n \times Z$
\item $a_I' = (a_I^*,z_I)$
\item $\Omega_n ' (B,z) = \Omega(z)$
\item $\Delta_n'((B,z),c) = \Delta_n(B,s)[ (B',\delta(B,z))/ B']$
\end{itemize}
It is not difficult to check that
$\mathbb{A}_n \odot \mathbb{Z}$ is equivalent to $\mathbb{A}_n$.
Since $\mathbb{A}_n\odot \mathbb{Z}$ is clearly still a non-deterministic automaton, this ends the proof of the simulation theorem.
Combining Proposition \ref{existentialclosure} with
Theorem~\ref{simulationtheorem},
we easily obtain the following closure property.
\begin{prop}
\label{p:clos-exist}
Let $\Lambda$ be a set of monotone predicate liftings for a set functor $\mathsf{T}$.
Over $\mathsf{T}$-tree models, the class of second-order $\Lambda$-automata is closed
under existential projection.
\end{prop}
We can now use the closure properties we have established for second-order
automata to give the desired translation of $\MSO_{\fun}$ into second-order automata.
\begin{prop}
\label{p:automatachar}
For every formula $\varphi\in \MSO_{\Lambda}$ with free variables in $P$, there exists
a $P$-chromatic automaton $\mathbb{A}_\varphi \in \mathit{Aut}(\mathlang{SO}_{\Lambda})$ which is
equivalent to $\varphi$ over $\mathsf{T}$-tree models.
\end{prop}
\begin{proof}
Proceeding by a straightforward induction on the complexity of $\phi$, we leave
it to the reader to construct appropriate automata for the atomic formulas.
The inductive cases for disjunction and negation follow by the
Propositions~\ref{closureunion} and~\ref{closurecomplementation}, together with
the easy observation that the one-step language $\mathlang{SO}_{\Lambda}$ is closed under
disjunction and Boolean duals.
The case of existential quantification is taken care of by
Proposition~\ref{p:clos-exist}.
\end{proof}
Theorem~\ref{t:automatachar} is immediate from this, as is the following.
\begin{coro}
\label{c:automatachar}
Suppose $\Lambda$ is any set of monotone predicate liftings for $\mathsf{T}$ such that $\MSO_{\fun} \equiv \mathlang{MSO}_\Lambda$.
Then for every formula of $\MSO_{\fun}$, there exists an equivalent
second-order $\Lambda$-automaton over $\mathsf{T}$-tree models. In particular, this holds whenever $\Lambda$ is expressively complete.
\end{coro}
\section{Bisimulation invariance}
\label{sec:bisinv}
This section continues the program of~\cite{vene:expr14}, making use of the
automata-theoretic translation of $\MSO_{\fun}$ we have just established.
The gist of our approach is that, in order to characterize a coalgebraic
fixpoint logic $\muML_{\fun}$ as the bisimulation-invariant fragment of $\MSO_{\fun}$, it
suffices to establish a certain type of translation between the corresponding
one-step languages.
First we need some definitions.
\begin{defi}
Given sets $X,Y$, a mapping $h : X \rightarrow Y$ and a valuation $V : A
\rightarrow \mathcal{Q} (Y)$, we define the valuation $V_{[h]} : A \rightarrow \mathcal{Q} (X)$ by
setting $V_{[h]}(b) = h^{-1}[V(b)]$ for each $b \in A$.
\end{defi}
The most important concept that we take from $\cite{vene:expr14}$ is that
of a \textit{uniform translation} (called \textit{uniform correspondence}
in \cite{vene:expr14}):
\begin{defi}
Given a functor $\mathsf{T}$, a \textit{uniform construction} $F$ for
$\mathsf{T}$ takes each pair $(X,\alpha)$ with $\alpha \in \mathsf{T} X$ to a tuple consisting
of a set $X_*$, an $\alpha_* \in \mathsf{T} (X_*) $,
and a map $h_\alpha : X_* \rightarrow X $ such that
$$\mathsf{T} (h_\alpha)(\alpha_*) = \alpha$$
We say that the second-order one-step language $\mathlang{SO}_{\Lambda}^{1}(A)$
\textit{admits uniform translations} if, given any natural number $k$, there
exists a uniform construction $F$ and an assignment of a monotone (natural)
predicate lifting
$$\varphi^* : \mathcal{Q}(-)^A \rightarrow \mathcal{Q} \circ \mathsf{T}$$
to each monotone one-step formula $\varphi \in \mathlang{SO}_{\Lambda}^{1}$
with free variables $A$ and quantifier depth at most $k$, such that for any
one-step model $(X,\alpha,V)$, we have
$$(X,\alpha,V) \vDash_1 \varphi^* \text{ iff }
(X_*,\alpha_*,V_{[h_\alpha]})\vDash_1 \varphi.
$$
\end{defi}
\begin{remark}
It is easy to see that every monotone predicate lifting $\lambda : \mathcal{Q}(-)^A \to \mathcal{Q} \circ \mathsf{T}$ is equivalent to an atomic formula of $\mathlang{ML}_{\fun}^{1}(A)$.
In the following we shall not take care to distinguish between such a monotone
predicate lifting and the corresponding
atomic formula.
\end{remark}
\begin{defi}
Any translation $(\cdot)^{*}: \mathlang{SO}_{\Lambda}^{1} \to \mathlang{ML}_{\fun}^{1}$ induces a construction
on automata, transforming a second-order $\Lambda$-automaton
$\mathbb{A} = (A,\Delta,a_I,\Omega)$ into the modal automaton
$\mathbb{A}^{*} = (A,\Delta^{*},a_I,\Omega)$, with $\Delta^{*}$ given by
$\Delta^{*}(a,c) \mathrel{:=} (\Delta(a,c))^{*}$.
\end{defi}
The proof of the following result closely follows that of the main result
in \cite{vene:expr14}.
The main difference with \cite{vene:expr14} is that here we need an
``unravelling''-like component.
\begin{prop}
\label{p:unr}
Assume that $\mathlang{SO}_{\Lambda}^{1}$ admits a uniform translation $(\cdot)^{*}$, and let
$\mathbb{A}$ be a second-order $\Lambda$-automaton.
Then for each pointed $\mathsf{T}$-model $(\mathbb{S},s)$ there is a $\mathsf{T}$-tree model
$(\mathbb{T},R,t)$, with a $\mathsf{T}$-model homomorphism $f$ from $\mathbb{T}$ to
$\mathbb{S}$, mapping $t$ to $s$, and such that
\[
\mathbb{A} \text{ accepts } (\mathbb{T},R,t) \text{ iff }
\mathbb{A}^{*} \text{ accepts } (\mathbb{S},s).
\]
Furthermore, given that $\mathbb{S} = (S,\sigma,V)$, if the map $h_{\sigma(s)} : S_* \to S$ is surjective, so is $f$.
\end{prop}
\begin{proof}
Consider any given pointed $\mathsf{T}$-model $(\mathbb{S}_1,s_1)$ where $\mathbb{S}_1 = (S_1,\sigma_1,V_1)$. We are going to construct a $\mathsf{T}$-tree model $(\mathbb{S}_2,R,s_2)$, $\mathbb{S}_2 = (S_2,\sigma_2,V_2)$, together with a model homomorphism from the underlying pointed $\mathsf{T}$-model $\mathbb{S}_2$ to $\mathbb{S}_1$ mapping $s_2$ to $s_1$, and such that $\mathbb{A}$ accepts the $\mathsf{T}$-tree model $(\mathbb{S}_2,s_2)$ if and only if $\mathbb{A}^*$ accepts the pointed $\mathsf{T}$-model $(\mathbb{S}_1,s_1)$.
We construct this $\mathsf{T}$-tree model as follows: for each $u \in S_1$, we define an associated pair $(X_u,\alpha_u)$ as follows: set $X_u = (S_1)_*$ and set $\alpha_u = \sigma_1(u)_*$.
Observe that, by the construction of these one-step models, for each $u \in S_1$, there is a mapping
$$\xi_u : X_u \rightarrow S_1$$
such that
\begin{enumerate}
\item $\mathsf{T}(\xi_u)(\alpha_u) = \sigma_1(u)$
\item For each valuation $U : A \rightarrow \mathcal{Q} (S_1)$, every $u \in S_1$ and every one-step formula $\Delta(a,c)$ appearing in $\mathbb{A}$, we have
$$(S_1,\sigma_1(u),U)\vDash_1 \Delta(a,c) \text{ iff } (X_u,\alpha_u,U_{[\xi_u]})\vDash_1 \Delta^*(a,c)$$
\end{enumerate}
The map $\xi_u$ is given by $h_{\sigma_1(u)}$. We now construct the $\mathsf{T}$-tree model $(S_2,R,\sigma_2,s_2,V_2)$ as follows: first, consider the set of all non-empty finite (non-empty) tuples $(v_1,...,v_n)$ of elements in
$$\{s_1\}\cup\bigcup_{u \in S_1} X_u$$
such that $v_1 = s_1$. We define, by induction, for each natural number $n > 0$ a subset $M_n$ of this set, and a mapping $\gamma_n : M_n \rightarrow S_1$, as follows:
\begin{itemize}
\item set $M_1 = \{(s_1)\}$, and define $\gamma_1(s_1) = s_1$.
\item Set $M_{n + 1} = \{\vec{v} * w \mid \vec{v} \in M_n, w \in X_{\gamma_n(\vec{v})}\}$. Define $\gamma_{n + 1} (\vec{v} * w) = \xi_{\gamma_n(\vec{v})}(w)$.
\end{itemize}
Here, we write $\vec{v}*w$ to denote the tuple $(v_1,...,v_n,w)$ if $\vec{v} = (v_1,...,v_n)$.
Set $S_2 = \bigcup_{n > 0} M_n$, and define $\gamma = \bigcup_{n> 0} \gamma_n$. Define the relation $R \subseteq S_2 \times S_2$ to be
$$\{(\vec{v},\vec{v} * w) \mid \vec{v} \in S_2, w \in X_{\gamma(\vec{v}})\}$$
Note that there is, for every $\vec{v} \in S_2$, a bijection $i_{\vec{v}} : X_{\gamma(\vec{v})} \simeq R[\vec{v}]$ given by $w \mapsto \vec{v} * w$. Note also that, for each $\vec{v} \in S_2$, we have
$$ \gamma \circ i_{\vec{v}} = \xi_{\gamma(\vec{v})}$$
With this in mind, we define the coalgebra structure $\sigma_2$ by setting
$$\sigma_2(\vec{v}) = T(i_{\vec{v}})(\alpha_{\gamma(\vec{v})})$$
Finally, set $s_2$ to be the unique singleton tuple with sole element $s_1$, and define the valuation $V_2$ by setting $V_2^\dagger(\vec{v}) = V_1^\dagger(\gamma(\vec{v}))$.
Clearly, $(S_2,R,\sigma_2,s_2,V_2)$ is a $\mathsf{T}$-tree model. Denote the underlying $T$-model by $\mathbb{S}_2$.
We can then prove the following two claims:
\textbf{Claim 1}: The map $\gamma$ is a $T$-model homomorphism from $\mathbb{S}_2$ to $\mathbb{S}_1$.
\textbf{Claim 2}: $\mathbb{A}$ accepts $(S_2,R,\sigma_2,s_2,V_2)$ iff $\mathbb{A}^*$ accepts $\mathbb{S}_1$.
The proof of Claim 2 is left to the reader. We prove the first claim:
The map $\gamma$ clearly respects the truth values of all propositional atoms, and $\gamma(s_2) = s_1$. It suffices to show that $\gamma$ is a coalgebra morphism, i.e. that $\mathsf{T} \gamma (\sigma_2(\vec{v})) = \sigma_1 (\gamma(\vec{v}))$ for all $\vec{v}$. Pick any $\vec{v} \in S_2$. We have:
\begin{eqnarray}
\mathsf{T}\gamma(\sigma_2(\vec{v})) & = & \mathsf{T}\gamma \circ \mathsf{T}(i_{\vec{v}})(\alpha_{\gamma(\vec{v})}) \\
& = & \mathsf{T}(\gamma \circ i_{\vec{v}})(\alpha_{\gamma(\vec{v})}) \\
& = & \mathsf{T}(\xi_{\gamma(\vec{v})})(\alpha_{\gamma(\vec{v})}) \\
& = & \sigma_1(\gamma(\vec{v}))
\end{eqnarray}
as required. \end{proof}
From this, a routine argument yields the following result.
\begin{theo}[Characterization Theorem 1]
\label{generalcharacterization}
Let $\Lambda$ be an expressively complete set of monotone predicate liftings
for a set functor $\mathsf{T}$, and assume that $\mathlang{SO}_{\Lambda}^{}(A)$ (for any set of
variables $A$) admits uniform translations.
Then $\muML_{\Lambda}$ is the bisimulation-invariant fragment of $\MSO_{\Lambda}$.
\end{theo}
The existence of uniform translations for the one-step language
$\cite{vene:expr14}$ involves two components: a translation
on the syntactic side and a uniform construction on the semantic side.
However, as we shall now see, we can focus entirely on finding a suitable
uniform construction for the one-step models; the syntactic translation will
come for free.
\begin{defi}
Let $\varphi$ be any formula of $\mathlang{SO}_{\Lambda}^{1}(A)$ of quantifier depth $\leq k$,
and let $F$ be a uniform construction for $k$.
Then, we define the generalized predicate lifting $\varphi^* :
\mathcal{Q}(-)^A \rightarrow \mathcal{Q} \circ \mathsf{T}$ by setting, for a given set $X$,
$\alpha \in \mathsf{T} X$ and $V : A \rightarrow \mathcal{Q}(X)$:
$$
\alpha \in\varphi^*_X(V) \text{ iff }
(X_*,\alpha_*,V_{[h_{\alpha}]})\vDash_1 \varphi.
$$
\end{defi}
The following is obvious:
\begin{prop}
If $\varphi$ is a monotone formula then $\varphi^*$ is a monotone generalized predicate
lifting.
\end{prop}
Note that, in order for $\mathlang{SO}_{\Lambda}^{1}(A)$ to admit a uniform translation, it
suffices that there exists for any $k$ a uniform construction $F$ such that,
for every formula $\varphi$ of quantfier depth $\leq k$, the generalized
lifting $\varphi^*$ is natural. An equivalent formulation of this condition
is the following.
\begin{prop}
Let $\varphi$ be any one-step formula in $\mathlang{SO}_{\Lambda}^{1}(A)$ and let $F$ be a uniform construction.
Then the lifting $\varphi^*$ is natural if, for any pair of sets $X,Y$,
any map $f : X \rightarrow Y$ and any valuation $V : A \rightarrow \mathcal{Q}(Y)$, we
have
$$(\star) \quad (X_*,\alpha_*,V_{[f \circ h_\alpha]})\vDash_1 \varphi
\text{ iff } (Y_*, \beta_*, V_{[h_\beta]}) \vDash_1 \varphi
$$
provided that $\mathsf{T} f(\alpha) = \beta$.
\end{prop}
A uniform construction $F$ is said to be \textit{adequate} for $k$, and with
respect to the liftings $\Lambda$, if the equivalence $(\star)$ holds for all
(monotone) formulas in $\mathlang{SO}_{\Lambda}^{1}(A)$ of quantifier depth
$\leq k$ (for any finite set of variables $A$).
Since we could of course take the quantifier depth $k$ and the set of liftings
as extra inputs for the uniform construction, we shall simply say that the
functor $\mathsf{T}$ admits an adequate uniform construction if there is an adequate
uniform construction for $\mathsf{T}$ with respect to every $k$ and every set of
monotone liftings.
If $\Lambda$ is an expressively complete set of liftings, this is equivalent to
requiring an adequate uniform construction with respect to $\Lambda$, for every
$k$.
The following theorem, from which we obtain Theorem~\ref{t:main2} by taking for
$\Lambda$ the set of all monotone liftings for $\mathsf{T}$, summarizes
the results of this section.
\begin{theo}
\label{t:bisinv2}
Let $\Lambda$ be any expressively complete set of monotone predicate
liftings for the set functor $\mathsf{T}$.
If $\mathsf{T}$ admits an adequate uniform construction, then
\[
\muML_{\Lambda} \equiv \mathlang{MSO}_{\Lambda}/{\simeq}.
\]
\end{theo}
\begin{example}
\label{ex:psf}
As a first application, the standard Janin-Walukiewicz characterization of the
modal $\mu$-calculus can be seen as an instance of the result by taking
$\Lambda = \{ \Box, \Diamond \}$ and $\mathsf{T} = \mathcal{P}$, recalling that $\mathlang{MSO} = \mathlang{MSO}_{\{\Diamond\}} \equiv \mathlang{MSO}_{\{\Box,\Diamond\}}$.
The adequate uniform construction for $\mathcal{P}$ is given as follows: consider a
pair $(X,\alpha)$ with $\alpha \in \mathcal{P}(X)$.
We take this to $X_* = \alpha_* = \alpha \times \omega$, and we let $h_\alpha:
\alpha \times \omega \to \alpha$ be the projection map.
\end{example}
It turns out that several other applications of this result can be obtained in
a particularly simple way.
Say that a uniform construction $F$ is \textit{strongly adequate} if, for any
mapping $f : X \to Y$ and any $\alpha \in \mathsf{T} X$, $\beta \in \mathsf{T} Y$ with
$\mathsf{T} f (\alpha) = \beta$, there is a bijection $g : X_{*}\to Y_{*}$ such that
$\mathsf{T} g (\alpha_{*}) = \beta_{*}$ and $f \circ h_\alpha = h_\beta \circ g$.
Since it is easy to check that any strongly adequate uniform construction is
adequate, we get:
\begin{coro}
If there is a strongly adequate uniform construction for $\mathsf{T}$, then
$\mu \mathtt{ML}_\mathsf{T}\equiv\mathtt{MSO}_\mathsf{T} {/} {\simeq} $.
\end{coro}
\begin{example}
\label{ex:bag}
As a first example, consider the finitary multiset (``bags'') functor
$\mathcal{B}$, which sends a set $X$ to the set of mappings $f : X \rightarrow
\omega$ such that the set $\{u \in X \mid f(u) = 0\}$ is cofinite.
The action on morphisms is given by letting, for $f \in \mathcal{B}X$ and
$h : X \rightarrow Y$, the multiset $\mathcal{B}h(f) : Y \rightarrow \omega$
be defined by $w \mapsto \sum_{h(v) = w} f(v)$.
Given a pair $X,\alpha$ where $\alpha : X \rightarrow \omega$ has finite support,
we define
$$X_{*} = \bigcup \{\{u\} \times \alpha(u) \mid u \in X\}.
$$
Here, we identify each each $n \in \omega$ with the set $\{0,...,n-1\}$.
The mapping $\alpha_{*} : X_{*} \rightarrow \omega$ is defined by setting
$\alpha_{*}(w) = 1$ for all $w \in X_{*}$.
The map $h_\alpha : X_* \rightarrow X$ is defined by $(u,i) \mapsto u$.
It is easy to check that the construction $F$ is strongly adequate, hence
$\mu \mathtt{ML}_{\mathcal{B}}\equiv\mathtt{MSO}_{\mathcal{B}} {/} {\simeq}$.
\end{example}
As a final application, consider the set of all \textit{exponential polynomial
functors}~\cite{jaco:intr12} defined by the ``grammar''
$$\mathsf{T} \mathrel{::=} \mathsf{C}
\mathrel{|} \mathsf{Id} \mathrel{|} \mathsf{T} \times \mathsf{T}
\mathrel{|} \coprod_{i \in I} \mathsf{T}_i \mathrel{|} \mathsf{T}(-)^\mathsf{C}
$$
where $\mathsf{C}$ is any constant functor for some set $C$, and $\mathsf{Id}$
is the identity functor on $\mathbf{Set}$.
These functors cover many important applications: streams, binary trees,
deterministic finite automata and deterministic labelled transition systems
are all examples of coalgebras for exponential polynomial functors, as is
the socalled \textit{game functor} whose coalgebras provide the semantics for
``Coalition Logic'' \cite{cirs:moda11}.
\begin{prop}
\label{p:epf}
Every exponential polynomial functor admits a strongly adequate uniform
translation.
\end{prop}
\begin{coro}
\label{c:epf}
For every exponential polynomial functor $\mathsf{T}$, we have
$\mu \mathtt{ML}_{\mathsf{T}}\equiv\mathtt{MSO}_{\mathsf{T}} {/} {\simeq}$.
\end{coro}
The cases where we can find a strongly adequate uniform construction are the
most straightforward applications of Theorem \ref{t:bisinv2} that we know of.
The Janin-Walukiewicz theorem is a less direct application: there is no strongly
adequate uniform construction for the powerset functor, but there is an adequate
uniform construction.
In the next section, we shall study an example of a functor where there is no
adequate uniform construction at all.
\section{The monotone neighborhood functor}
\label{sec:mon}
The final section of our paper concerns the monotone neighborhood functor
$\mathcal{M}$.
Our main result concerns a characterization of the fragment of $\mathlang{MSO}_{\mathcal{M}}$
that is invariant under \textit{global} neighborhood bisimulations, to be
introduced below.
Our proof applies the method of section~\ref{sec:bisinv}, but not directly:
we will first see that the functor $\mathcal{M}$ itself does \emph{not} admit an
adequate uniform construction.
\subsection{No adequate uniform construction for $\mathcal{M}$}
We first consider the negative result.
\begin{prop}
\label{p:no-adc-mon}
There is no adequate uniform construction for the monotone neighborhood
functor $\mathcal{M}$.
\end{prop}
\begin{proof}
To arrive at a contradiction assume that $F$ is adequate.
Fix some $a \in A$ and consider the formula $\varphi = \forall Z. (a \subseteq
Z)$ expressing that $a$ has empty extension.
Let $Y$ be the set $\{u,v\}$ and let $\beta \in \mathcal{M}Y$ be the
neighborhood structure $\{\{u\},\{u,v\}\}$.
Let $V$ be any valuation with $V(a) = \{v\}$.
First, we prove that
$(Y,\beta,V) \vDash_1 \varphi^*$:
to see this, consider the one-step model $(Y',\beta',V')$ where $\beta' =
\{\{u\}\}$ and we recall that $Y' = \{u\}$, and where $V'$ is simply the
restriction of $V$ to $Y'$. It is easy to show that
$( Y'_*, \beta'_*,V'_{[h_{\beta'}]}) \vDash_1 \varphi$, and hence
$(Y',\beta',V') \vDash_1 \varphi^*$.
Since the generalized predicate lifting $\varphi^*$ is natural by
assumption and $\mathcal{M} \iota_{Y',Y}(\beta')
= \beta$, we get $(Y,\beta,V) \vDash_1 \varphi^*$ as required.
With this in mind, let $X$ be the set $\{u^*,v^*,w^*\}$ and let $\alpha \in
\mathcal{M} X$ be the neighborhood structure
$$\{\{u^*,v^*\}, \{u^*,w^*\}, \{u^*,v^*,w^*\}\}$$
Define the map $f : X \rightarrow Y$ by setting $u^* \mapsto u$, $v^* \mapsto v$
and $w^* \mapsto u$.
It can easily be checked that $\mathcal{M} f (\alpha) = \beta$.
By naturality of the formula $\varphi^*$, it follows that
$(X,\alpha,V_{[f]}) \vDash_1 \varphi^*$.
Hence we must have
$$( X_*, \alpha_* ,V_{[f \circ h_{\alpha}]}) \vDash_1 \varphi$$
hence $V_{[f \circ h_\alpha]}(a) = \emptyset$. Since $v^* \in V_{[f]}(a)$, this
means that we have $v^* \notin h_\alpha[X_*]$.
But since $\mathcal{M} h_{\alpha}(\alpha_*) = \alpha$, this means $h_\alpha[X_*]$
must be a support for $\alpha$. But it is easy to show that $\alpha$ cannot
have a support $S$ with $v^* \notin S$,
so we have now reached a contradiction showing that $F$ cannot be an
adequate construction.
\end{proof}
\subsection{The functor $\mathcal{M}^{\star}$}
In this section, as a step towards our main characterization result, we shall
consider the language $\mathlang{\mu ML}_{\mathcal{M}^{\star}}$, where the functor
$\mathcal{M}^{\star}$ is a slight variation of the monotone neighborhood functor
$\mathcal{M}$.
The functor $\mathcal{M}^{\star}$ is obtained as the subfunctor of $\mathcal{M}
\times \mathcal{P}$ given by
$$X \mapsto \{(\alpha,Y) \in \mathcal{M} X \times \mathcal{P} X \mid Y
\text{ supports } \alpha\}$$
This is indeed a subfunctor of $\mathcal{M} \times \mathcal{P}$, because given a map
$h : X \rightarrow Y$, if $Z$ is a support for $\alpha \in \mathcal{M} X$, then
$h[Z]$ is a support for $\mathcal{M} h(\alpha)$. Given $\alpha \in
\mathcal{M}^{\star} X$, we will write $\alpha = (N_\alpha,S_\alpha)$.
\begin{defi}
For the functor $\mathcal{M}^{\star}$ we define the unary predicate liftings $\Box$
and $E$ by
\begin{align*}
\Box_X(Z) & \mathrel{:=} \{\alpha \in \mathcal{M}^{\star} X \mid Z \in N_\alpha\}
\\
E_X (Z) & \mathrel{:=}
\{\alpha \in \mathcal{M}^{\star} X \mid Z \cap S_\alpha \neq \emptyset\},
\end{align*}
and we let $\Diamond$ be the dual of $\Box$ and let $E^d$ be the dual of $E$.
The set of liftings $\{\Box,\Diamond, E,E^d\}$ is denoted as $\Theta$.
\end{defi}
The set $\Theta$ is an expressively complete set of liftings for $\mathcal{M}^{\star}$. We shall omit the proof of this fact here, and merely state it as the following proposition:
\begin{prop}
\label{p:lyndon}
Every monotone natural predicate lifting
$\lambda : \mathcal{Q}(-)^A \rightarrow \mathcal{Q} \circ \mathcal{M}^{\star}$ is equivalent to a
formula in $\mathlang{ML}_{\Theta}(A)$.
\end{prop}
The main technical result of this section states the existence, for all $k$,
of a uniform construction $F$ that is adequate for $k$ and with respect to
the set of liftings $\{\Box,E\}$.
\begin{defi}
\label{d:F-monstar}
Fix a natural number $k$.
Given a set $X$, and object $\alpha \in \mathcal{M}^{\star} X$, put
$$
X_* \mathrel{:=}
\{(u,i,Z,j) \in (X \times 2^{k} \times \mathcal{P}(S_\alpha) \times \omega)
\mid u \in Z \},
$$
and let $\pi_X$ be the projection map from $X_*$ to $X$.
Define $\alpha_* = (N_{\alpha_*},S_{\alpha_*}) \in \mathcal{M}^{\star}(X_*)$
by setting $S_{\alpha_*} = X_*$, and set $Z \in N_{\alpha_*}$ for $Z
\subseteq S_{\alpha_*}$ iff $\left\lceil Y,j\right\rceil \subseteq Z$ for some
$Y \in \alpha$, $Y \subseteq S_\alpha$ and some $j < \omega$, where
$$\left\lceil Y,j \right\rceil := \{(u,i,Y,j) \mid u \in Y, i < 2^k
\}.
$$
The sets of the form $\left\lceil Z,j\right\rceil$ will be called the
\textit{basic members} of $N_{\alpha^*}$.
\end{defi}
The main goal of this section is to prove the following:
\begin{prop}
\label{p:ad-monstar}
The construction given in
Definition~\ref{d:F-monstar} is an adequate, uniform construction for $k$.
\end{prop}
It is easy to check that, for all sets $X$ and $\alpha \in \mathcal{M}^{\star} X$, we have $\mathcal{M}^{\star}\pi_X(\alpha_F) = \alpha$.
Our main goal in this section is to prove the following result, from which Proposition~\ref{p:ad-monstar} now follows:
\begin{lemma}
\label{monuniform}
Let $X,Y$ be any sets, $\alpha \in \mathcal{M}^{\star} X$, $\beta \in \mathcal{M}^{\star} Y$
and $V : A \rightarrow Q(Y)$. Suppose that we have a map $h : X \rightarrow Y$ such that $\mathcal{M}^{\star} h(\alpha) = \beta$. Then we have
$$(X_*,\alpha_*,V_{[h \circ \pi_X]}) \equiv^k (Y_*,\beta_*,V_{ [\pi_Y]})$$
\end{lemma}
Here, and throughout this section, we write $(X,\alpha,V) \equiv^k (Y,\beta,U)$ to say that two one-step models satisfy the same formulas of $\mathlang{MSO}_{\{\Box,E\}}^{1}(A)$ with at most $n$ nested quantifiers. Let us keep the data $X,Y,\alpha,\beta,V$ and $h$ fixed throughout the proof, and assume that $\mathcal{M}^{\star} h(\alpha) = \beta$. We will also assume, from now on, that $N_\alpha$ and $N_\beta$ are both non-empty sets: if one of them is empty then both of them are, and in this case the lemma can be proved essentially using an easier version of the argument we use below.
\begin{defi}
Given a finite set of variables $A$, a propositional $A$-\textit{type} $\tau$ is a subset of $A$. Given a set $X$ and a valuation $V : A \rightarrow Q(X)$, the propositional $A$-type of $v \in X$ is defined to be $V^\dagger(v) = \{a \in A \mid v \in V(a)\}$.
\end{defi}
\begin{defi}
Given a basic member $\left\lceil Z,j \right\rceil$ either in $N_{\alpha_*}$ or in $N_{\beta_*}$, a valuation
$V : A \rightarrow \mathcal{Q}(X_*)$ or $V : A \rightarrow \mathcal{Q}(Y_*)$, and a natural number $m$, the \textit{$m$-signature} of $\left\lceil Z,j \right\rceil $ over variables $A$ and relative to the valuation $V$ is the mapping $\sigma : \mathcal{P}(A) \rightarrow \{0,...,m\} $ defined by setting $\sigma(t)$ to be $n < m$ if $\left\lceil Z,j \right\rceil$ contains exactly $n$ elements of type $t$ under the valuation $V$, or $\sigma(t) = m$ if $\left\lceil Z,j \right\rceil$ contains \textit{at least} $m$ elements of type $t$.
\end{defi}
\begin{defi}
Let $B$ be any set of variables containing $A$, and let $V_1 : B \rightarrow \mathcal{Q}(X_*)$ and $V_2 : B \rightarrow \mathcal{Q}(Y_*)$. Then for any natural number $n$ we write
$$(X_*,\alpha_*,V_1) \approx^{n} (Y_*,\beta_*,V_2) $$
and say that these one-step models \textit{match up to depth $n$}, if: for every $n$-signature $\sigma$ over variables $B$, either the number of basic elements of signature $\sigma$ in $N_{\alpha_*}$ and $N_{\beta_*}$ respectively are both finite and the same, or both infinite.
\end{defi}
\begin{lemma}
\label{kstep}
$(X_*,\alpha_*,V_{[h \circ \pi_X]}) \approx^{2^k} (Y_*,\beta_*,V_{ [\pi_Y]})$.
\end{lemma}
\begin{proof}
First note that, for any $2^k$-signature $\sigma$, $N_{}$ either contains no basic elements of signature $\sigma$, or infinitely many: if there is some basic element $\left\lceil Z,j \right\rceil$ if signature $\sigma$, then for any $i \neq j$, the basic element $\left\lceil Z,i \right\rceil$ has the same $2^k$-signature as $\left\lceil Z,j \right\rceil$ with respect to the valuation $V_{h \circ \pi_X}$. The same holds for $N_{\beta_*}$ with respect to the valuation $V_{ \pi_Y}$.
Thus, it suffices to show that $N_{\alpha_*}$ contains a basic element of signature $\sigma$ w.r.t. $V_{[h \circ \pi_X]}$ iff $N_{\beta_*}$ contains a basic element of signature $\sigma$ w.r.t $V_{[\pi_Y]}$. Suppose that $N_{\beta_*}$ contains a basic element $\left\lceil Z,j\right\rceil$ of signature $\sigma$, where $Z \in N_\beta$ and $Z \subseteq S_\beta$. Since $N_\beta = \mathcal{M}h(N_\alpha)$, we have $h^{-1}[Z] \in N_\alpha$ and so $h^{-1}[Z] \cap S_\alpha \in N_\alpha$. Hence we get
$$\left\lceil h^{-1}[Z] \cap S_\alpha, 0\right\rceil \in N_{\alpha_*}$$
It is easy to see that any basic element of $N_{\alpha_*}$ contains either $0$ or at least $2^k$ members of a propositional type $t$ w.r.t. $V_{[h \circ \pi_X]}$, and the same is true of any basic element of $N_{\beta_*}$ w.r.t. $V_{[\pi_Y]}$. Hence, to show that $\left\lceil h^{-1}[Z] \cap S_\alpha, 0\right\rceil $ has the same $2^k$-signature as $\left\lceil Z,j\right\rceil$, it suffices to show that these basic elements realize the same propositional types over $B$. So suppose $\left\lceil h^{-1}[Z] \cap S_\alpha, 0\right\rceil $ contains some element $v$ of type $t$. Then $\pi_X(v) \in h^{-1}[Z] \cap S_\alpha$, so $h \circ \pi_X(v) \in Z$. This means that
$$(h \circ \pi_X(v),0,Z,j) \in \left\lceil Z,j\right\rceil $$
and this will have the same propositional type as $v$. Conversely, suppose that $\left\lceil Z,j\right\rceil $ contains some element $w$ of type $t$. Then $\pi_Y(w) \in Z$. Since $Z \subseteq S_\beta$, and $S_\beta = h[S_\alpha]$, there exists some $w' \in h^{-1}[Z] \cap S_\alpha$ with $h(w') = \pi_Y(w)$. We get
$$(w',0,h^{-1}[Z] \cap S_\alpha,0) \in \left\lceil h^{-1}[Z] \cap S_\alpha, 0\right\rceil $$
and this will have the type $t$.
Now, suppose that $N_{\alpha_*}$ contains a basic element $\left\lceil Z,j\right\rceil$ of $2^k$-signature $\sigma$, so that $Z \in N_\alpha$ and $Z \subseteq S_\alpha$. Then $h[Z] \in N_\beta$, and furthermore $h[Z] \subseteq S_\beta = h[S_\alpha]$. Again, it suffices to check that $\left\lceil Z,j\right\rceil$ and $\left\lceil h[Z],0\right\rceil$ realize the same propositional types. Given $v\in \left\lceil Z,j\right\rceil$, we have $h \circ\pi_X(v) \in h[Z]$ and so
$$ (h \circ \pi_X(v),0,h[Z],0) \in \left\lceil h[Z],0\right\rceil$$
and this will have the same propositional type as $v$. Conversely, if $w \in \left\lceil h[Z],0\right\rceil$ then $\pi_Y(w) \in h[Z]$, so there is $w' \in Z$ with $h(w') = \pi_Y(w)$. We get
$$(w',0,Z,j) \in \left\lceil Z,j\right\rceil$$
and this has the same propositional type as $w$.
\end{proof}
We are going to show, by induction on a natural number $m \leq k$, that if two one-step models of the form $(X_*,\alpha_*,V_1)$ and $(Y_*,\beta_*,V_2)$ match up to depth $2^m$, then they satisfy the same formulas of quantifier depth $m$. For the basis case of $2^0 = 1$, we need the following result:
\begin{lemma}
\label{atomicstep}
Let $B$ be a set of variables containing $A$, and let $V_1 : B \rightarrow \mathcal{Q}(X_*)$ and $V_2 : B \rightarrow \mathcal{Q}(F^{\beta}(Y))$ be valuations such that
$$(X_*,\alpha_*,V_1) \approx^{1} (Y_*,\beta_*,V_2) $$
Then these two one-step models satisfy the same atomic formulas of the one-step
language $\mathlang{MSO}^{1}_{\{\Box,E\}}$.
\end{lemma}
\begin{proof}
Suppose first that
$$X_*,\alpha_*,V_1)\vDash_1 p \subseteq q$$
where $p,q \in B$. Suppose that $V_2(p) \nsubseteq V_2(q)$. Then there is some $(u,i,Z,j) \in Y_*$ such that $(u,i,Z,j) \in V_2(p)\setminus V_2(q)$. We have $(u,i,Z,j) \in \left\lceil Z,j\right\rceil$, and so there must be some basic element $\left\lceil Z',j'\right\rceil$ of $N_{\alpha_*}$ of the same $1$-signature over variables $B$ as $\left\lceil Z,j\right\rceil$. It follows that there is some element of $\left\lceil Z',j'\right\rceil$ of the same propositional type as $(u,i,Z,j)$, and then we cannot have $V_1(p) \subseteq V_1(q)$. The converse direction is proved in the same manner.
Now, suppose that
$$(X_*,\alpha_*,V_1)\vDash_1 \Box p$$
Then $V_1(p) \in N_{\alpha_*}$, so there is some basic element $\left\lceil Z,j\right\rceil \in N_{\alpha_*}$ with $\left\lceil Z,j\right\rceil \subseteq V_1(p)$. There must be some basic $\left\lceil Z',j'\right\rceil \in N_{\beta_*}$ of the same $1$-signature over $B$ as $\left\lceil Z,j\right\rceil$, and clearly it follows that $\left\lceil Z',j'\right\rceil \subseteq V_2(p)$ and so $V_2(p) \in N_{\beta_*}$ as required. The converse direction is proved in the same way.
Finally, suppose that
$$(X_*,\alpha_*,V_1)\vDash_1 E p$$
Then there is some $(u,i,Z,j) \in S_{\alpha_*}$ with $(u,i,Z,j) \in V_1(p)$. We have $(u,i,Z,j) \in \left\lceil Z,j\right\rceil$, and there must be some basic element $\left\lceil Z',j'\right\rceil$ of $N_{\beta_*}$ with the same $1$-signature as $\left\lceil Z,j\right\rceil$. Hence, $\left\lceil Z',j'\right\rceil$ contains some element $(u',i',Z',j')$ of the same propositional type as $(u,i,Z,j)$, and it follows that
$$(Y_*,\beta_*,V_2)\vDash_1 E p$$
as required. The converse direction is proved in the same way.
\end{proof}
To clinch the proof of Proposition \ref{p:ad-monstar}, we now only need the following lemma:
\begin{lemma}
\label{mainlemmamonstar}
Let $B$ be a finite set of variables containing $A$, let $0 < m \leq k$ and let $V_1 : B \rightarrow Q(X_*)$ and $V_2 : B \rightarrow \mathcal{Q} (Y_*)$ be valuations such that
$$(X_*,\alpha_*,V_1) \approx^{2^m} (Y_*,\beta_*,V_2)$$
Let $q$ be any fresh variable. Then for any valuation
$V_1'$
extending $V_1$ with some value for $q$, there exists a valuation
$V_2' $
extending $V_2$, such that
$$(X_*,\alpha_*,V_1') \approx^{2^{(m-1)}} (Y_*,\beta_*,V_2')$$
and vice versa.
\end{lemma}
\begin{proof}
We only prove one direction since the other direction can be proved by a symmetric argument. Let $V_1'$ be given. By the hypothesis, for any $2^m$-signature $\sigma$ over the variables $B$, either the number of basic elements of signature $\sigma$ in $N_{\alpha_*}$ and $N_{\beta_*}$ relative to $V_1$ and $V_2$ are both finite and the same, or both infinite. Let $\sigma_1,...,\sigma_k$ be a list of all the distinct $2^m$-signatures over $B$ such that the set of basic elements of $N_{\alpha_*}$ and $N_{\beta_*}$ of signature $\sigma_i$, with $1 \leq i \leq k$, is non-empty but finite, and let $\sigma_{k+1},...,\sigma_{l}$ be a list of all the $2^m$-signatures such that, for $k + 1 \leq i \leq l$, there are infinitely many basic elements of $N_{\alpha_*}$ and of $N_{\beta_*}$ of signature $\sigma_i$. Then, for each $i \in \{1,...,l\}$, let $\alpha_*[\sigma_i]$ denote the set of basic elements in $N_{\alpha_*}$ of signature $\sigma_i$, and similarly let $\beta_*[\sigma_i]$ denote the set of basic elements of $N_{\beta_*}$ of signature $\sigma_i$. Then $\alpha_*[\sigma_1],...,\alpha_*[\sigma_l]$ is a partition of the set of basic elements of $N_{\alpha_*}$ into non-empty cells, and similarly $\beta_*[\sigma_1],...,\beta_*[\sigma_l]$ is a partition of the set of basic elements of $N_{\beta_*}$.
Given the extended valuation $V_1'$ in $X_*$ defined on variables $B \cup \{q\}$, we similarly let $\tau_1,...,\tau_{k^*}$ be a list of all the $2^{m - 1}$-signatures over $B \cup \{q\}$ such that, for $1 \leq i \leq k^*$, the set of basic elements of $N_{\alpha_*}$ of $2^{m - 1}$-signature $\tau_i$ is non-empty but finite. We let $\tau_{k^* + 1},...,\tau_{l^*}$ be a list of all the $2^{m - 1}$-signatures over $B \cup \{q\}$ such that, for each $i$ with $k^* + 1 \leq i \leq l^*$, the set of basic elements of $N_{\alpha_*}$ of $2^{m - 1}$-signature $\tau_i$ is infinite. Let $\alpha_*[\tau_i]$ denote the set of basic elements of $N_{\alpha_*}$ of $2^{m - 1}$-signature $\tau_i$, so that the collection $\alpha_*[\tau_1],...,\alpha_*[\tau_{l^*}]$ constitutes a second partition of the set of basic elements of $N_{\alpha*}$. It will be useful to introduce the abbreviation $D_1$ for the finite set $\alpha_*[\sigma_1] \cup...\cup \alpha_*[\sigma_k]$, and the abbreviation $D_2$ for the finite set $\alpha_*[\tau_1] \cup...\cup \alpha_*[\tau_{k^*}]$.
For each $i$ with $1 \leq i \leq k$, there is a bijection between the set $\alpha_*[\sigma_i]$ and $\beta_*[\sigma_i]$, and we can paste all these bijections together into a bijective map
$$f : \alpha_*[\sigma_1] \cup...\cup \alpha_*[\sigma_k] \rightarrow \beta_*[\sigma_1] \cup ... \cup \beta_*[\sigma_k]$$
Since every basic element of $N_{\alpha_*}$ not in $D_1$ belongs to a $2^m$-signature of which there are infinitely many basic elements in $\beta_*$, and since $D_1 \cup D_2$ is finite, it is easy to see that we can extend the map $f$ to a map $g$ which is an injection from the set $D_1 \cup D_2$
into the set of basic elements of $N_{\beta_*}$, such that for each basic element $\left\lceil Z,j\right\rceil$ in $D_1 \cup D_2$, $\left\lceil Z,j\right\rceil$ and $g(\left\lceil Z,j\right\rceil)$ have the same $2^m$-signature over $B$, and such that $g\!\upharpoonright\!D_1 = f$. Each basic element of $N_{\beta_*}$ not in the image of $g$ must then be of one of the $2^m$-signatures $\sigma_{k + 1},...,\sigma_{l}$, and so we can partition the set of basic elements of $N_{\beta_*}$ outside the image of $g$ into the cells
$\beta_*[\sigma_{k + 1}]\setminus g[D_2],...,\beta_*[\sigma_l]\setminus g[D_2]$.
For each $i$ with $k + 1 \leq i \leq l$, let $\gamma^i_{1},...,\gamma^i_{r}$ list all infinite sets of the form
$\alpha_*[\sigma_i] \cap \alpha_*[\tau_j]$ for $k^* + 1 \leq j \leq l^*$.
The list $\gamma^i_{1},...,\gamma^i_{r}$ must be non-empty, and so since the set $\beta_*[\sigma_{i}]\setminus g[D_2]$ is also infinite, we may
partition it into $r$ many infinite cells and list these as $\delta^i_{1},...,\delta^i_{r}$.
Now, for each basic element $\left\lceil Z,j\right\rceil$ of $\beta_*$, we
define a map
$W_{\left\lceil Z,j\right\rceil}$ from $B \cup \{q\}$ to
$\mathcal{P}(\left\lceil Z,j\right\rceil)$
by a case distinction as follows:
\emph{Case 1:} $\left\lceil Z,j\right\rceil = g(\left\lceil Z',j'\right\rceil)$
for some $\left\lceil Z',j'\right\rceil \in D_1 \cup D_2$.
Then $\left\lceil Z,j\right\rceil$ and $\left\lceil Z',j'\right\rceil$ have the
same $2^m$-signature over $B$. Using this fact we define the valuation
$W_{\left\lceil Z,j\right\rceil}$ so that, for each $p \in B$, we have
$W_{\left\lceil Z,j\right\rceil}(p) = V_2(p) \cap \left\lceil Z,j\right\rceil$,
and so that $\left\lceil Z',j'\right\rceil$ and $\left\lceil Z,j\right\rceil$
have the same $2^{m - 1}$-signature over $B \cup \{q\}$ with respect to the
valuations $V_1'$ and $W_{\left\lceil Z,j\right\rceil}$.
We show how to assign the value of the variable $q$: for each propositional type $t$ over $B \cup \{q\}$, there are three different possible cases to consider. If $\left\lceil Z',j'\right\rceil$ has $l < 2^{m-1}$ elements of type $t \cup \{q\}$ over $B \cup \{q\}$, then pick $l$ many elements of $\left\lceil Z,j\right\rceil$ of type $t$ and mark them by $q$. This is possible since $l < 2^{m-1} \leq 2^m$ and $\left\lceil Z',j'\right\rceil$ and $\left\lceil Z,j\right\rceil$ have the same $2^m$-signature. If there are $l < 2^{m-1}$ elements of $\left\lceil Z',j'\right\rceil$ of type $t$ over $B \cup \{q\}$, then pick $l $ elements of $\left\lceil Z,j\right\rceil$ of type $t$ over $B$, and mark all the other elements of $\left\lceil Z,j\right\rceil$ of type $t$ by $q$. Finally, if there are at least $2^{m-1}$ elements of $\left\lceil Z',j'\right\rceil$ of type $t \cup \{q\}$ over $B \cup \{q\}$ and at least $2^{m - 1}$ elements of $\left\lceil Z',j'\right\rceil$ of type $t$ over $B \cup \{q\}$, then all in all there must be at least $2^m$ elements of $\left\lceil Z',j'\right\rceil$ of type $t$ over $B$, and so there must be at least $2^m$ elements of $\left\lceil Z,j\right\rceil$ of type $t$ over $B$. Pick $2^{m-1}$ of these and mark them by $q$. Finally, let $W_{\left\lceil Z,j\right\rceil}(q)$ be the set of elements of $\left\lceil Z,j\right\rceil$ marked by $q$.
\emph{Case 2:} $\left\lceil Z,j\right\rceil$ is not in the image of $g$. Then there must be some $i \in \{k^* + 1,...,l^*\}$ such that $\left\lceil Z,j\right\rceil \in \beta_*[\sigma_{k + 1}]\setminus g[D_2]$, and this set is partitioned into $\delta^i_1,...,\delta^i_r$. Let $\left\lceil Z,j\right\rceil \in \delta^i_j$, and pick some arbitary element $\left\lceil Z',j'\right\rceil$ of the set $\gamma^i_j$. Then $\left\lceil Z',j'\right\rceil$ and $\left\lceil Z,j\right\rceil$ have the same $2^m$-signature over $B$ and we can proceed as in Case 1.
We define the valuation $V_2'$ by setting $V_2'(q)$ to be the union of the sets $W_{\left\lceil Z,j\right\rceil}(q)$ for $\left\lceil Z,j\right\rceil$ a basic element in $N_{\beta_*}$. It is now fairly straightforward to check that
$$(X_*,\alpha_*,V_1') \approx^{2^{(m-1)}} (Y_*,\beta_*,V_2')$$
as required.
First, suppose there are infinitely many basic elements of $N_{\alpha_*}$ of some $2^{m-1}$ signature $\tau_j$, meaning that $k^* \leq j \leq l^*$. Then since the set $\alpha_*[\tau_j]$ is infinite, $D_1$ is finite and $\alpha_*[\tau_j]$ is equal to
$$ (D_1 \cap \alpha_*[\tau_j]) \cup (\alpha_*[\sigma_{k +1}] \cap \alpha_*[\tau_{j}]) \cup ... \cup (\alpha_*[\sigma_l] \cap \alpha_*[\tau_{j}])$$
there must be some $i \in \{k+1,...,l\}$ such that the set $\alpha_*[\sigma_i] \cap \alpha_*[\tau_j]$ is infinite. This means that $\alpha_*[\sigma_i] \cap \alpha_*[\tau_j]$ appears in the list $\gamma^i_1,...,\gamma^i_r$, and so we see that all elements of some member of the list $\delta^i_1,...,\delta^i_r$ will have the $2^{m-1}$-signature $\tau_j$. Since each member of this list is infinite, we see that there must be infinitely many basic elements of $N_{\beta_*}$ of signature $\tau_j$.
Conversely, suppose there are infinitely many basic elements of $N_{\beta_*}$ of $2^{m-1}$-signature $\tau_j$ over $B \cup \{q\}$. Then since the image of $g$ is finite, some of these elements must be outside the image of $g$, which means that for some $i \in \{k + 1,...,l\}$, some member of the list $\delta^i_1,...,\delta^i_r$ will consist of elements of signature $\tau_j$. This means that some member of the list $\gamma_1^i,...,\gamma_r^i$ will consist of elements of signature $\tau_j$, and since each member of this list is infinite we see that $N_{\alpha_*}$ has infinitely many basic elements of $2^{m-1}$-signature $\tau_j$ over $B \cup \{q\}$.
Finally, suppose that there are finitely many basic elements of $N_{\alpha_*}$ and $N_{\beta_*}$ of $2^{m-1}$-signature $\tau_j$. We check that the mapping $g$ restricts to a bijection between the basic elements of $N_{\alpha_*}$ and $N_{\beta_*}$ of this signature. First, $g$ is injective and maps basic elements of $N_{\alpha_*}$ of signature $\tau_j$ to basic elements of $N_{\beta_*}$ of signature $\tau_j$. It only remains to show that (the restriction of) $g$ is surjective, i.e. each basic element $\left\lceil Z,r\right\rceil$ of signature $\tau_j$ is equal to $g(\left\lceil Z',r'\right\rceil)$ for some $\left\lceil Z',r'\right\rceil$. But suppose $\left\lceil Z,r\right\rceil$ is not in the image of $g$; then it is in one of the members of the list $\delta_1^i,...,\delta_r^i$ for some $i$, and since each of these members is an infinite set of basic elements of the same signature, we see that there are infinitely many basic elements of $N_{\beta_*}$ of signature $\tau_j$, contrary to our assumption. Hence, the proof is done.
\end{proof}
Lemma \ref{monuniform} can now be deduced by combining the last three lemmas, by a straightforward argument using Ehrenfeucht-Fraïssé games for the one-step language.
\subsection{Global neighborhood bisimulations}
Since the set of liftings $\{\Box,\Diamond\}$ can be shown to be expressively complete for $\mathcal{M}$, and since $\Diamond$ is just the dual of $\Box$,
the monadic second order language $\mathlang{MSO}_\mathcal{M}$ is equivalent to
the logic $\mathtt{MMSO}$ which has its syntax given by
$$
\varphi \mathrel{::=}
\mathtt{sr}(p) \mid p \subseteq q \mid \Box(p,q) \mid \exists p. \varphi
\mid \neg \varphi \mid \varphi \vee \varphi.
$$
The semantics of an atomic formula $\Box(p,q)$ in a neighborhood model
$\mathbb{S}$ is given, concretely, by the clause:
$(S,\sigma,V,u)\vDash \Box(p,q)$ if, for all $v \in V(p)$, there is
$Z \in \sigma(v)$ such that $Z \subseteq V(q)$.
At the present time, we do not know how to characterize the fragment of the
language $\mathtt{MMSO}$ that is invariant for arbitrary neighborhood
bisimulations.
However, the situation changes if we consider \textit{global} bisimulations
between neighborhood models.
\begin{defi}
A \textit{global neighborhood bisimulation} between $\mathcal{M}$-models
$\mathbb{S}_1$ and $\mathbb{S}_2$ is a neighborhood bisimulation $R$ that
satisfies the conditions:
\begin{description}
\item[Forth] For every $u \in S_1$ there is some $v\in S_2$ with $u R v$
\item[Back] For every $v \in S_2$ there is some $u \in S_1$ with $u R v$
\end{description}
\end{defi}
We now ask: what is the fragment of $\mathtt{MMSO}$ that is invariant
for global neighborhood bisimulations? Since global bisimulations are the natural
equivalence relation for modal logic with the
\textit{global modalities}, the most reasonable candidate would
be: the monotone modal $\mu$-calculus extended with the global modalities.
To be precise, let the \textit{monotone modal $\mu$-calculus with global
modalities}, denoted $\mu \mathtt{MML}_g$, be the language defined by
the grammar:
\begin{align*}
\varphi \mathrel{::=} p & \mid \neg p \mid \bot \mid \top \mid \Box \varphi
\mid \Diamond\varphi \mid [\forall] \varphi \mid [\exists] \varphi
\\ & \mid \varphi \vee \varphi \mid \varphi \wedge \varphi \mid\mu p. \varphi
\mid \nu p.\varphi
\end{align*}
where the formula $\varphi$ in $\mu p. \varphi$ and $\nu p . \varphi$ must be
positive in the variable $p$.
The new operators $[\forall]$ and $[\exists]$ are the global universal and
existential modalities, with their standard semantics:
$(\mathbb{S},u) \vDash [\forall] \varphi$ if $(\mathbb{S},v) \vDash \varphi$
for all $v \in S$, and $(\mathbb{S},u) \vDash [\exists] \varphi$ if
$(\mathbb{S},v) \vDash \varphi$ for some $v \in S$.
Given an $\mathcal{M}^{\star}$-model $\mathbb{S}$, let $\mathbb{S}_\mathcal{M}$ be the underlying $\mathcal{M}$-model. Conversely, given an $\mathcal{M}$-model $\mathbb{S} = (S,\sigma,V)$, define the $\mathcal{M}^{\star}$-model $\mathbb{S}^G = (S,\sigma^G,V)$ by setting $\sigma^G(s) = (\sigma(s),S)$.
The main result of this section is the following.
\begin{theo}
\label{janinwalukiewicztheorem}
A formula in $\mathtt{MMSO}$ is invariant for global neighborhood
bisimulations if, and only if, it is equivalent to a formula of the logic
$\mu \mathtt{MML}_g$.
\end{theo}
\begin{proof}
Clearly $\mu \mathtt{MML}_g$ translates into
$\mathtt{MMSO}$ and is invariant for global bisimulations.
Conversely, suppose $\varphi \in \mathtt{MMSO} $ is invariant for
global neighborhood bisimulations.
First observe that $\varphi$ can be regarded as a formula in
$\mathlang{MSO}_{\mathcal{M}^{\star}}$ as well.
More precisely, there is a formula $\varphi^* \in \mathlang{MSO}_{\mathcal{M}^{\star}}$
such that
\begin{equation}
\label{eq:y0}
(\mathbb{T},t) \vDash \varphi^* \text{ iff }
(\mathbb{T}_\mathcal{M},t) \vDash \varphi
\end{equation}
for any $\mathcal{M}^{\star}$-model $(\mathbb{T},t)$.
By Corollary~\ref{c:automatachar} there is a second-order $\{\Box,E\}$-automaton
$\mathbb{A}_{\phi}$ such that
\begin{equation}
\label{eq:y1}
\mathbb{A}_{\phi} \equiv \phi^{*} \text{ (on all $\mathcal{M}^{\star}$-tree models)}.
\end{equation}
Now we use the existence of an adequate, uniform construction for $\mathcal{M}^{\star}$
(Proposition~\ref{p:ad-monstar}).
Let $\mathbb{A}_{\phi}^{t}$ be the corresponding modal $\Lambda$-automaton given
by Proposition~\ref{p:unr}, where $\Lambda$ is the collection of all monotone,
natural predicate liftings for $\mathcal{M}^{\star}$.
By Proposition~\ref{p:lyndon} we may in fact assume that $\mathbb{A}_{\phi}^{t}$ is
a $\Theta$-automaton, where $\Theta = \{ \Box, \Diamond, E, E^{d} \}$.
Let $\psi = \psi_{\mathbb{A}_{\phi}^{t}}$ be the corresponding formula in
$\mathlang{\mu ML}_{\Theta }$.
We claim that, for any pointed neighborhood model $(\mathbb{S},s )$ we have
\begin{equation}
\label{eq:y2}
\mathbb{S},s \vDash \varphi \text{ iff }
\mathbb{S}^{G},s \vDash \psi.
\end{equation}
To prove this, consider the $\mathcal{M}^{\star}$-tree model $(\mathbb{T},R,r)$ given by
Proposition~\ref{p:unr}, applied to the pointed $\mathcal{M}^{\star}$-model
$(\mathbb{S}^{G},s)$.
Then there is a surjective $\mathcal{M}^{\star}$-coalgebra morphism $f: (\mathbb{T},r) \to
(\mathbb{S}^{G},s)$, and so in particular, $f$ is the graph of a global
neighborhood bisimulation between $\mathbb{T}_{\mathcal{M}}$ and $\mathbb{S}$ relating $r$ to $s$.
Gathering some facts we obtain the following chain of equivalences:
\begin{align*}
\mathbb{S},s \vDash \phi & \text{ iff } \mathbb{T}_{\mathcal{M}},r \vDash \phi
& \text{(assumption on $\phi$)}
\\ & \text{ iff } \mathbb{T},r \vDash \phi^{*}
& \text{\eqref{eq:y0}}
\\ & \text{ iff } \mathbb{T},R,r \vDash \mathbb{A}_{\phi}
& \text{\eqref{eq:y1}}
\\ & \text{ iff } \mathbb{S}^{G},s \vDash \mathbb{A}_{\phi}^{t}
& \text{(Proposition~\ref{p:unr})}
\\ & \text{ iff } \mathbb{S}^{G},s \vDash \psi
& \text{(assumption on $\psi$)}
\end{align*}
which proves \eqref{eq:y2} indeed.
Finally, let $\psi^{\forall} \in \mu\mathtt{MML}_g$ be the formula
we obtain from $\psi$ by replacing every occurrence of $E$ with $[\exists]$ and
every occurrence of $E^d$ with $[\forall]$.
It is a routine check to verify that
\begin{equation}
\label{eq:y3}
\mathbb{S}^{G},s \vDash \psi \text{ iff }
\mathbb{S},s \vDash \psi^{\forall}.
\end{equation}
But then the equivalence of $\phi \in \mathtt{MMSO}$ and
$\psi^{\forall} \in \mu \mathtt{MML}_g$ is immediate from
\eqref{eq:y2} and \eqref{eq:y3}.
\end{proof}
\bibliographystyle{abbrv}
|
\section{Introduction}
The spectral ball is the set of square matrices with spectral radius less than $1$. It appears naturally in Control Theory \cites{robustcontrol-proc, robustcontrol}, but is also of theoretical interest in Several Complex Variables.
\begin{definition}
The \emph{spectral ball} of dimension $n \in \NN$ is defined to be
\[
\sball_n := \{ A \in \mat{n}{n}{\CC} \;:\; \rho(A) < 1 \}
\]
where $\rho$ denotes the spectral radius, i.e.\ the modulus of the largest eigenvalue.
\end{definition}
Note that in dimension $n=1$, the spectral ball is just the unit disc. We will assume throughout the paper that $n \geq 2$.
\smallskip
The Nevanlinna--Pick problem is an interpolation problem for holomorphic functions on the unit disc $\udisk$. The classical \emph{Nevanlinna--Pick problem} for holomorphic functions $\udisk \to \udisk$ with interpolation in a finite set of points has been solved by Pick \cite{Pick} and Nevanlinna \cite{Nevanlinna}. The \emph{spectral Nevanlinna--Pick problem} is the analogue interpolation problem for holomorphic maps $\udisk \to \sball_n$:
\begin{quote}
Given $m \in \NN$ distinct points $a_1, \dots, a_m \in \udisk$, decide whether there is a holomorphic map $F \colon \udisk \to \sball_n$ such that
\[F(a_j) = A_j, \; j = 1, \dots, m\]
for given matrices $A_1, \dots, A_m \in \sball_n$.
\end{quote}
It has been studied by many authors, in particular by Agler and Young for dimension $n=2$ and generic interpolation points \cites{AglerYoung2000, AglerYoung2004}. Bercovici \cite{Bercovici} has found solutions for $n=2$, and Costara \cites{Costara2} has found solutions for generic interpolation points in higher dimensions. In general, the problem is still open.
The spectral ball $\sball_n$ can also be understood in the following way: Denote by $\sigma_1, \dots, \sigma_n \colon \CC^n \to \CC$ the elementary symmetric polynomials in $n$ complex variables. Let $ \mathrm{EV} \colon \mat{n}{n}{\CC} \to \CC^n$ assign to each matrix a vector of its eigenvalues.
Then we denote by $\pi_1 := \sigma_1 \circ \mathrm{EV}, \dots, \pi_n := \sigma_n \circ \mathrm{EV}$ the elementary symmetric polynomials in the eigenvalues. By symmetrizing we avoid any ambiguities of the order of eigenvalues and obtain a polynomial map $\pi = (\pi_1, \dots, \pi_n)$, symmetric in the entries of matrices in $\mat{n}{n}{\CC}$, actually
\[
\chi_A(\lambda) = \lambda^n + \sum_{j=1}^n (-1)^j \cdot \pi_j(A) \cdot \lambda^{n-j}
\]
where $\chi_A$ denotes the characteristic polynomial of $A$.
Now we can consider the holomorphic surjection $\pi \colon \Omega_n \to \sdisk_n$ of the spectral ball onto the symmetrized polydisc $\sdisk_n := (\sigma_1, \dots, \sigma_n)(\udisk^n)$. A generic fibre, i.e.\ a fibre above a base point with no multiple eigenvalues, consists exactly of one equivalence class of similar matrices. Thus, a generic fibre is actually a $\slgrp_n(\CC)$-homogeneous manifold where the group $\slgrp_n(\CC)$ acts by conjugation. A singular fibre decomposes into several strata which are $\slgrp_n(\CC)$-homogeneous manifolds as well, but not necessarily connected.
Given this holomorphic surjection, it is natural to consider a weaker version of the spectral Nevanlinna--Pick problem, which is called the \emph{spectral {Nevan\-linna}\-{--}Pick lifting problem}:
\begin{quote}
Given $m \in \NN$ distinct points $a_1, \dots, a_m \in \udisk$, and a holomorphic map $f \colon \udisk \to \sdisk_n$ with $f(a_j) = \pi(A_j)$ for given matrices $A_1, \dots, A_m \in \sball_n$,
decide whether there is a holomorphic map $F \colon \udisk \to \sball_n$ such that \[F(a_j) = A_j, \; j = 1, \dots, m.\]
i.e.\ such that the following diagram commutes:
\begin{center}
$
\begin{diagram}
\node[2]{\sball_n} \arrow{s,r}{\pi} \\
\node{\udisk} \arrow{ne,t,--}{F} \arrow{e,t}{f} \node{\sdisk_n} \\
\end{diagram}
\vspace*{-12mm}
$
\\
$
a_1, \dots, a_m \mapsto \pi(A_1), \dots, \pi(A_m)
$
\end{center}
\end{quote}
When this lifting problem is solved, the spectral Nevanlinna--Pick problem reduces to an interpolation problem $\udisk \to \sdisk_n$. In contrast to the spectral ball, the symmetrized polydisc $\sdisk_n$ is a taut domain, and should be more accessible with techniques from hyperbolic geometry.
Solutions to this lifting problem have been found for dimensions $n = 2, 3$ by Nikolov, Pflug and Thomas \cite{TwoThree} and recently for dimensions $n = 4, 5$ by Nikolov, Thomas and Tran \cite{locallift}. They also provide the solution to a localised version of the spectral Nevanlinna--Pick lifting problem:
\begin{proposition}[\cite{locallift}*{Proposition 7}]
\label{locallifting}
Let $f \in \hol(U, \sdisk^n)$, where $U$ is a neighborhood of $p \in \udisk$, and let $M \in \sball_n$.
Then there exists a neighborhood $U^\prime \subset U$ of $p$ and
$F \in \hol(U^\prime, \sball_n)$ such that
$\pi \circ F = f$, $F(p) = M$ and $F(v)$ is cyclic for $v \in U^\prime \setminus \{p\}$ if and only if
\begin{equation}
\label{localcondition}
\tag{$\ast$}
\left.\frac{d^k}{d \lambda^k} {{\mbox{\Large$\chi$}}}_{f(v)}(\lambda)\right|_{\lambda = \lambda_j} = O((v - p)^{d_{n_j-k}(B_j)}), \quad 0 \leq k \leq n_j - 1, \; 1 \leq j \leq s
\end{equation}
where $B_1 \in \mat{n_1}{n_1}{\CC}, \dots, B_s \in \mat{n_s}{n_s}{\CC}$ are the Jordan blocks of $M \simeq B_1 \oplus \dots \oplus B_s$ and
\[
\begin{split}
d_\ell(B_j) = \min\Big\{ d \in \NN \,:\; \exists x_1, \dots x_d \in \CC^n\, \text{s.t.} \\ \dim \mathop{\mathrm{span}_\CC}\left\{ B^{k_1} x_1, \dots B^{k_d} x_d; \, k_1, \dots, k_d \geq 0\right\} \geq \ell \Big\}
\end{split}
\]
\end{proposition}
A matrix $M \in \mat{n}{n}{\CC}$ is called \emph{cyclic}, if it admits a cyclic vector $x \in \CC^n$, i.e.\ $\mathop{\mathrm{span}_\CC} \left\{ M^k \cdot x, \; k \in \NN\right\} = \CC^n$. This is equivalent to saying that there is only one Jordan block per eigenvalue in the Jordan block decomposition of $M$.
\begin{remark}
It is easy to see that the condition \eqref{localcondition} is necessary: it is a direct consequence of the factorisation $\pi \circ F = f$ on the jet level, explicitly stated using the Jordan block decomposition.
\end{remark}
In this paper we prove that the solution of the localised spectral Nevanlinna--Pick lifting problem implies the global one:
\begin{theorem}
\label{mainthm}
The spectral Nevanlinna--Pick lifting problem can be solved if and only if it can be solved locally around the interpolation points which means that \eqref{localcondition} holds for each interpolation point.
\end{theorem}
As a by-product of the proof we obtain (see Corollary \ref{noSPSH}) that the spectral ball admits no bounded from above strictly plurisubharmonic functions, but non-constant bounded holomorphic functions.
\pagebreak
\section{Oka Theory}
In this section we provide some background in Oka theory that will be needed in the proof of the main theorem. The goal is to find conditions for the existence of holomorphic liftings or holomorphic sections. We first recall the following equivalence:
\begin{remark}
\label{pullbacklifting}
For a holomorphic surjection $\pi \colon Z \to X$ and a holo\-morphic map $f \colon Y \to X$, we denote the pull-back of $Z$ under $f$ by
\begin{equation*}
f^\ast(Z) = \left\{ (y, z) \in Y \times Z \;:\; f(y) = \pi(z) \right\}
\end{equation*}
Together with the natural projections to the first resp. second factor $f^\ast(\pi) \colon f^\ast(Z) \to Y$ resp. $f^\ast(Z) \to Z$ we obtain the following commutative diagram
\[
\begin{diagram}
\node{f^\ast(Z)} \arrow[2]{e,t}{} \arrow{s,l}{f^\ast(\pi)} \node[2]{Z} \arrow{s,r}{\pi} \\
\node{Y} \arrow{e,=} \node{Y} \arrow{e,t}{f} \arrow{ne,t,--}{F} \arrow{nw,t,--}{f^\ast(F)} \node{X}
\end{diagram}
\]
We see that the existence of the lifting $F$ is equivalent to the existence of the lifting $f^\ast(F), \, y \mapsto (y, F(y))$. Thus, the lifting problem can always be reduced to finding a section of $f^\ast(Z) \to Y$.
\end{remark}
The crucial ingredient will be an Oka theorem for branched holomorphic maps due to Forstneri\v{c}.
From \cite{Okamulti} we recall the following definitions:
\begin{definition}
A point $z \in Z$ is a \emph{branching point} of a holomorphic surjection between complex spaces $\pi \colon Z \to X$ if $\pi$ is not submersive in $z$. The \emph{branching locus} of $\pi$ is denoted by $\mathop{\mathrm{br}} \pi$ and consists of all branching points. For maps between complex spaces, also $Z_{\mathrm{sing}}$ and $\pi^{-1}\left(X_{\mathrm{sing}}\right)$ are included in $\mathop{\mathrm{br}} \pi$ by convention.
\end{definition}
\begin{definition}
A holomorphic surjection $\pi \colon Z \to X$ between complex spaces is called an \emph{elliptic submersion} over an open subset $V \subseteq X_{\mathrm{reg}}$ if
\nopagebreak
\begin{enumerate}
\item $\pi|\pi^{-1}(V)$ is a submersion of complex manifolds
\item each point $x \in V$ has an open neighborhood $U \subseteq V$ such that there exists a holomorphic vector bundle $E \to \pi^{-1}(U)$ together with a so-called holomorphic \emph{dominating spray} $s \colon E \to \pi^{-1}(U)$ satisfying the following conditions for each $z \in \pi^{-1}(U)$:
\begin{enumerate}
\item $s(E_z) \subseteq Z_{\pi(z)}$
\item $s(O_z) = z$
\item the derivative $d s \colon T_{0_z} E \to T_z Z$ maps the subspace $E_z \subseteq T_{0_z} E$ surjectively onto the vertical tangent space $\mathop{\mathrm{ker}} d_z \pi$.
\end{enumerate}
\end{enumerate}
\end{definition}
A examination of the proof in \cite{Okamulti} shows that a small variation of the statement of \cite{Okamulti}*{Theorem 2.1} is valid with exactly the same proof:
\begin{theorem}
\label{Okabranched}
Let $X$ be a Stein space and $\pi \colon Z \to X$ be a holomorphic map of a complex space $Z$ onto $X$. Assume that $Z \setminus \mathop{\mathrm{br}}{\pi}$ is an elliptic submersion over $X$.
Let $X_0$ be a (possibly empty) closed subvariety of $X$.
Given a continuous section $F \colon X \to Z$ which is holomorphic in an open set containing $X_0$ and such that $F(X \setminus X_0) \subseteq Z \setminus \mathop{\mathrm{br}}{\pi}$,
there is for any $k \in \NN$ a homotopy of continuous sections $F_t \colon X \to Z, \; t \in [0, 1],$ such that $F_0 = F$, for each $t \in [0, 1]$ the section $F_t$ is holomorphic in a neighborhood of $X_0$ and tangent to order $k$ along $X_0$, and $F_1$ is holomorphic on $X$.
If $F$ is holomorphic in a neighborhood of $K \cup X_0$ for some compact, holomorphically convex subset $K$ of $X$ then we can choose $F_t$ to be holomorphic in a neighborhood of $K \cup X_0$ and to approximate $F = F_0$ uniformly on $K$.
\end{theorem}
\begin{remark}
The difference to \cite{Okamulti}*{Theorem 2.1} is only that we allow $X_0$ to be smaller than $\pi (\mathop{\mathrm{br}} \pi)$ provided that the section over $X_0 \setminus \pi (\mathop{\mathrm{br}} \pi)$ does not hit the branching locus $\mathop{\mathrm{br}} \pi$.
\end{remark}
\section{Proof}
In case of the spectral Nevanlinna--Pick lifting problem, we have $X = \sdisk_n, Y = \udisk, Z = \sball_n$, and $X_0 := \{a_1, \dots, a_m \}$ is a closed subvariety in $\udisk$. We note that $Y$ is indeed Stein and that $\pi$ is an elliptic submersion in the cyclic matrices since the largest strata of all fibres are $\slgrp_n(\CC)$-homogeneous spaces of dimension $n$, see for example \cite{Okabook}*{Example 5.5.13}.
We write down the spray explicitly for any $U \subseteq \sdisk_n$: we choose the bundle $E := \slalg_n(\CC) \times \pi^{-1}(U) \to \pi^{-1}(U)$ with the natural projection, and define the spray $s \colon E \to \pi^{-1}(U)$ as
\[
(B, A) \mapsto \exp(B) \cdot A \cdot \exp(-B)
\]
The derivative in $(0, A)$, evaluated for a tangent vector $C$, is then given by the adjoint representation of $\slalg_n(\CC)$:
\[
d_{(0, A)} s (C) = [A, C]
\]
and the conditions of the elliptic submersion are easily verified.
Now, by remark \ref{pullbacklifting} and the fact that the pullback of an elliptic submersion is still an elliptic submersion (see \cite{Okamulti}*{Lemma 3.1}) we translate the problem of finding a lifting $F \colon X \to Z$ with $\pi \circ F = f$ into a problem of finding a section of $f^{\ast}(\pi) \colon f^{\ast}(\sball_n) \to \udisk$.
\medskip
Due to a lack of reference, we also include the following lemma and its proof:
\begin{lemma}
\label{connected}
Each fibre, the largest stratum of each fibre and also each connected component of any stratum of each fibre of $\pi \colon \sball_n \to \sdisk_n$ is $\CC$-connected.
\end{lemma}
\begin{proof}
Given any two similar matrices $B$ and $C$, we can connect them through finitely many holomorphic images of complex lines:
there exists a finite number of matrices $Q_1, \dots, Q_s \in \slalg_n(\CC)$ such that
\[
C = \exp(Q_s) \cdots \exp(Q_1) \cdot B \cdot \exp(-Q_1) \cdots \exp(-Q_s).
\]
By appending a complex time parameter $t_s$ in front of each $Q_s$ we obtain $s$ lines connecting $B$ to $C$. This proves that the largest stratum (hence also a generic fibre) or any connected component of a stratum is $\CC$-connected.
For the connectedness of a singular fibre it is now sufficient to connect the strata. This can be done by introducing parameters in the off-diagonal entries of the Jordan-block decomposition.
\end{proof}
We obtain the following corollary which will however not play any role in the proof of the main theorem:
\begin{cor}
\label{noSPSH}
The spectral ball $\sball_n, n \geq 2,$ is a union of immersed complex lines. Hence there exists no bounded from above strictly pluri\-subharmonic function on $\sball_n$.
\end{cor}
\begin{proof}
The pull-back of any (strictly) plurisubharmonic function to any of these complex lines is (strictly) subharmonic, and Liouville's Theorem applies.
\end{proof}
\bigskip
Now we are ready to prove Theorem \ref{mainthm}:
First we need understand the branching locus of the pull-back of $\pi \colon \sball_n \to \udisk$ by $f$, $f^\ast(\pi) \colon f^\ast(\sball_n) \to \udisk$.
Let $(z, A) \in f^\ast(\sball_n) \subset \udisk \times \sball_n \subset \CC \times \CC^{n^2}$. Assume that $(z, A) \in f^\ast(\sball_n)$ is a singularity. This means that the derivative of the defining equations of $f^\ast(\sball_n)$ has not full rank in $(z, A)$,
\begin{equation*}
\exists v \in \mat{n}{1}{\CC}, v \neq 0, \;:\; \left( f'(z), -d_A \pi \right)^\intercal \cdot v = 0
\end{equation*}
But then also $(d_A \pi)^\intercal \cdot v = 0$ and necessarily $A \in \mathop{\mathrm{br}} \pi$.
Assume that $(z, A) \in f^\ast(\sball_n)$ is a branching point of $f^\ast(\pi)$ which projects $(z, A)$ to $z$, i.e.\ for $d_{(z, A)} f^\ast(\pi) = (1, 0, \dots, 0)^t$ it holds outside singularities that
\begin{equation*}
\exists w \in \mat{n}{1}{\CC}, w \neq 0, \;:\; \left( f'(z), -d_A \pi \right)^\intercal \cdot w = (1,0, \dots, 0)^\intercal
\end{equation*}
But then again $(d_A \pi)^\intercal \cdot w = 0$ and necessarily $A \in \mathop{\mathrm{br}} \pi$.
Hence, $\mathop{\mathrm{br}} f^\ast(\pi)$ which by definition includes also the singularities of $f^\ast(\sball_n)$, is contained in the pull-back of $\mathop{\mathrm{br}} \pi$, and no new branching points are introduced.
\smallskip
Theorem \ref{Okabranched} gives the desired section $f^{\ast}(F) = g_1$, once the existence of $g_0$ is clear. Proposition \ref{locallifting} provides the necessary local liftings in each point $a_1, \dots, a_m$ which are pulled back by $f$ to local holomorphic sections such that they do not hit $\mathop{\mathrm{br}}{f^\ast(\pi)}$ except for possibly the interpolation point itself.
Choosing the neighborhoods for the local sections small enough and contractible, we achieve that $g_0$ is holomorphic in a neighborhood $U$ of $X_0$.
\smallskip
It remains to show that we can extend $g_0$ as continuous section over $\udisk \setminus U$. Since all the components of $U$ are contractible, this is equivalent to find a continuous section $g_0$ which interpolates the points.
We can connect the finitely many points $a_1, \dots, a_m$ by arcs such that the union $\gamma$ of these arcs is homeomorphic to an interval, and then contract the unit disc to $\gamma$. Since all the fibres are connected (even $\CC$-connected by Lemma \ref{connected}) there is no topological obstruction to construct a continuous interpolating section $\gamma \to \sball_n$.
We have proved Theorem \ref{mainthm}.
\begin{bibdiv}
\begin{biblist}
\bib{AglerYoung2000}{article}{
author={Agler, J.},
author={Young, N. J.},
title={The two-point spectral Nevanlinna-Pick problem},
journal={Integral Equations Operator Theory},
volume={37},
date={2000},
number={4},
pages={375--385},
issn={0378-620X},
review={\MR{1780117 (2001g:47025)}},
doi={10.1007/BF01192826},
}
\bib{AglerYoung2004}{article}{
author={Agler, Jim},
author={Young, N. J.},
title={The two-by-two spectral Nevanlinna-Pick problem},
journal={Trans. Amer. Math. Soc.},
volume={356},
date={2004},
number={2},
pages={573--585 (electronic)},
issn={0002-9947},
review={\MR{2022711 (2005d:47032)}},
doi={10.1090/S0002-9947-03-03083-6},
}
\bib{Bercovici}{article}{
author={Bercovici, Hari},
title={Spectral versus classical Nevanlinna-Pick interpolation in
dimension two},
journal={Electron. J. Linear Algebra},
volume={10},
date={2003},
pages={60--64},
issn={1081-3810},
review={\MR{2001974 (2004g:47019)}},
}
\bib{robustcontrol-proc}{article}{
author={Bercovici, Hari},
author={Foias, Ciprian},
author={Tannenbaum, Allen},
title={Spectral radius interpolation and robust control},
conference={
title={Proceedings of the 28th IEEE Conference on Decision and
Control, Vol.\ 1--3},
address={Tampa, FL},
date={1989},
},
book={
publisher={IEEE, New York},
},
date={1989},
pages={916--917},
review={\MR{1038978 (91b:93052)}},
}
\bib{robustcontrol}{article}{
author={Bercovici, Hari},
author={Foias, Ciprian},
author={Tannenbaum, Allen},
title={A spectral commutant lifting theorem},
journal={Trans. Amer. Math. Soc.},
volume={325},
date={1991},
number={2},
pages={741--763},
issn={0002-9947},
review={\MR{1000144 (91j:47006)}},
doi={10.2307/2001646},
}
\bib{Costara2}{article}{
author={Costara, Constantin},
title={On the spectral Nevanlinna-Pick problem},
journal={Studia Math.},
volume={170},
date={2005},
number={1},
pages={23--55},
issn={0039-3223},
review={\MR{2142182 (2006d:30054)}},
doi={10.4064/sm170-1-2},
}
\bib{Okamulti}{article}{
author={Forstneri{\v{c}}, Franc},
title={The Oka principle for multivalued sections of ramified mappings},
journal={Forum Math.},
volume={15},
date={2003},
number={2},
pages={309--328},
issn={0933-7741},
review={\MR{1956971 (2003m:32015)}},
doi={10.1515/form.2003.018},
}
\bib{Okabook}{book}{
author={Forstneri{\v{c}}, Franc},
title={Stein manifolds and holomorphic mappings},
series={Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge.
A Series of Modern Surveys in Mathematics
},
volume={56},
note={The homotopy principle in complex analysis},
publisher={Springer, Heidelberg},
date={2011},
pages={xii+489},
isbn={978-3-642-22249-8},
isbn={978-3-642-22250-4},
review={\MR{2975791}},
doi={10.1007/978-3-642-22250-4},
}
\bib{Nevanlinna}{article}{
author={Nevanlinna, Rolf},
title={\"Uber beschr\"ankte Funktionen, die in gegebenen Punkten vorgeschriebene Werte annehmen.},
journal={Ann. Acad. Sci. Fenn., Ser. A},
volume={13},
date={1920},
}
\bib{TwoThree}{article}{
author={Nikolov, Nikolai},
author={Pflug, Peter},
author={Thomas, Pascal J.},
title={Spectral Nevanlinna-Pick and Carath\'eodory-Fej\'er problems for
$n\leq3$},
journal={Indiana Univ. Math. J.},
volume={60},
date={2011},
number={3},
pages={883--893},
issn={0022-2518},
review={\MR{2985860}},
doi={10.1512/iumj.2011.60.4310},
}
\bib{locallift}{article}{
author={Nikolov, Nikolai},
author={Thomas, Pascal J.},
author={Tran, Duc-Anh},
title={Lifting maps from the symmetrized polydisk in small dimensions},
date={2014},
eprint={arXiv:1410.8567},
}
\bib{Pick}{article}{
author={Pick, Georg},
title={\"Uber die Beschr\"ankungen analytischer Funktionen, welche durch
vorgegebene Funktionswerte bewirkt werden},
language={German},
journal={Math. Ann.},
volume={77},
date={1915},
number={1},
pages={7--23},
issn={0025-5831},
review={\MR{1511844}},
doi={10.1007/BF01456817},
}
\end{biblist}
\end{bibdiv}
\end{document}
|
\section*{Introduction}
Although the problem of deciding the splitting of vector bundles is very classical, only relatively few cases have been settled: there are cohomological criteria for products of projective spaces and quadrics \cite{c-mr,ba-ma}, Grassmannians \cite{ottv,mals}, hypersurfaces in projective spaces \cite{swd,bis+rav}. Splitting criteria corresponding to restrictions are useful because they yield dimensional reductions: the problem is reduced to a (usually much) lower dimensional variety, where one can use further cohomological tools.
Horrocks' criterion \cite{hor} states that a vector bundle on $\mbb P^n$, $n\ges 3$, splits if and only if its restriction to some hyperplane $D\cong\mbb P^{n-1}$ splits. This was generalized in \cite{ba}, where the author restricted vector bundles on `\emph{Horrocks varieties}' to \emph{ample divisors}. The ampleness assumption excludes several natural situations, \textit{e.g.} the case of morphisms, where one wishes to restrict vector bundles either to pre-images of ample divisors or to relatively ample ones. Also, the `Horrocks variety' assumption is a rather restrictive cohomological property.
The goal of this note is to generalize the splitting criterion in \textit{op. cit.} to include $q$-ample divisors; this covers the case of morphisms mentioned before. We obtain two types of results: \emph{splitting} and \emph{triviality} criteria.
\begin{thm-nono}{(splitting criteria)}
Let $(X,\eO_X(1))$ be a smooth, complex projective variety, with $\dim X\ges3$. Let $\crl V$ be a vector bundle on $X$, $\crl E:=\cEnd(\crl V)$ the bundle of endomorphisms, $\eL\in\Pic(X)$ be $q$-ample, and $D\in|d\eL|$. The equivalence
$$\;[\;\crl V\;\text{splits}\;\Leftrightarrow\;\crl V\otimes\eO_D\;\text{splits}\;]\;$$
holds in any of the following cases:
\begin{enumerate}
\item[(a)]
If $q\les\dim X-3$, $D$ is either reduced or irreducible, and $H^1(\crl E_D\otimes\eL_D^{-a})=0$ for all $a\ges d$. The parameter $d$ is bounded from below by a linear function in the regularity of $\crl E$ with respect to $\eO_X(1)$.
\smallskip
\item[(b)]
If $X$ is $2$-split, $q\les\dim X-4$, and $D$ is smooth.
\smallskip
\item[(c)]
If $X$ is $1$-split, $q\les\dim X-4$, $\eL$ is globally generated, and $D\in|\eL|$ is very general.
\end{enumerate}
\end{thm-nono}
The conditions $1$-, $2$-split are respectively the notions of `splitting' and `Horrocks variety' in \cite{ba}. Bakhtary's result corresponds to the case (b) above with $q=0$.
The statement (a) is more effective in the case where $\eL$ is relatively ample with respect to a morphism (cf. Theorem \ref{thm:rel-ample}): it suffices $D\in|\eL|$ to be weakly normal and $\crl E\otimes\eL$ be relatively ample.
\smallskip
We illustrate the advantage of allowing $q$-ample line bundles by discussing several explicit examples. First we simplify the cohomological splitting criteria for products of projective spaces and quadrics \cite{c-mr,ba-ma}. They involve a large number of cohomological tests and, by restricting to `sub-products', the number of the tests is massively reduced.
Second, we use the results in \cite{hal2} and we deduce a splitting criterion for vector bundles on varieties $X$ which are products of minuscule homogeneous varieties (cf. Theorem \ref{thm:minuscule}). To our best knowledge, currently there are \emph{no known results} in this direction; even a product of Grassmannians seems to be uncovered. We prove that, for such $X$, it is enough to test the splitting on appropriate products of planes $\mbb P^2$ (Schubert subvarieties) in $X$. \smallskip
The trivializable vector bundles are particular cases of the split ones, so the triviality criteria below hold in greater generality. Notably, one can eliminate the conditions $1$-, $2$-split.
\begin{thm-nono}{(triviality criteria)}
Let $X,\eL,\crl V$ be as above, and $D\in|\eL|$.
\nit The equivalence
$\;[\,\crl V\;\text{is trivial}\;\Leftrightarrow\;\crl V_D\;\text{is trivial}\,]\;$
holds in any of the following cases:
\begin{enumerate}
\item[(a)]
If $\eL\in\Pic(X)$ is semi-ample and $(\dim X-3)$-ample.
\item[(b)]
If $\eL$ is relatively ample for a morphism $X\srel{f}{\to} Y$ of relative dimension at least three.
\item[(c)]
If the anti-canonical bundle $\omega_X^{-1}$ is $(\dim X-3)$-ample, $X$ is Frobenius split by a power of a section $\si$ in $\omega_X^{-1}$, and $D=\dvs(\si)$.
\end{enumerate}
\end{thm-nono}
The condition (c) above is particularly suited for spherical varieties (\textit{e.g.} toric varieties), because they satisfy the assumption about the Frobenius splitting. We elaborate on the case of toric varieties.
\begin{thm-nono}{\kern-.75ex}
Let $X$ be a smooth toric variety and $\Delta$ be its boundary divisor. We assume that
$$
\codim\big(\sbloc(\omega_{X}^{-1})\,\big) \ges 3.
$$
Then, for a vector bundle $\crl V$ on $X$, one has the equivalences:
\begin{enumerate}
\item[(a)]
$[\,\crl V\mbox{ splits}\;\Leftrightarrow\;\crl V_{\Delta_m}\mbox{ splits}\,]$, for $m\gg0$.
\item[(b)]
$[\,\crl V\mbox{ is trivial}\;\Leftrightarrow\;\crl V_{\Delta}\mbox{ is trivial}\,]$.
\end{enumerate}
\end{thm-nono}
The splitting criteria obtained in this article are based on two technical ingredients: the `universal' criterion \ref{prop:h1}, on one hand, and various Kodaira-type vanishing theorems for $q$-ample line bundles, on the other hand.
For this reason, the role of the appendices is twofold: first, to recall the definitions and properties of the $q$-ampleness (cf. \cite{totaro,ottem}) and Frobenius splitting (cf. \cite{brion-kumar}) which are used in the body of the article; second, they contain a few (possibly) noteworthy results:
\begin{itemize}[leftmargin=*]
\item
a Kodaira vanishing theorem for relatively ample line bundles on weakly normal varieties (cf. \ref{thm:global-gen}(iii)) and for $q$-ample line bundles on Frobenius-split varieties (cf. \ref{thm:F-split});
\item
a Picard-Lefschetz property in the relative setting (cf. the Theorem \ref{thm:pic-iso});
\item
a $q$-ampleness criterion for line bundles which are not necessarily globally generated (cf. Theorem \ref{thm:q-ample}).
\end{itemize}
Throughout this article, $X$ stands for a smooth projective variety over $\kk$ of dimension at least three.
\section{The general splitting principle}\label{sct:split-vb}
\begin{m-definition}\label{def:split}
Let $T$ be a scheme defined over $\kk$ and $S$ be a closed subscheme of it; we assume that
$H^0(\eO_S)=H^0(\eO_T)=\kk.$
For a locally free sheaf (a vector bundle) $\crl V_T$ of rank $r$ on $T$, we denote $\crl E_T:=\cEnd(\crl V_T)$ the sheaf of endomorphisms; let $\crl V_S:=\crl V_T\otimes_{\eO_T}\eO_S$, \textit{etc}.
An \emph{eigenvalue} of $h_S\in H^0(\crl E_S)$ is a complex root of the polynomial
\begin{equation}\label{eq:ph}
p_{h_S}:=\det\big(t\bone-h_S\big)\in
H^0\big(\,
\cEnd(\det\crl V_S)
\,\big)[t]
= H^0(\eO_S)[t]=\kk[t].
\end{equation}
We say that $\crl V_T$ \textit{splits} if it is isomorphic to a direct sum of $r$ invertible sheaves (line bundles) on $T$.
\end{m-definition}
Let $S,h_S$ be as above. If $\veps\in\kk$ is an eigenvalue of $p_{h_S}$, then
$\Ker(\veps\bone-h_S)\subset\crl V_S$ is a non-zero $\eO_S$-module: indeed, for a closed point $x\in S_{\rm red}\subset S$ with maximal ideal $\mfrak m_x\subset\eO_S$, $\veps$ is an (usual) eigenvalue of $h_S\otimes\frac{\eO_S}{\mfrak m_x}\in\End\big( \crl V_S\otimes\frac{\eO_S}{\mfrak m_x} \big)$.
\begin{m-lemma}\label{lm:split}
Let the notation be as above. Then the following statements hold:
\begin{enumerate}
\item
$\crl V_S$ splits if and only if there is $h_S\in H^0(\crl E_S)$ with $r$ pairwise distinct eigenvalues.
\item
If $H^0(\crl E_{T})\to H^0(\crl E_{S})$ is surjective, then $\crl V_{T}$ splits if and only if $\crl V_S$ splits.
\end{enumerate}
\end{m-lemma}
\begin {m-proof}
(i)
Suppose that $h_S\in H^0(\crl E_S)$ has pairwise distinct eigenvalues $\veps_1,\ldots,\veps_r\in\kk$, and let $\ell_j:=\Ker(\veps_j\bone-h_S)\neq0$. The (polynomial) identity
\\ \centerline{$
\mbox{$
1=\ouset{j=1}{r}{\sum}c_jp_j(t),\quad\text{with}\;
c_j:=\biggl(\underset{k\neq j}{\prod}(\veps_j-\veps_k)\biggr)^{-1}\in\kk,
\quad p_j(t):=\underset{k\neq j}{\prod}(t-\veps_k)\in\kk[t],
$}
$}\\[.5ex]
implies $\crl V_S\,{=}\ouset{j=1}{r}{\sum}\,{\rm Image}\big(p_j(h_S)\big)$.
Moreover, the Cayley-Hamilton theorem yields:
\\ \centerline{$
{\rm Image}\big(p_j(h_S)\big)\subset\ell_j\;\Rightarrow\;
\mbox{$\crl V_S\,{=}\ouset{j=1}{r}{\sum}\,\ell_j.$}
$}\\[1ex]
We claim that the sum is direct. Indeed, if $v_j\in\ell_j$ for all $j$, then holds:
$$
\begin{array}{rcll}
v_1+v_2+\ldots+v_r=0 &{\Rightarrow}& p_1(h_S)(v_1)=0,
&
\mbox{(since $t-\veps_j\mid p_1(t)$, for $j\ges 2$),}
\\[1ex]
v_1=\ouset{j=2}{r}{\sum}c_j\cdot p_j(h_S)(v_1)&=&0,\;etc.
&
\mbox{(since $v_1\in\Ker(\veps_1\bone-h_S)$).}
\end{array}
$$
Hence $\ell_j$, $j=1,\ldots,r$, are (locally) projective $\eO_S$-modules, so they are locally free (cf. \cite[Ch.\,II, \S 5.2, Th\'eor\`eme 1]{bour-comm1-4}) of rank one.
\nit(ii)
If $\crl V_S$ splits, there is $h_S\in H^0(\crl E_S)$ with $r$ pairwise distinct complex eigenvalues; take an extension $h_{T}\in H^0(\crl E_{T})$ of it. The equation \eqref{eq:ph} shows that $p_{h_T}{=}\,\det(t\bone-h_{T})\in\kk[t]$, so $p_{h_T}=p_{h_S}$, hence $h_T$ has the same eigenvalues as $h_S$.
\end {m-proof}
\begin{m-definition}\label{def:D-thick}
Let $\eL$ be an invertible sheaf (a line bundle) on $X$ and $D\in|d\eL|$ be an effective divisor. For $m\ges 0$, the $m$-th order thickening $D_m$ of $D$ is the subscheme of $X$ defined by the ideal $\eI_D^{m+1}$, where $\eI_D=\eO_X(-D)\cong\eL^{-d}$.
\end{m-definition}
The structure sheaves of successive thickenings fit into the exact sequences:
\begin{equation}\label{eqn:thick}
0\to\eL_D^{-dm}\to\eO_{D_m}\to\eO_{D_{m-1}}\to0,\quad m\ges1.
\end{equation}
We will apply the Lemma \ref{lm:split} mostly in the case $T=X, S=D_m$, for suitable $m$.
\begin{m-remark}
If $D$ is an effective divisor of $X$, the surjectivity of $H^0(\crl E_X)\to H^0(\crl E_D)$ is implied by the vanishing of $H^1(\eO_X(-D)\otimes\crl E_X)$. Thus, at a certain extent, the $(\dim X-2)$-amplitude of $\eO_X(D)$ is the weakest possible assumption which allows to deduce splitting criteria for vector bundles by restricting them to $D$.
\end{m-remark}
In the framework of formal schemes, we have the following very general statement.
\begin{m-proposition}\label{prop:formal}
Let $D\subset X$ be an effective divisor, $\dim X\ges2$, and $\hat X:=\disp\varprojlim_{m}D_m$ denote the formal completion of $X$ along $D$. If the cohomological dimension $\cd(X\sm D)\les\dim X{-}2$, then $\crl V$ splits if and only if $\crl V\otimes\eO_{\hat X}$ does. The assumption is satisfied if $D$ is $(\dim X{-}2)$-ample.
\end{m-proposition}
\begin{m-proof}
By \cite[Theorem 3.4]{hart-as}, the restrictions
$$
\kk\cong H^0(\eO_X)\to H^0(\eO_{\hat X}),\quad
H^0(\crl E)\to H^0(\crl E_{\hat X})
$$
are both isomorphisms; we conclude by \ref{lm:split}(ii). The last claim is \cite[Proposition 5.1]{ottem}.
\end{m-proof}
\begin{m-lemma}\label{lm:q-split}
Let $\eL$ be a $q$-ample line bundle on $X$, with $q\les\dim X-2$. Consider $D\in|\eL|$ which is either reduced or irreducible. Then the following statements hold:
\begin{enumerate}
\item
$H^0(\eO_{D_m})=\kk$, for all $m\ges0$;
\item
$\crl V_X$ splits if and only if its restriction $\crl V_{D_m}$ splits, for some $m\gg 0$.
\end{enumerate}
The conclusions (i), (ii) above hold for an arbitrary $D\in|\eL|$ in any of the situations enumerated below:
\begin{equation}\label{eq:C}
\mbox{
\begin{minipage}{0.85\linewidth}
\nit{\rm(a)}
$\eL$ is semi-ample;
\nit{\rm(b)}
$\eL$ is relatively ample for a morphism $X\srel{f}{\to}Y$, $\dim X-\dim Y\ges2$;
\nit{\rm(c)}
$X$ is an F-split variety (cf. appendix \ref{sct:frob}).
\end{minipage}
}
\end{equation}
\end{m-lemma}
\begin {m-proof}
(i) The $(\dim X-2)$-amplitude of $\eL$ implies that $\kk= H^0(\eO_X)\to H^0(\eO_{D_m})$ is an isomorphism for $m\gg0$, so $D$ is connected.
If $D$ is reduced, then we have $H^0(\eO_D)=\kk$. As $H^0(\eL_D^{-m})=0$ for $m\gg0$, the same holds for all $m\ges1$. The conclusion follows from \eqref{eqn:thick}.
If $D$ is irreducible, then $D=m_0D_{\rm red}$ for some $m_0\ges1$, and the previous argument shows that $H^0(\eO_{D_{{\rm red},m}})=\kk$ for all $m\ges1$.
Concerning the final claim, in the cases \eqref{eq:C} we have $H^j(X,\eL^{-m})=0$, for $j=0,1$ and all $m\ges 1$ (cf. \ref{thm:global-gen} and \ref{thm:F-split}), which implies that $H^0(\eO_{D_m})=\kk$ for all $m\ges 0$.
\nit(ii) We apply the Lemma \ref{lm:split}: $H^0(\crl E_X)\to H^0(\crl E_{D_m})$ is surjective if $H^1(\crl E_X\otimes\eL^{-d(m+1)})=0$. This is indeed the case, for large $m$.
\end {m-proof}
The following general splitting principle, corresponding to restrictions to partially ample divisors, is the root of the results obtained in this article.
\begin{m-proposition}\label{prop:h1}
Let $\eL\in\Pic(X)$ be $q$-ample, with $q\les\dim X-2$, and $D\in|d\eL|$ be an effective divisor which is either \emph{reduced} or \emph{irreducible}. Assume that
\begin{equation}\label{eq:h1}
H^1(D,\crl E_D\otimes\eL_D^{-a})=0,\quad\forall a\ges c.
\end{equation}
Then the following properties hold:
\begin{enumerate}
\item
$H^1\bigl(X,\crl E\otimes\eL^{-a}\bigr)=0$, for all $a\ges c$.
\item
If $d\ges c$ and $\crl V_D$ splits, then $\crl V$ splits too.
\end{enumerate}
In any of the cases \eqref{eq:C}, $D$ can be \emph{arbitrary}.
\end{m-proposition}
\begin {m-proof}
(i) Denote
$$a_0:=\max\{a\mid H^1(X,\crl E\otimes\eL^{-a})\neq0\}<\infty.$$
The exact sequence $0\to\eL^{-d}\to\eO_X\to\eO_D\to0$ yields
$$
\ldots\to
H^1\bigl(\crl E\otimes\eL^{-d-a_0}\bigr)
\to
H^1\bigl(\crl E\otimes\eL^{-a_0}\bigr)\to
H^1\bigl(\crl E_D\otimes\eL_D^{-a_0}\bigr)
\to\ldots,
$$
with $-d-a_0\les-(a_0+1)$, so the leftmost term vanishes. If $a_0\ges c$, then the rightmost and the middle terms vanish too. This contradicts the definition of $a_0$, hence $a_0<c$.
\nit (ii) We have $H^0(\eO_D)=\kk$, by \ref{lm:q-split}; since $d\ges c$, $H^0(\crl E)\to H^0(\crl E_D)$ is surjective.
\end {m-proof}
\begin{m-remark}\label{rmk:c}
(i) The uniform $q$-ampleness property \cite[Theorem 6.4]{totaro} implies that there is a linear function $l(r)=\lambda r+\mu$, with $\lambda,\mu$ depending only on $\eL$, such that \ref{prop:h1}(i) holds for all $a\ges l\big({\rm reg}(\crl E)\big)$, where ${\rm reg}(\crl E)$ stands for the regularity of $\crl E$ with respect to a (fixed) ample line bundle $\eO_X(1)$.
\nit(ii) The condition \eqref{eq:h1} involves \emph{only} the restriction of $\crl E$ to $D$, which splits by assumption. This feature is helpful because it is easier to decide the vanishing of the cohomology of line bundles, rather than of vector bundles (\textit{i.e.} \ref{prop:h1}(i)). Also, for $q\les\dim X-3$, the condition \eqref{eq:h1} is indeed fulfilled for $c\gg0$.
\nit(iii) There are two important classes of $q$-ample line bundles: the relatively ample and the pulls-back of ample line bundles with respect to a morphism. We will constantly elaborate on these two situations; the case of a pull-back typically requires stronger hypotheses.
\end{m-remark}
\begin{m-theorem}\label{thm:rel-ample}
Let $\crl V$ be a vector bundle on $X$, $\crl E:=\cEnd(\crl V)$, and
$f:X\to Y$ be a surjective morphism with $Y$ projective,
$$
\dim X-\dim Y\ges 3,\;\eL\in\Pic(X) \text{ $f$-relatively ample}.
$$
Let $D\in|\eL|$ be a \emph{reduced}, \emph{weakly normal} divisor, and assume that $\crl E_D\otimes\eL_D$ is relatively ample with respect to $D\to Y$. (In particular, it suffices $\crl E\otimes\eL$ to be relatively ample.) Then we have the equivalence:
$[\,\crl V\;\mbox{splits} \;\Leftrightarrow\; \crl V_D\;\mbox{splits}\,].$
\end{m-theorem}
The weak normality condition for a divisor (cf. \cite[Proposition 4.1]{le-vi}) is explicit, but it is somewhat technical. However, one can see that the condition is satisfied in the following fairly general situation (a particular case of the $WN1$-property \cite[Definition 3.2]{cgm1}):
\begin{itemize}[leftmargin=*]
\item
$D=D_1+\dots+D_t$ is reduced, and $D$ is normal away from its self-intersections of a single irreducible component or of two different components;
\item
For any point $p\in D$, the local equations (in the analytic topology) of the components of $D$, which are passing through $p$, form a regular sequence in $\hat\eO_{X,p}\cong\kk\{\xi_1,\dots,\xi_{\dim X}\}$. Hence the locus of the points which belong to at least three branches of $D$ have codimension at least two in $D$.
\item
$D$ has \emph{generically} normal crossings: at the generic intersection point between two local (analytic) branches of $D$ there are local (analytic) coordinates $\{\xi_1,\dots,\xi_{\dim X}\}$ in $\hat{\eO}_{X,p}$ such that the germ of $D$ at $p$ is given by $\{\xi_1\xi_2=0\}$.
\end{itemize}
\begin {m-proof}
Let $\ell\in\Pic(D)$ be a direct summand of $\crl E_D$; by hypothesis, $\ell_a:=\ell\otimes\eL^a$ is relatively ample, for all $a\ges 1$. Note that $D$ is Gorenstein, so it satisfies Serre's condition $S_2$. Then the Theorem \ref{thm:global-gen}(iii) implies that $H^1(D,\ell_a^{-1})=0$, that is the condition \eqref{eq:h1} is fulfilled and we may apply the Proposition \ref{prop:h1}.
\end {m-proof}
\section{Splitting along divisors: a `deterministic' approach} \label{sct:q-ample}
\begin{m-definition}\label{def:s-split}
For $s\ges 1$, we say that a scheme $S$ is $s$-split if
\begin{equation}\label{eq:s-split}\tag{$s$-split}
H^j(S,\ell)=0,\text{ for }j=1,\ldots,s,\;\forall\,\ell\in\Pic(S).
\end{equation}
\end{m-definition}
\begin{m-remark}\label{rmk:s-split}
For $s=1,2$ one gets respectively the `splitting' and `Horrocks scheme' notions introduced in \cite{ba}. Examples of projective varieties satisfying \eqref{eq:s-split} are as follows:
\smallskip
\nit(i) arithmetically Cohen-Macaulay varieties $X$---\textit{e.g.} homogeneous spaces, complete intersections in them---with cyclic Picard group (where $s=\dim X-1$), and products of them;
\nit(ii) projective bundles: if $Y$ is $s$-split, and $\euf M_1,\ldots,\euf M_r$, $r\ges s+2$, are line bundles on $Y$, then $X:=\mbb P(\euf M_1\oplus\ldots\oplus\euf M_r)$ is also $s$-split (cf. \cite[Example 4.9]{ba}).
\end{m-remark}
\begin{m-proposition}\label{prop:s-1s}
Let $D$ be an effective $q$-ample divisor on $X$. The following statements hold:
\begin{enumerate}
\item
If $D$ is $s$-split, with $s\les\dim X-(q+1)$, then $X$ is $s$-split.
\item
If $D$ is $s$-split, with $q\les s\les\dim X-(q+1)$, then $X$ is $(s+1)$-split.
\item
If $X$ is $(s+1)$-split, then $D$ is $s$-split in the following cases:
\begin{enumerate}
\item[(a)] $D$ is smooth and $\eO_X(D)$ is $(\dim X-4)$-positive;
\item[(b)] $D$ is arbitrary, relatively ample for a morphism $X\to Y$, with $Y$ projective and $\dim X-\dim Y\ges4$.
\end{enumerate}
\end{enumerate}
\end{m-proposition}
\begin {m-proof}
(i) We consider the exact sequences
$$
0\to\eO_X((k-1)D)\to\eO_X(kD)\to\eO_D(kD)\to 0,\;\forall\,k\in\mbb Z,
$$
and tensor them by $\eL\in\Pic(X)$. We obtain surjective homomorphisms
$H^i(X,\eL((k-1)D))\surj H^i(X,\eL(kD))$, for $i\les s$.
The $q$-ampleness of $D$ implies that $H^i(X,\eL(kD))=0$ for $k\ll0$, which yields $H^i(X,\eL)=0$.
\nit(ii) We should verify only that $H^{s+1}(X,\eL)=0$. The previous exact sequence yields
inclusions $H^{s+1}(X,\eL((k-1)D))\subset H^{s+1}(X,\eL(kD))$, for all $k\in\mbb Z$. Again, the $q$-ampleness of $D$ implies $H^{s+1}(X,\eL(kD))=0$, for $k\gg0$, so $H^{s+1}(X,\eL)=0$.
\nit(iii) In both cases $\Pic(X)\to\Pic(D)$ is an isomorphism, by the Theorem \ref{thm:pic-iso}. Thus, for any $\ell\in\Pic(D)$ there is $\tld\ell\in\Pic(X)$ such that $\tld\ell_D=\ell$. It remains to take the cohomology of the sequence $0\to\tld\ell(-D)\to\tld\ell\to\ell\to0$.
\end {m-proof}
\begin{m-theorem}\label{thm:1}
Let $D$ be an effective divisor on $X$ which is either \emph{reduced} or \emph{irreducible};
in the cases enumerated below, $D$ can be \emph{arbitrary}:
\begin{enumerate}
\item[(a)]
$\eO_X(D)$ is semi-ample;
\item[(b)]
$D$ is relatively ample for a morphism $X\to Y$, with $Y$ projective, $\dim X-\dim Y\ges2$;
\item[(c)]
$X$ is an F-split variety (cf. appendix \ref{sct:frob}).
\end{enumerate}
Assume, moreover, that $D$ is $1$-split and $\eO_X(D)$ is $(\dim X-2)$-ample. Then we have the equivalence:
$
\,[\crl V\;\text{splits}\;\Leftrightarrow\;\crl V_D\;\text{splits}\,].
$
\end{m-theorem}
\begin {m-proof}
Since $\crl V_D$ splits and $D$ is $1$-split, it holds $H^1(\crl E_D\otimes\eL_D^{-a})=0$ for all $a\ges 1$. The conclusion follows from the Proposition \ref{prop:h1}.
\end {m-proof}
The interest in allowing partial ampleness for line bundles, which is considerably weaker than amplitude, is to apply the result for morphisms (\textit{e.g.} fibre bundles).
\begin{m-corollary}\label{cor:relative}
Let $X$ be a smooth, $2$-split, projective variety and $D$ be an effective divisor on it; let $X\srel{f}{\to} Y$ be a morphism. Then the splitting of $\crl V_D$ implies the splitting of $\crl V$ in any of the following cases:
\begin{enumerate}
\item[(a)]
$f$ is smooth, $D=f^{-1}(D_Y)$, with $D_Y\subset Y$ a smooth $(\dim Y-4)$-positive divisor;
\item[(b)]
$D$ is arbitrary, $f$-relatively ample, and $\dim X{-}\dim Y\ges4$.
\end{enumerate}
\end{m-corollary}
\begin {m-proof}
In both situations, the Proposition \ref{prop:s-1s}(iii) implies that $D$ is $1$-split.
\end {m-proof}
Note that (b) above generalizes \cite[Corollary 4.14]{ba} to the relative case. The result seems to be new even in the case when $X$ is a product (cf. \ref{cor:pxq} below).
\begin{m-example}\textbf{(Projective bundles).}\label{expl:proj-bdl}
Let $Y$ be a smooth, projective, $1$-split variety. Consider $\euf M,\euf M_1,\ldots,\euf M_t\in\Pic(Y)$, $t\ges3$, and define
$$X:=\mbb P(\eO_Y\oplus\euf M_1\oplus\ldots\oplus\euf M_t)\srel{f}{\to}Y,\;\;
D:=\mbb P(\euf M_1\oplus\ldots\oplus\euf M_t)\in|\eO_f(1)|.$$
($D$ is $1$-split, by \ref{rmk:s-split}(ii).) For any vector bundle $\crl V$ on $X$, we have: [\,$\crl V$ splits $\Leftrightarrow$ $\crl V_D$ splits\,].
By repeatedly applying this method, one reduces the verification of the splitting of $\crl V$ to a $\mbb P^2$-sub-bundle of $X$ over $Y$.
\end{m-example}
\nit\textbf{(Vector bundles on products of projective spaces and quadrics).}
A splitting criterion for vector bundles on $X_1:=\mbb P^{n_1}\times\ldots\times\mbb P^{n_t}$ is obtained in \cite[Theorem 4.7]{c-mr}. It generalizes Horrocks' criterion, and involves the vanishing of $(n_1+1)\cdot\ldots\cdot(n_t+1)$ cohomology groups.
The result has been extended in \cite[Theorem 2.14, 2.15]{ba-ma} to products $X_1\times X_2$, where $X_1$ is as above and $X_2$ is a product of hyper-quadrics $Q_n\subset\mbb P^{n+1}$. The splitting criterion involves a very large number of cohomological conditions. By applying \ref{cor:relative}, we obtain:
\begin{m-corollary}\label{cor:pxq}
Let $X_1,X_2$ be as above.
\begin{enumerate}
\item
A vector bundle on $X_1$ splits if and only if it splits along a $\mbb P^2\times\ldots\times\mbb P^2\subset X_1$. (This reduces the number of cohomological tests to $3^t$.)
\item
A vector bundle on $X_1\times X_2$ splits if and only if it splits along some $X_1'\times X_2'\subset X_1\times X_2$, where $X_1'$ is a product of projective planes $\mbb P^2$ and $X_2'$ is a product of copies of $Q_3$. (The number of cohomological tests decreases dramatically.)
\end{enumerate}
\end{m-corollary}
\smallskip
\nit\textbf{(Vector bundles on products of minuscule varieties).}
The minuscule homogeneous varieties are the following: the projective spaces, Grassmannians, spinor varieties, quadrics, the Cayley plane, and Freudenthal's variety. In \cite{hal2} it is proved that a vector bundle on a minuscule homogeneous variety $M$, $\dim M\ges 2$, splits if and only if its restriction to the union $M_2\subset M$ of the two-dimensional Schubert subvarieties splits. (It turns out that $M_2$ is either $\mbb P^2$ or a union of two copies of $\mbb P^2$ glued along a $\mbb P^1$.) The proof, which does not fit within the frame of this article, exploits the compatible F-splitting of the Schubert varieties and the properties of the minuscule weights.
Rather \emph{surprisingly}, our approach reduces the problem concerning the splitting of vector bundles on products of minuscule varieties to the problem of splitting on products of projective $2$-planes. To our knowledge, there are no results in this direction, even for vector bundles on products of Grassmannians. Based on the articles \cite{ottv,mals}, one may speculate that, in the latter case, a cohomological splitting criterion should include a large number of tests, indexed by the Schur powers of the universal quotient bundles on the factors.
\begin{m-theorem}\label{thm:minuscule}
Let $M^{(j)}$, $j=1,\dots,t$, be minuscule homogeneous varieties, $\dim M^{(j)}\ges 2$, and
$X:=M^{(1)}\times\dots\times M^{(t)}.$
A vector bundle $\crl V$ on $X$ splits if and only if its restriction to $M^{(1)}_2\times\dots\times M^{(t)}_2$ splits, where each $M^{(j)}_2\subset M^{(j)}$ stands for the union of the $2$-dimensional Schubert subvarieties.
\end{m-theorem}
\begin {m-proof}
The proof is by induction on $t+\rk\crl V$. For $t=1$, see \cite{hal2}. Let us prove the statement for $\tld X=X\times M$, with $X$ as above and $M$ minuscule: we assume that $\crl V_{X_2\times M_2}$ splits, where $X_2:=M^{(1)}_2\times\dots\times M^{(t)}_2\subset X$. The proof consists of two steps.
\smallskip
\nit\textit{Claim 1}\quad $\crl V_{X\times M_2}$ splits.
We prove this step by induction on $\rk(\crl V)$.
Recall that $\mbb Z^{t+1}\cong\Pic(\tld X)\srel{\cong}{\to}\Pic(X_2\times M_2)$, so the line bundles on $\tld X$ are of the form
$
\eO_{\tld X}(\unl{\tld\alpha})=
\eO_{M^{(1)}}(\alpha_1)\boxtimes\dots\boxtimes\eO_{M^{(t)}}(\alpha_t)\boxtimes\eO_M(k)=
\eO_X(\unl\alpha)\boxtimes\eO_M(k).
$
We deduce that
\begin{equation}\label{eq:xm2}
\crl V_{X_2\times M_2}\cong
\uset{(\unl\alpha,k)\in\mbb Z^{t+1}}{\bigoplus} \eO_X(\unl\alpha)\boxtimes\eO_{M_2}(k)^{d_{\unl\alpha,k}}.
\end{equation}
Let us consider the diagram
$$
\xymatrix@R=1.75em@C=3em{
X_2\times M_2\,\ar@{^(->}[r]\ar[d]^-f&X\times M_2\ar[d]^-f
\\
M_2\ar@{=}[r]&M_2.
}
$$
The induction hypothesis implies that $\crl V$ splits on the fibres of $f$, so
$$
\crl V_{X\times\{z\}}\cong\uset{\unl\alpha\in\mbb Z^t}{\bigoplus}
\eO_X(\unl\alpha)^{d_{\unl\alpha}(z)},\; \forall\,z\in M_2.
$$
Actually, the multiplicities $d_{\unl\alpha}(z)$ are independent of $z\in M_2$. By restricting to $X_2\times\{z\}$ and by using \eqref{eq:xm2}, we deduce:
$$
d_{\unl\alpha}:=\uset{k\in\mbb Z}{\sum}\;d_{\unl\alpha,k}
=d_{\unl\alpha}(z),\;\forall\,z\in M_2.
$$
Let $\unl a\in\mbb Z^t$ be a maximal element of $\{\unl\alpha\in\mbb Z^t\mid d_{\unl\alpha}\neq0\}$, for the lexicographic order. Then $f_*\big(\eO_X(-\unl a)\otimes\crl V\big)$ is locally free on $M_2$ of rank $d_{\unl a}$ and \eqref{eq:xm2} yields:
$$
\crl R:=f_*\big(\eO_X(-\unl a)\otimes\crl V\big)\cong
\uset{k\in\mbb Z}{\bigoplus}\eO_{M_2}(k)^{d_{\unl a,k}}.
$$
(For this, observe that
$H^0\big(X,\eO_X(\unl\alpha-\unl a)\big), H^0\big(X_2,\eO_X(\unl\alpha-\unl a)\big)\neq 0$
if and only if all the components of $\unl\alpha-\unl a$ are positive; the only such $\unl\alpha$ is $\unl a$ itself.) It follows that we have the exact sequence
$\;0\to\eO_X(\unl a)\boxtimes\crl R\to\crl V\to\crl V'\to0.$
The first arrow is pointwise injective, so the quotient $\crl V'$ is locally free on $X\times M_2$, its restriction to $X_2\times M_2$ splits, and $\rk(\crl V')<\rk(\crl V)$. Hence, by the induction hypothesis, $\crl V'$ splits. Finally, we deduce that
$$
\crl V\in\Ext^1(\crl V',\eO_X(\unl a)\boxtimes\crl R)=H^1\big(X\times M_2,\cHom(\crl V',\eO_X(\unl a)\boxtimes\crl R)\big)=0,
$$
because $X\times M_2$ is $1$-split and $\cHom(\crl V',\eO_X(\unl a)\boxtimes\crl R)$ is a direct sum of line bundles. We conclude by recurrence that $\crl V_{X\times M_2}$ splits.
\smallskip
\nit\textit{Claim 2}\quad $\crl V_{X\times M}$ splits.
We denote by $M_d$ the union of all the $d$-dimensional Schubert subvarieties $S_d\subset M$. For any $(d+1)$-dimensional Schubert variety $S_{d+1}\subset M$, the intersection $\del S_{d+1}:=M_d\cap S_{d+1}$ is reduced and it is the union of the $d$-dimensional Schubert subvarieties of $S_{d+1}$; usually, it's called the boundary of $S_{d+1}$.
With this notation, for $d\ges2$, the following properties hold (cf. \cite{hal2} and the references therein):
\begin{itemize}[leftmargin=*]
\item
$\mbb Z\cdot\eO_M(1)=\Pic(M)\to\Pic(S_d)$ is an isomorphism;
\item
$\eO_M(1)\otimes \eO_{S_{d+1}}=\eO_{S_{d+1}}(\del S_{d+1});$
\item
$S_d, \del S_{d+1}$ are $1$-split.
\end{itemize}
Thus, for any Schubert variety $S_{d+1}\subset M$, the divisor $X\times\del S_{d+1}\subset X\times S_{d+1}$ is $1$-split and $\dim X$-positive. We deduce the implications:
$$
\crl V_{X\times M_d}\mbox{ splits }
\;\srel{\del S_{d+1}\subset M_d}{\Longrightarrow}\;
\crl V_{X\times\del S_{d+1}}\mbox{ splits }
\;\ouset{\ref{thm:1}}{\text{Thm.}}{\Longrightarrow}\;
\crl V_{X\times S_{d+1}}\mbox{ splits.}
$$
Clearly, for $S'_{d+1}, S''_{d+1}\subset M$, the splittings of
$\crl V_{X\times S'_{d+1}},\crl V_{X\times S''_{d+1}}$ coincide along the (reduced) intersection $X\times (S'_{d+1}\cap S''_{d+1})\subset X\times M_d$, so one gets a splitting of $\crl V_{X\times M_{d+1}}$. By repeating the argument, we deduce that $\crl V_{X\times M}$ splits.
\end {m-proof}
\begin{m-example}\label{expl:mr+bm}
Let $X=X^{(1)}\times\dots\times X^{(t)}$ be a product of Fano varieties of dimension at least four, with $\Pic(X^{(j)})\cong\mbb Z$ for all $j$. By applying the Theorem \ref{thm:1}, one reduces the problem of splitting of a vector bundle on $X$ to $S^{(1)}_3\times\dots\times S^{(t)}_3$, where each $S^{(j)}_3\subset X^{(j)}$ is an irreducible, $3$-dimensional complete intersection.
\end{m-example}
\section{Splitting along divisors: a `probabilistic' approach}
\label{sct:global-gener}
In this section we obtain splitting criteria for vector bundles by restricting them to zero loci of \textit{generic} sections of \textit{globally generated}, partially positive line bundles. The global generation allows to replace the {\rm($2$-split)} by the weaker {\rm($1$-split)} condition, but we will have to consider very general test divisors instead of arbitrary ones. This explains the `probabilistic' attribute used in the title.
Note that, if $\eL$ is a $q$-ample line bundle on $X$ such that $\eL^d$ is globally generated for some $d\ges 1$, then $\eL$ is $q$-positive and the fibres of the morphism $f:X\to|\eL^d|$ are at most $q$-dimensional (cf. \cite[Theorem 1.4]{mats}), thus $\kappa(\eL)\ges\dim({\rm Image}(f))\ges\dim X-q$. Henceforth we replace $\eL^d$ by $\eL$.
Let the situation be as above. We start with general considerations: the equations defining $X,\eL,\crl V$ involve finitely many coefficients in $\kk$. By adjoining them to $\mbb Q$, we obtain a field extension $\mbb Q\hra\Bbbk$ of finite type (which depends on $\crl V$), so we may assume that $\Bbbk\hra\kk$; its algebraic closure is a \emph{countable}. After replacing $\Bbbk$ by $\bar{\Bbbk}$, we may assume that $X,\eL,\crl V$ are defined over a countable, algebraically closed subfield $\Bbbk$ of $\kk$; we denote these objects by $X_\Bbbk,\eL_\Bbbk,\crl V_\Bbbk$.
The sheaf $\eG{:=}\,\Ker\big( H^0(X,\eL)\otimes\eO_X {\to}\,\eL\big)$ is locally free and the incidence variety
$$
\cal D{:=}\{([s\,],x)\mid s(x)=0\}\subset|\eL|\times X
$$
is naturally isomorphic to the projective bundle $\mbb P(\eG)$ over $X$. The projections of $\cal D$ onto $|\eL|, X$ are denoted by $\pi,\rho$:
$$
\xymatrix@R=.75em{
&\;\cal D\;\ar[dl]^-\pi\ar[dr]_-\rho\ar@{^(->}[rr]&&|\eL|\times X
\\
|\eL|&&\,X.
}
$$
For any open $S\subset|\eL|$, let $\cal D_S:=\pi^{-1}(S)$; for $s\in|\eL|$, let $D_s:=\dvs(s)\subset X$.
All these objects are defined over $\Bbbk$, and they are denoted by $\eL_\Bbbk, \cal D_\Bbbk, \pi_\Bbbk, \rho_\Bbbk$. Let $K_{\kk}:=\kk(|\eL|)$ and $K_{\Bbbk}:=\Bbbk(|\eL_\Bbbk|)$ be the function fields of the projective spaces $|\eL|$, $|\eL_\Bbbk|$, respectively.
\begin{m-definition}\label{def:very-gen}
We say that a property holds for a \emph{very general} point of some parameter space, if it holds on the complement of countably many proper subvarieties of that parameter space. In our case, we are interested in the splitting of $\crl V_{D_s}$, for $s\in|\eL|$ very general.
\end{m-definition}
\begin{m-lemma}\label{lm:very-gen}
\begin{enumerate}
\item
If $\crl V_{D_s}$ splits for a very general $s\in|\eL|$, then $(\rho^*\crl V)\otimes\bar{K}_\kk$ splits.
\item If $(\rho^*\crl V)\otimes\bar{K}_\kk$ splits, then there is an \emph{analytic} open ball $\ball\subset|\eL|$ such that $D_s$ is smooth, for all $s\in\ball$, and ${(\rho^*\crl V)}_\ball$ splits over $\cal D_\ball$.
\end{enumerate}
\end{m-lemma}
\begin {m-proof}
(i) Let $\tau:|\eL|\to |\eL_\Bbbk|$ be the trace morphism. Since $\Bbbk$ is countable and $\kk$ is not, $\tau(s)$ is the generic point of $|\eL_\Bbbk|$, for $s\in|\eL|$ very general. Hence $\Bbbk(\tau(s))=\Bbbk(|\eL_\Bbbk|)=K_\Bbbk$ and we obtain the Cartesian diagram:
$$
\xymatrix@R=1.75em@C=3em{
D_s\ar[r]\ar[d]
&
\cal D_{\bar{K}_\Bbbk}
\ar[d]\ar[r]
&
\cal D_{\Bbbk}\ar[d]
\\
\Spec(\kk)\ar[r]
&
\Spec(\bar{K}_\Bbbk)\ar[r]
&
\,|\eL_\Bbbk|.
}
$$
For varieties defined over algebraically closed fields ($\kk$ and $\bar{K}_\Bbbk$ in our case), the property of a vector bundle to be split commutes with base change. Then $\crl V_{\bar{K}_\Bbbk}$ splits on $\cal D_{\bar{K}_\Bbbk}$, so the same holds for $\crl V_{\bar{K}_\kk}$.
\nit{(ii)}
We note that $\crl V_{\cal D_{\bar{K}_\kk}}$ actually splits over an intermediate field $K_\kk\hra K'\hra\bar{K}_\kk$ finitely generated (and also algebraic) over $K_\kk$. Thus there is an open affine $S\subset|\eL|$, an affine variety $S'$ over $\kk$, and a \emph{finite morphism} $S'\srel{\si}{\to} S$ such that the direct summands of $\crl V_{\cal D_{\bar{K}_\kk}}$ are defined over $\kk[S']$; thus ${(\rho^*\crl V)}_{S'}$ splits. For $S$ sufficiently small, Bertini's theorem implies that $D_s$ is smooth, for all $s\in S$. Finally, there are open balls $\ball'\subset S'$ and $\ball\subset S$ such that $\si:\ball'\to\ball$ is an {\em analytic} isomorphism. Then the splitting of ${(\rho^*\crl V)}_{\ball'}$ descends to ${(\rho^*\crl V)}_{\ball}$ on $\cal D_{\ball}$.
\end {m-proof}
\begin{m-remark}\label{rmk:very-gen}
The previous lemma precises the meaning of a very general point $s\in|\eL|$: its coordinates should be algebraically independent over the definition field of $X,\eL,\crl V$. In particular, the notion of a very general point depends on $\crl V$ itself, actually on its field of definition.
Often one wishes to have statements which involve \emph{general} points, rather than very general ones. The splitting of $\crl V_{D_s}$ for general $s\in|\eL|$ means, by definition, that $(\rho^*\crl V)\otimes K_\kk$ splits. This condition is \emph{more restrictive} than the splitting of $(\rho^*\crl V)\otimes \bar K_\kk$.
\end{m-remark}
\begin{m-theorem}\label{thm:2}
Suppose that $\eL\in\Pic(X)$ is globally generated, $(\dim X-4)$-positive, and $D\in|\eL|$ is very general (thus smooth). If $X$ is $1$-split, then holds: $[\,\crl V$ splits $\Leftrightarrow$ $\crl V_D$ splits$\,]$.
\end{m-theorem}
The interest in this result is that it allows to test the splitting of vector bundles along divisors which are not $1$-split; this situation arises especially in low dimensions. Otherwise, of course, one applies the `deterministic' Theorem \ref{thm:1}.
\begin {m-proof}
Let $\ball$ be as in the Lemma \ref{lm:very-gen}. By \cite[Proposition 5.1]{ottem}, the cohomological dimension $\cd(X\sm D_s)\les \dim X-4$, for all $s\in\ball$, which implies that, for all $o,s,t\in\ball$, the intersections $D_{st}:=D_s\cap D_t$ and $D_{ost}:=D_o\cap D_s\cap D_t$ are non-empty and connected (cf. \cite[Ch. III, Corollary 3.9]{hart-as}). Note that the intersections are generically transverse, because $\eL$ is globally generated.
\smallskip
\nit\textit{Claim~1}\quad
Let $s,t\in\ball$ such that $D_s,D_t$ intersect transversally. Then holds:
$$
\begin{array}{cl}
\res^X_{D_s}:\Pic(X)\to\Pic(D_s)&\mbox{is an isomorphism};
\\[1ex]
\res^{D_s}_{D_{st}}:\Pic(D_s)\to\Pic(D_{st})&\mbox{is injective}.
\end{array}
$$
The first statement is proved in the Theorem \ref{thm:pic-iso} and the second in the Proposition \ref{prop:pic-inj}.
\smallskip
\nit\textit{Claim~2}\quad
$\rho^*:\Pic(X)\to\Pic(\cal D_\ball)$ is an isomorphism.
Indeed, fix $o\in\ball$ and consider the diagram:
$$\xymatrix@R=.75em@C=4em{
&\Pic(D_s)\ar[dr]^-{\res^{D_s}_{D_{os}}}&
\\
\Pic(X)\ar[ur]^-{\res^X_{D_s}}
\ar[dr]_-{\res^X_{D_o}}\ar[rr]|{\;\res^X_{D_{os}}\;}
&&
\Pic(D_{os}).
\\
&\Pic(D_o)\ar[ur]_-{\res^{D_o}_{D_{os}}}&
}
$$
The composition $\Pic(X)\,{\srel{\;\rho^*}{\to}}\Pic(\cal D_\ball){\srel{\res_{D_o}}{\to}}\Pic(D_o)$ is bijective, so $\rho^*$ is injective. For the surjectivity, take $\ell\,{\in}\Pic(\cal D_\ball)$. If $\ell_{D_o}{\cong}\,\eO_{D_o}$, then $\ell_{D_s}\in\Pic^0(D_s)$, for all $s\in\ball$, so
$$\{s\in\ball\mid\ell_{D_s}\not\cong\eO_{D_s}\}=\{s\in\ball\mid h^0(\ell_{D_s})=0\}$$
is open in $\ball$, hence
$S:=\{s\in\ball\mid\ell_{D_s}\cong\eO_{D_s}\}$
is closed.
On the other hand, the diagram above implies, by taking the restrictions to $D_{os}$, that $S$ contains all $s$ such that $D_s$ intersects $D_o$ transversally; thus $S$ is dense, so $S=\ball$. It follows that $\ell\cong\eO_{\cal D_\ball}$. For an arbitrary $\ell\in\Pic(\cal D_\ball)$, let $\eL\in\Pic(X)$ such that $\ell_{D_o}\cong\eL_{D_o}$, so ${\big((\rho^*\eL^{-1})\otimes\ell\big)}_{D_o}$ is trivial. This concludes the proof of the Claim~2.
\smallskip
Since $\crl V_\ball$ splits, we deduce that
${(\rho^*\crl V)}_\ball\,{\cong}\,\rho^*\big(\underset{j\in J}{\bigoplus}\eL_j^{\oplus d_j}\big)$,
with $\eL_j\in\Pic(X)$ pairwise non-isomor\-phic. We consider the following partial order on line bundles:
$$
\eL\prec\euf M\;\Leftrightarrow\;\eL\not\cong\euf M
\text{ and } H^0(\eL^{-1}\otimes\euf M)\neq0.
$$
For $s\in\ball$, let $J_{s,\mx}\subset J$ be the set of maximal elements for $\prec$ on $\Pic(D_s)\cong\Pic(X)$. By semi-continuity, the set $\{t\in\ball\mid J_{s,\mx}\subset J_{t,\mx}\}$ is open. Thus, after shrinking $\ball$, we may assume that $J_{s,\mx}\subset J$ is independent of $s$; we denote it by $J_\mx$.
The maximality property implies that there is a natural, pointwise injective homomorphism
\begin{equation}\label{eq:h-inj}
h:\underset{\mu\in J_\mx}{\bigoplus}
\rho^*\eL_\mu\otimes
\underbrace{\pi^*\pi_*\big(\rho^*(\eL_\mu^{-1}\otimes\crl V)\big)}
_{\srel{\circledast}{\cong} \;\; \eO_{\cal D_\ball}^{\oplus d_\mu}}
\to{\big(\rho^*\crl V\big)}_{\cal D_\ball}.
\end{equation}
\nit\textit{Claim~3}\quad \begin{minipage}[t]{.8\textwidth}
$h$ descends to $X$, after a suitable base change in $\eO_{\cal D_\ball}^{\oplus d_\mu}$ by an analytic map
$\beta:\ball\to\kern-1ex\underset{\mu\in J_\mx}{\prod}\kern-1.5ex\Gl(d_\mu)$.\end{minipage}
Indeed, we fix $o\in\ball$ and, for each $\mu\in J_\mx$, we fix a basis in
$H^0(D_o,\eL_\mu^{-1}\crl V)\cong\kk^{d_\mu}$. (Bases are represented as square matrices whose columns are the vectors of the basis.) For any $s\in\ball$, $D_{os}$ is non-empty and connected, so there is a \emph{unique} basis in $H^0(D_s,\eL_\mu^{-1}\crl V)\cong\kk^{d_\mu}$ whose restriction to $D_{os}$ coincides with the restriction of the basis along $D_o$. (As $s\in\ball$ varies, the transition matrices from the trivialization $\circledast$ in \eqref{eq:h-inj} to these new bases yields the map $\beta$.)
We observe that, after this reparameterization, $h$ descends to the open set $\cU:=\rho(\cal D_{\ball})\subset X$. Indeed, define
\begin{equation}\label{eq:bar-h}
\bar h:
\big(\underset{m\in J_\mx}{\mbox{$\bigoplus$}}\kern-1ex\eL_\mu^{\oplus d_\mu}\big)\otimes\eO_\cU
\to\crl V_\cU,\;\bar h(x):=h_s(x)\text{ for some }s\in\ball\text{ such that }x\in D_s.
\end{equation}
In order to prove that $\bar h(x)$ is independent of $s\in\ball$, we must show that, for any $s,t\in\ball$, the restrictions to $D_{st}$ of the new bases in $H^0(D_s,\eL_\mu^{-1}\crl V),H^0(D_t,\eL_\mu^{-1}\crl V)$ coincide: it is enough to check this on the triple intersection $D_{ost}=D_o\cap D_{st}$ (which is non-empty, connected), where both bases are induced from $D_o$.
The homomorphism \eqref{eq:bar-h} yields the extension of locally free sheaves on $\cU$:
$$
0\to
\big(\underset{m\in J_\mx}{\bigoplus}\kern-1ex\eL_\mu^{\oplus d_\mu}\big)\otimes\eO_\cU
\to\crl V_\cU\to\crl W_\cU\to0,
\quad
\rho^*(\crl W_\cU)\cong\kern-.5ex\rho^*\Big(
\underset{j\in J\sm J_\mx}{\bigoplus}\kern-1.5ex\eL_\mu^{\oplus d_\mu}
\otimes\eO_\cU\Big).
$$
The homomorphism on the left is pointwise injective. Recursively, we deduce that $\crl V_\cU$ is obtained as a successive extension of the line bundles $\eL_j\otimes\eO_\cU$, $j\in J$. (Note that, in the gluing process, we did not use that $\crl V$ is defined on all $X$; we used only its restriction to $\cU$.)
Since $\cU$ is an analytic neighbourhood of $D_o$, one gets induced extensions on the thickenings ${(D_{o})}_m$, $m\ges 0$, (cf. Definition \ref{def:D-thick}). But $X$ is $1$-split and $D_o$ is $(\dim X-4)$-ample, so
\begin{equation}\label{eq:ext1}
0=\Ext^1(\eL,\euf M)\to \Ext^1\big( \eL_{{(D_{o})}_m}, \euf M_{{(D_{o})}_m}\big)
\end{equation}
is an isomorphism, for all $\eL,\euf M\in\Pic(X)$, $m\gg0$. It follows that $\crl V_{{(D_o)}_m}$ splits, for $m\gg0$. By applying the Lemma \ref{lm:q-split}, we deduce that $\crl V$ splits on $X$. (Here it is necessary to have $\crl V$ defined on $X$.)
\end {m-proof}
\begin{m-example}\label{expl:p2p2}
We have seen in \ref{expl:proj-bdl} that the splitting of a vector bundle on a projective bundle over a $1$-split variety can be verified along an arbitrary $\mbb P^2$-sub-bundle. At this stage, the reduction process given by the Theorem \ref{thm:1} stops. In this example we show that the Theorem \ref{thm:2} allows to decrease further the dimension of the test subvarieties.
Let $(S,\eA)$ be an arithmetically Cohen-Macaulay surface, with $\eA$ ample and $\Pic(S)=\mbb Z\eA$. (A necessary and sufficient cohomological condition for the splitting of vector bundles on such surfaces has been obtained in \cite{ha+ta}.) The four-fold
$$
X:=\mbb P(\eO_S\oplus\eA^{-m}\oplus\eA^{-m-n})\srel{f}{\to} S,\;\;m,n\ges0,
$$
is $1$-split. The line bundle $\eL:=\eO_f(1)\otimes f^*\eA$ is ample on $X$, and the general $D\in|\eL|$ is a smooth $\mbb P^1$-fibre bundle over $S$. In particular, $D$ is not $1$-split, so the theorem \ref{thm:1} can not be applied. However, the `probabilistic' Theorem \ref{thm:2} still applies: a vector bundle $\crl V$ on $X$ splits if and only if it does split on a very general $D$.
\end{m-example}
\section{Triviality criteria}\label{sct:triv}
Finally, in this section we restrict our discussion to the case of the trivializable vector bundles. The motivation is, first, that the general `effective splitting criterion' \ref{prop:h1} is not explicit enough, especially for partially ample line bundles which are pulls-back (cf. Remark \ref{rmk:c}, Theorem \ref{thm:rel-ample}). Second, it is desirable to remove the $1$-, $2$-split conditions which appear throughout the sections \ref{sct:q-ample}, \ref{sct:global-gener}, and which are imposed precisely to ensure the vanishing \eqref{eq:h1}.
Unfortunately, the Kodaira vanishing does not hold for $q$-ample line bundles. Thus, to obtain effective results in this situation, one must find appropriate conditions which imply the Kodaira vanishing. These lines of thought lead to the triviality criteria below.
\begin{m-lemma}\label{lm:pic-inj}
Assume that $\eL\in\Pic(X)$ is $(\dim X-2)$-ample and satisfies $H^i(X,\eL^{-a})=0$, for all $a\ges 1$ and $i=0,1,2$. For any vector bundle $\crl V$ on $X$ and $D\in|\eL|$, one has the equivalence:
$\;[\,\crl V\cong\eO_X^{\oplus r}\;\Leftrightarrow\;\crl V_D\cong\eO_D^{\oplus r}\,].$
\end{m-lemma}
\begin {m-proof}
The hypothesis implies that $H^i(\eL_D^{-a})=0$, for all $a\ges 1$, $i=0,1$, and $H^0(\eO_{D_a})=\kk$, for $a\ges0$, so $H^0(\crl E)\to H^0(\crl E_D)=\End(\kk^r)$ is an isomorphism, by the Proposition \ref{prop:h1}. Hence $\crl V$ splits, actually $\crl V\cong\euf M^{\oplus r}$ for some $\euf M\in\Pic(X)$. As $\crl V_D=\eO_D^{\oplus r}$, we deduce that both $\euf M_D$ and $\euf M_D^{-1}$ admit non-trivial sections, so $\euf M_D\cong\eO_D$. According to the Proposition \ref{prop:pic-inj}, $\Pic(X)\to\Pic(D)$ is injective, so $\euf M$ is trivial.
\end {m-proof}
\begin{m-theorem}\label{thm:q-global-gener}
Consider a vector bundle $\crl V$ on $X$, $\eL\in\Pic(X)$, and $D\in|\eL|$. In any of the following cases, we have:
$\;[\,\crl V\cong\eO_X^{\oplus r}\;\Leftrightarrow\;\crl V_D\cong\eO_D^{\oplus r}\,].$
\begin{enumerate}
\item[(a)]
$\eL\in\Pic(X)$ is semi-ample and $(\dim X-3)$-ample;
\item[(b)]
$\eL$ is relatively ample for a morphism $X\srel{f}{\to}Y$, with $\dim(X)-\dim(Y)\ges3$.
\end{enumerate}
\end{m-theorem}
\begin {m-proof}
In both cases, the Theorem \ref{thm:global-gen} implies $H^i(X,\eL^{-a})=0$ for $a\ges 1$, $i=0,1,2$, so we can apply the Lemma \ref{lm:pic-inj}.
\end {m-proof}
\subsection{The case of Frobenius split (F-split) varieties}\label{ssct:frob}
These objects are ubiquitous, especially in representation theory. Examples of F-split varieties (defined in characteristic zero) include Fano varieties (cf. \cite[Exercise 1.6E(5)]{brion-kumar}), spherical varieties, in particular projective homogeneous varieties and toric varieties (cf. \cite{br-in}, \cite[Section 31]{tim}). The notions and properties which are relevant for us are summarized in the appendix \ref{sct:frob}.
\begin{m-theorem}\label{thm:F-q-split}
Let $D$ be a $(\dim X-3)$-ample, effective divisor, which is F-split. Then holds:
$$
\crl V\cong\eO_X^{\oplus r}
\;\Leftrightarrow\;
\crl V_D\cong\eO_D^{\oplus r}.
$$
\end{m-theorem}
The F-splitting allows to handle more `exotic' situations. Many examples arise from varieties $X$ which are compatibly split with respect to a divisor $D$.
\begin {m-proof}
Since $\eO_D(D)$ is $(\dim D-2)$-ample, the Theorem \ref{thm:F-split}
implies $H^1(D,\eO_D(-aD))=0$, for all $a\ges 1$. The conclusion follows from the Proposition \ref{prop:h1}.
\end {m-proof}
\begin{m-corollary}\label{cor:f-split}
Let $X$ be a smooth projective variety whose anti-canonical line bundle $\omega_X^{-1}$ is $(\dim X-3)$-ample. Assume that $X$ is F-split by $\si\in H^0(\omega_X^{-1})$, and denote $D:=\dvs(\si)$. Then holds:
$[\,\crl V\cong\eO_X^{\oplus r}\;\Leftrightarrow\;\crl V_D\cong\eO_D^{\oplus r}\,].$
\\ The criterion applies, in particular, in the following cases:
\begin{enumerate}
\item[(a)] $X$ is a Fano variety of dimension at least three;
\item[(b)] $X$ is a spherical variety whose anti-canonical bundle is $(\dim X-3)$-ample.
\end{enumerate}
\end{m-corollary}
\begin {m-proof}
The hypothesis implies that $D$ is F-split, compatibly with the splitting defined by $\si$.
Fano and spherical varieties are Frobenius split, compatibly with suitable anti-canonical divisors (cf. \cite[1.6.E(5), pp. 56]{brion-kumar} and \cite[Theorem 1]{br-in}, respectively).
\end {m-proof}
\subsection{The case of toric varieties}\label{toric}
A non-trivial application of the ideas developed inhere arises when $X:=X_\Si$ is the smooth projective toric variety defined by the regular fan $\Si$.
\begin{m-remark}\label{rmk:toric}
{\rm(i)}
$X$ is F-split, compatibly with the invariant divisors $D_\rho, \rho\in\Si(1)$, and their intersections (cf. \cite[Exercise 1.3E(6)]{brion-kumar}).
\\{\rm(ii)} Let
\begin{equation}\label{eq:dta}
\Delta:=\underset{\rho\in\Si(1)}{\mbox{$\sum$}}{D_\rho}
\end{equation}
be the torus-invariant, anti-canonical divisor. Its complement is $X\sm\Delta\cong(\mbb C ^*)^{\dim X}$, so the cohomological dimension equals $\cd(X\sm\Delta)=0$.
\end{m-remark}
\begin{m-theorem}\label{thm:split-toric}
Let $\crl V$ be an arbitrary vector bundle on the smooth toric variety $X$. The following statements hold:
\begin{enumerate}
\item
Let $\disp\hat X:=\varprojlim_m\Delta_m$ be the formal completion of $X$ along $\Delta$. If $\dim X\ges 2$, then one has the equivalence:
$
\;[\,\crl V\mbox{ splits}\;\Leftrightarrow\;\crl V\otimes\eO_{\hat X}\mbox{ splits}\,].
$
\item
Assume that
$\dim X,\;\codim\big(\bloc(\omega_{X}^{-1})\,\big)
\ges 3.$
Then we have:
\begin{enumerate}
\item[(a)]
$[\,\crl V\mbox{ splits}\;\Leftrightarrow\;\crl V_{\Delta_m}\mbox{ splits}\,]$, for $m\gg0$.
(See \ref{rmk:c}(i) for a lower bound on $m$.)
\item[(b)]
$[\,\crl V\mbox{ is trivial}\;\Leftrightarrow\;\crl V_{\Delta}\mbox{ is trivial}\,]$.
\end{enumerate}
\end{enumerate}
\end{m-theorem}
\begin{m-proof}
\nit(i) See the Proposition \ref{prop:formal}.
\nit(ii) The Theorem \ref{thm:q-ample} implies that $\omega_{X}^{-1}$ is $(\dim X-3)$-ample. Now the conclusion follows from the Proposition \ref{prop:h1} and the Corollary \ref{cor:f-split}, respectively.
\end{m-proof}
\begin{m-remark}
\nit{\rm(i)}
One may wonder if it is possible to have a splitting criterion for toric varieties which involves an \emph{irreducible}, torus-invariant, $(\dim X-2)$-ample divisor. In general, the answer is ``no''; \emph{reducible} divisors are necessary for the following reason. If $D$ is such an irreducible divisor, then ${\rm cd}(X\sm D)\les\dim X-2$, so $D$ must intersect all the other torus-invariant divisors. Hence $\Si$ has the following property: if $\xi_D\in\Si(1)$ defines $D$, then $\xi_D,\xi$ span a cone of $\Si$, for all $\xi\in\Si(1)\sm\{\xi_D\}$. This condition is clearly not satisfied in general.
\smallskip
\nit{\rm(ii)}
It is rather surprising, but the issue concerning the bare existence of non-trivial vector bundles on toric varieties is not settled yet, in general (cf. \cite{gharib+karu}).
\end{m-remark}
|
\section{INTRODUCTION}
\label{sec:introduction}
Agglomerative hierarchical clustering, which is one of the most popular algorithms in cluster analysis,
builds a binary tree representing the cluster structure of a dataset~\citep{DudaRO2001book}.
Given a dataset and a dissimilarity measure between clusters,
agglomerative hierarchical clustering starts from leaf nodes corresponding to individual data points
and successively merges pairs of nodes with smallest dissimilarities to complete a binary tree.
Bayesian hierarchical clustering (BHC) \citep{HellerKA2005icml} is a probabilistic alternative of agglomerative
hierarchical clustering. BHC defines a generative model on binary trees and compute the probability that
nodes are merged under that generative model to evaluate the (dis)similarity between the nodes.
Since this (dis)similarity is written as a probability, one can naturally decide a level, where to stop
merging according to this probability. Hence, unlike traditional agglomerative clustering algorithms,
BHC has a flexibility to infer a proper number of clusters for given data. The source of this
flexibility is Dirichlet process mixtures (DPM) \citep{FergusonTS73as,AntoniakCE74aos} used to define the generative model of binary trees.
BHC was shown to provide a tight lower bound on the marginal likelihood of DPM~\citep{HellerKA2005icml,WallachHM2010aistats}
and to be an alternative posterior inference algorithm for DPM.
However, when evaluating the dissimilarity between nodes, one has to repeatedly compute the marginal likelihood
of clusters and careful tuning of hyperparameters are required.
In this paper, we study BHC when the underlying distributions are conjugate exponential families.
Our contributions is twofold.
First, we derive a non-probabilistic relaxation of BHC, referred to as RBHC,
by performing {\em small variance asymptotics,} i.e., letting the variance of the underlying distribution in the model go to zero.
To this end, we use the technique inspired by the recent work \citep{KulisB2012icml,JiangK2012nips},
where the Gibbs sampling algorithm for DPM with conjugate exponential family was shown to approach
a $k$-means-like hard clustering algorithm in the small variance limit.
The dissimilarity measure in RBHC is of a simpler form, compared to the one in the original BHC.
It does not require careful tuning of hyperparameters, and yet has the flexibility of the original BHC to infer
a proper number of the clusters in data.
It turns out to be equivalent to the dissimilarity proposed in \citep{TelgarskyM2012icml},
which was derived in different perspective, minimizing a cost function involving Bregman information~\citep{BanerjeeA2005jmlr}.
Second, we study the {\em reducibility}~\citep{BruynoogheM78} of the dissimilarity measure in RBHC.
If the dissimilarity is reducible, one can use \emph{nearest-neighbor chain algorithm}~\citep{BruynoogheM78} to
build a binary tree with much smaller complexities, compared to the greedy algorithm.
The nearest neighbor chain algorithm builds a binary tree of $n$ data points in $O(n^2)$ time and $O(n)$ space,
while the greedy algorithm does in $O(n^2\log n)$ time and $O(n^2)$ space.
We argue is that even though we cannot guarantee that the dissimilarity in RBHC is always reducible,
it satisfies the reducibility in many cases, so it is fine to use the nearest-neighbor chain algorithm in practice to speed up
building a binary tree using the RBHC.
We also present the conditions where the dissimilarity in RBHC is more likely to be reducible.
\section{BACKGROUND}
\label{sec:background}
We briefly review agglomerative hierarchical clustering, Bayesian hierarchical clustering, and Bregman clustering,
on which we base the development of our clustering algorithm RBHC.
Let $\bX = \{\bx_i\}_{i=1}^n$ be a set of $n$ data points.
Denote by $[n] = \{1,\dots,n\}$ a set of indices.
A {\em partition} $\pi_n$ of $[n]$ is a set of disjoint nonempty subsets of $[n]$ whose union is $[n]$.
The set of all possible partitions of $[n]$ is denoted by $\Pi_{n}$
For instance, in the case of $[5]=\{1,2,3,4,5\}$, an exemplary random partition that consists of three clusters is
$\pi_5=\bigl\{ \{1\}, \{2,4\}, \{3,5\} \bigr\}$; its members are indexed by $c \in \pi_5$.
Data points in cluster $c$ is denoted by $\bX_c \defeq \{\bx_i | i \in c\}$ for $c \in \pi_n$.
Dissimilarity between $c_0 \in \pi_n$ and $c_1 \in \pi_n$ is given by $\mathsf d(c_0,c_1)$.
\subsection{Agglomerative Hierarchical Clustering}
\label{subsec:ahc}
Given $\bX$ and $\mathsf d(\cdot,\cdot)$, a common approach to building a binary tree for agglomerative hierarchical clustering
is the greedy algorithm, where pairs of nodes are merged as one moves up the hierarchy, starting in its leaf nodes
(\textbf{\small{Algorithm~\ref{alg:greedy_ahc}}}).
A naive implementation of the greedy algorithm requires $O(n^3)$ in time since each iteration needs $O(n^2)$
to find the pair of closest nodes and the algorithm runs over $n$ iterations.
It requires $O(n^2)$ in space to store pairwise dissimilarities.
The time complexity can be reduced to $O(n^2\log n)$
with priority queue, and can be reduced further for some special cases; for example,
in single linkage clustering where
\bee
\mathsf d(c_0,c_1) = \displaystyle{\min_{i\in c_0, j\in c_1}} \mathsf d(\{i\},\{j\}),
\eee
the time complexity is $O(n^2)$ since building a binary tree is equivalent to finding the minimum spanning tree in the dissimilarity graph.
Also, in case of the centroid linkage clustering where the distance between clusters are defined as the Euclidean distance between
the centers of the clusters, one can reduce the time complexity to $O(kn^2)$ where $k \ll n$~\citep{DayWHE84joc}.
\begin{algorithm}[htp]
\small
\caption{\small Greedy algorithm for agglomerative hierarchical clustering}
\label{alg:greedy_ahc}
\begin{algorithmic}[1]
\REQUIRE $\bX = \{\bx_i\}_{i=1}^n$, $\mathsf d(\cdot,\cdot)$.
\ENSURE Binary tree.
\STATE Assign data points to leaves.
\STATE Compute $\mathsf d(\{i\},\{j\})$ for all $i,j\in [n]$.
\WHILE{the number of nodes $>$ 1}
\STATE Find a pair $(c_0,c_1)$ with minimum $\mathsf d(c_0,c_1)$.
\STATE Merge $c \leftarrow c_0\cup c_1$ and compute $\mathsf d(c, c')$ for all $c'\neq c$.
\ENDWHILE
\end{algorithmic}
\end{algorithm}
\begin{algorithm}[htp]
\small
\caption{\small Nearest neighbor chain algorithm}
\label{alg:nnchain_ahc}
\begin{algorithmic}[1]
\REQUIRE $\bX = \{\bx_i\}_{i=1}^n$, reducible $\mathsf d(\cdot,\cdot)$.
\ENSURE Binary tree.
\WHILE{the number of nodes $>$ 1}
\STATE Pick any $c_0$.
\STATE Build a chain $c_1 = \mathrm{nn}(c_0), c_2 = \mathrm{nn}(c_1),\dots$,
where $\mathrm{nn}(c) \defeq \argmin_{c'} \mathsf d(c,c')$.
Extend the chain until $c_i = \mathrm{nn}(c_{i-1})$ and $c_{i-1} = \mathrm{nn}(c_i)$.
\STATE Merge $c\leftarrow c_i\cup c_{i-1}$.
\STATE If $i \geq 2$, go to line 3 and extend the chain from $c_{i-2}$. Otherwise, go to line 2.
\ENDWHILE
\end{algorithmic}
\end{algorithm}
\subsection{Reducibility}
\label{subsec:reducibility}
Two nodes $c_0$ and $c_1$ are {\em reciprocal nearest neighbors} (RNNs) if the dissimilarity $\mathsf d(c_0,c_1)$
is minimal among all dissimilarities from $c_0$ to elements in $\pi_n$ and also minimal among all dissimilarities from $c_1$.
Dissimilarity $\mathsf d(\cdot,\cdot)$ is {\em reducible}~\citep{BruynoogheM78}, if for any $c_0,c_1,c_2 \in \pi_n$,
\be
\label{eq:reducibility}
&\mathsf d(c_0,c_1) \leq \min\{\mathsf d(c_0,c_2),\mathsf d(c_1,c_2)\}\nn
& \Rightarrow \min\{\mathsf d(c_0,c_2),\mathsf d(c_1,c_2)\} \leq \mathsf d(c_0\cup c_1, c_2).
\ee
The reducibility ensures that if $c_0$ and $c_1$ are reciprocal nearest neighbors (RNNs),
then this pair of nodes are the closest pair that the greedy algorithm will eventually find by searching on an entire space.
Thus, the reducibility saves the effort of finding a pair of nodes with minimal distance.
Assume that $(c_0,c_1)$ are RNNs. Merging $(c_0,c_1)$ become problematic only if, for other RNNs $(c_2,c_3)$,
merging $(c_2,c_3)$ changes the nearest neighbor of $c_0$ (or $c_1$) to $c_2\cup c_3$.
However this does not happen since
\bee
&\mathsf d(c_2,c_3) \leq \min\{\mathsf d(c_2,c_0),\mathsf d(c_3,c_0)\} \\
&\Rightarrow \min\{\mathsf d(c_2,c_0),\mathsf d(c_3,c_0)\} \leq \mathsf d(c_2\cup c_3,c_0)\\
&\Rightarrow \mathsf d(c_0,c_1) \leq \mathsf d(c_2\cup c_3,c_0).
\eee
The nearest neighbor chain algorithm \citep{BruynoogheM78} enjoys this property and
find pairs of nodes to merge by following paths in the nearest neighbor graph of the nodes until the paths terminate
in pairs of mutual nearest neighbors ({\small{\textbf{Algorithm~\ref{alg:nnchain_ahc}}}}).
The time and space complexity of the nearest neighbor chain algorithm are $O(n^2)$ and $O(n)$, respectively.
The reducible dissimilarity includes those of single linkage and Ward's method~\citep{WardJH63jasa}.
\subsection{Agglomerative Bregman Clustering}
\label{subsec:abc}
Agglomerative clustering with Bregman divergence \citep{Bregman67} as a dissimilarity measure was recently developed in
\citep{TelgarskyM2012icml}, where the clustering was formulated as the minimization of the sum of cost based on
the Bregman divergence between the elements in a cluster and center of the cluster.
This cost is closely related with the Bregman Information used for Bregman hard clustering~\citep{BanerjeeA2005jmlr}).
In \citep{TelgarskyM2012icml}, the dissimilarity between two clusters $(c_0, c_1)$ is defined as the change of cost function
when they are merged. As will be shown in this paper, this dissimilarity turns out to be identical to the one we derive
from the asymptotic limit of BHC.
Agglomerative clustering with Bregman divergence showed better accuracies than traditional agglomerative hierarchical
clustering algorithms on various real datasets~\citep{TelgarskyM2012icml}.
\subsection{Bayesian Hierarchical Clustering}
\label{subsec:bhc}
Denote by $\calT_c$ a tree whose leaves are $\bX_c$ for $c \in \pi_n$.
A binary tree constructed by BHC \citep{HellerKA2005icml} explains the generation of $\bX_c$
with two hypotheses compared in considering each merge:
(1) the first hypothesis $\calH_c$ where all elements in $\bX_c$ were generated from a single cluster $c$;
(2) the alternative hypothesis where $\bX_c$ has two sub-clusters $\bX_{c_0}$
and $\bX_{c_1}$, each of which is associated with subtrees $\calT_{c_0}$ and $\calT_{c_1}$, respectively.
Thus, the probability of $\bX_c$ in tree $\calT_c$ is written as:
\be
\label{eq:bhc_recursive}
\lefteqn{p(\bX_c|\calT_c) = p(\calH_c)p(\bX_c|\calH_c) }\nn
&+& \{1-p(\calH_c) \}p(\bX_{c_0}|\calT_{c_0})p(\bX_{c_1}|\calT_{c_1}),
\ee
where the prior $p(\calH_c)$ is recursively defined as
\be
&\gamma_{\{i\}} \defeq \alpha,\,\,\gamma_c \defeq \alpha\Gamma(|c|)+\gamma_{c_0}\gamma_{c_1}, \\
&p(\calH_c) \defeq \alpha\Gamma(|c|)/\gamma_c,
\ee
and the likelihood of $\bX_c$ under $\calH_c$ is given by
\be\label{eq:bhc_single}
p(\bX_c|\calH_c) \defeq \int \bigg\{\prod_{i\in c} p(\bx_i|\btheta)\bigg\} p(d\btheta).
\ee
Now, the posterior probability of $\calH_c$, which is the probability of merging $(c_0,c_1)$, is computed by Bayes rule:
\be
p(\calH_c|\bX_c,\calT_c) = \frac{p(\calH_c)p(\bX_c|\calH_c)}{p(\bX_c|\calT_c)}.
\ee
In \citep{LeeJH2014aistats}, an alternative formulation for the generative probability was proposed, which writes
the generative process via the unnormalized probabilities (potential functions):
\be
&\phi(\bX_c|\calH_c) \defeq \alpha\Gamma(|c|)p(\bX_c|\calH_c),\\
&\phi(\bX_c|\calT_c)\defeq \gamma_c p(\bX_c|\calT_c).
\ee
With these definitions, \eqref{eq:bhc_recursive} is written as
\be
\phi(\bX_c|\calT_c)= \phi(\bX_c|\calH_c) + \phi(\bX_{c_0}|\calT_{c_0})\phi(\bX_{c_1}|\calT_{c_1}),
\ee
and the posterior probability of $\calH_c$ is written as
\be
p(\calH_c|\bX_c,\calT_c) = \bigg\{1 + \frac{\phi(\bX_{c_0}|\calT_{c_0})\phi(\bX_{c_1}|\calT_{c_1})}{\phi(\bX_c|\calH_c)}\bigg\}^{-1}.
\ee
One can see that the ratio inside behaves as the dissimilarity between $c_0$ and $c_1$:
\be\label{eq:bhc_dis_1}
\mathsf d(c_0,c_1)\defeq \frac{\phi(\bX_{c_0}|\calT_{c_0})\phi(\bX_{c_1}|\calT_{c_1})}{\phi(\bX_c|\calH_c)}.
\ee
Now, building a binary tree follows {\small \textbf{Algorithm}~\ref{alg:greedy_ahc}} with the distance in (\ref{eq:bhc_dis_1}).
Beside this, BHC has a scheme to determine the number of clusters.
It was suggested in \citep{HellerKA2005icml} that the tree can be cut at points $p(\calH_c|\bX_c,\calT_c) < 0.5$.
It is equivalent to say that the we stop the algorithm if the minimum over $\mathsf d(c_0,c_1)$ is greater than 1.
Note that once the tree is cut, the result contains forests, each of which involves a cluster.
BHC is closely related to the marginal likelihood of DPM; actually, the prior $p(\calH_c)$ comes from
the predictive distribution of DP prior. Moreover, it was shown that computing $\phi(\bX_c|\calT_c)$ to build a tree naturally
induces a lower bound on the marginal likelihood of DPM, $p_{\mathrm{\tiny DPM}}(\bX_c)$~\citep{HellerKA2005icml}:
\be
\frac{\Gamma(\alpha)}{\Gamma(\alpha+n)}\phi(\bX_c|\calT_c)\leq p_{\mathrm{\tiny DPM}}(\bX_c).
\ee
Hence, in the perspective of the posterior inference algorithm for DPM, building tree in BHC is equivalent to computing the approximate
marginal likelihood. Also, cutting the tree at the level where $\mathsf d(c_0,c_1) > 1$ corresponds finding the MAP clustering of $\bX$.
In \citep{HellerKA2005icml}, the time complexity was claimed to be $O(n^2)$.
However, this does not count the complexity required to find a pair with the smallest dissimilarity via sorting.
For instance, with a sorting algorithm using priority queues, BHC requires $O(n^2\log n)$ in time.
The dissimilarity is very sensitive to tuning the hyperparameters involving the distribution over parametes$p(\btheta)$
required to compute $p(\bX_c|\calH_c)$.
An EM-like iterative algorithm was proposed in \citep{HellerKA2005icml} to tune the hyperparameters, but the repeated
execution of the algorithm is infeasible for large-scale data.
\subsection{Bregman Diverences and Exponential Families}
\label{subsec:brdiv_and_ef}
\begin{defn}
(\citep{Bregman67}) Let $\varphi$ be a strictly convex differentiable function defined on a convex set.
Then, the Bregman divergence, $\sfB_\varphi(\bx,\by)$, is defined as
\be
\sfB_\varphi(\bx,\by) = \varphi(\bx)-\varphi(\by) - \ip{\bx-\by,\nabla\varphi(\by)}.
\ee
\end{defn}
Various divergences belong to the Bregman divergence.
For instance, Euclidean distance or KL divergence is Bregman divergence, when
$\varphi(\bx)= \frac{1}{2}\bx^{\top} \bx$ or $\varphi(\bx) = \bx \log \bx$, respectively.
The exponential family distribution over $\bx\in\bbr^d$ with
natural parameter $\btheta\in\bTheta$ is of the form:
\be
p(\bx|\btheta) = \exp \Big\{ \ip{t(\bx),\btheta} - \psi(\btheta) - h(\bx) \Big\},
\ee
where $t(\bx)$ is sufficient statistics, $\psi(\btheta)$ is a log-partition function,
and $\exp\{- h(\bx)\}$ is a base distribution.
We assume that $p(\bx|\btheta)$ is \emph{regular} ($\bTheta$ is open) and $t(\bx)$ is \emph{minimal} ($\nexists \ba \in\bbr^d$ s.t. $\forall \bx,\,\, \ip{\ba,t(\bx)} = \mathrm{const}$).
Let $\varphi(\bmu)$ be the convex conjugate of $\psi$:
\be
\varphi(\bmu) = \sup_{\theta\in\Theta} \Big\{\ip{\bmu,\btheta}-\psi(\btheta) \Big\}.
\ee
Then, the Bregman divergence and the exponential family has the following relationship:
\begin{thm}
\citep{BanerjeeA2005jmlr}
Let $\varphi(\cdot)$ be the conjugate function of $\psi(\cdot)$.
Let $\btheta$ be the natural parameter and $\bmu$ be the corresponding expectation parameter, i.e.,
$\bmu = \bbe[t(\bx)] = \nabla\psi(\btheta)$.
Then $p(\bx|\btheta)$ is uniquely expressed as
\be
p(\bx|\btheta) & = & \exp \Big\{-\sfB_\varphi(t(\bx),\bmu) \Big\} \nonumber \\
& & \exp \Big\{\varphi(t(\bx))-h(\bx) \Big\}.
\ee
\end{thm}
The conjugate prior for $\btheta$ has the form:
\be
p(\btheta|\nu,\btau) = \exp \Big\{ \ip{\btau,\btheta} - \nu\psi(\btheta) - \xi(\nu,\btau) \Big\}.
\ee
$p(\btheta|\nu,\btau)$ can also be expressed with the Bregman divergence:
\be
\lefteqn{p(\btheta|\nu,\btau) = \exp\{-\nu\sfB_\varphi(\btau/\nu,\bmu)\}} \nn
&& \times \exp\Big\{\nu\varphi(\btau/\nu)-\xi(\nu,\btau)\Big\}.
\ee
\subsection{Scaled Exponential Families}
\label{subsec:scaled_ef}
Let $\widetilde\btheta = \beta\btheta$, and $\widetilde\psi(\widetilde\btheta) = \beta\psi(\widetilde\btheta/\beta) = \beta\psi(\btheta)$.
The scaled exponential family with scale $\beta$ is defined as follows~\citep{JiangK2012nips}:
\be
p(\bx|\widetilde\btheta) &=& \exp \Big\{ \ip{t(\bx),\widetilde\btheta}-\widetilde\psi(\widetilde\btheta)-h_\beta(\bx) \Big\}\nn
&=&\exp \Big\{ \beta \ip{t(\bx),\btheta}-\beta\psi(\btheta)-h_\beta(\bx) \Big\}.
\ee
For this scaled distribution, the mean $\bmu$ remains the same, and the covariance $\bSigma$ becomes
$\bSigma/\beta$ ~\citep{JiangK2012nips}.
Hence, the distribution is more concentrated around its mean. The scaled distribution in the Bregman divergence form is
\be
p(\bx|\widetilde\btheta) & = & \exp \Big\{-\beta\sfB_\varphi(t(\bx),\bmu) \Big\} \nonumber \\
& & \exp \Big\{\beta\varphi(t(\bx))-h_\beta(\bx) \Big\}.
\ee
According to $p(\bx|\widetilde\btheta)$, $p(\widetilde\btheta|\widetilde\btau,\widetilde\nu)$ is defined with $\widetilde\btau = \btau/\beta$, $\widetilde\nu = \nu/\beta$.
Actually, this yields the same prior as non-scaled distribution.
\be
p(\widetilde\btheta|\widetilde\nu,\widetilde\btau) &=&
\exp\Big\{\ip{\widetilde\btheta,\widetilde\btau}-
\widetilde\nu\widetilde\psi(\widetilde\btheta)-\xi_\beta(\widetilde\nu,\widetilde\btau)\Big\}\nn
&=& \exp \Big\{ \ip{\btheta,\btau}-\nu\psi(\btheta)-\xi(\nu,\btau) \Big\}.
\ee
\section{MAIN RESULTS}
\label{sec:main}
We present the main contribution of this paper.
From now on, we assume that the likelihood and prior in Eq.~\eqref{eq:bhc_single}
are scaled exponential families defined in Section \ref{subsec:scaled_ef}.
\subsection{Small-Variance Asymptotics for BHC}
\label{subsec:sva_bhc}
The dissimilarity in BHC can be rewritten as follows:
\be
\label{eq:bhc_dis_2}
\mathsf d(c_0,c_1) &=& \frac{\phi(\bX_{c_0}|\calT_{c_0})\phi(\bX_{c_1}|\calT_{c_1})}{\phi(\bX_c | \calH_c)}\nn
&=& \frac{\alpha\Gamma(|c_0|)\Gamma(|c_1|)p(\bX_{c_0}|\calH_{c_0})p(\bX_{c_1}|\calH_{c_1})}{\Gamma(|c|)p(\bX_c|\calH_c)}\nn
& \times & \Big\{ 1 + \mathsf d(c_{00},c_{01}) \Big\} \Big\{1+\mathsf d(c_{10},c_{11}) \Big\},
\ee
where $c_0 = c_{00 }\cup c_{01}$ and $c_1 = c_{10} \cup c_{11}$.
We first analyze the term $p(\bX_c|\calH_c)$, as in \citep{JiangK2012nips}.
\be
\label{eq:integral}
\lefteqn{p(\bX_c|\calH_c) = \disint \bigg\{\prod_{i\in c} p(\bx_i|\widetilde\btheta)\bigg\} p(\widetilde\btheta|\nu,\btau)d\widetilde\btheta}\nn
&=&\beta^d\disint\exp \bigg\{ \bigg\langle \btheta, \btau + \beta\sum_{i\in c} t(\bx_i)\bigg\rangle - (\nu+\beta|c|)\psi(\btheta)\nn
& & -\sum_{i\in c} h_{\beta}(\bx_i) - \xi(\nu,\btau)\bigg\} d\btheta\nn
&=&\beta^d \exp\bigg\{(\nu+\beta|c|)\varphi(\bmu_c)-\sum_{i\in c} h_{\beta}(\bx_i) - \xi(\nu,\btau)\bigg\}\nn
& & \times \int \exp\{-(\nu+\beta|c|)\sfB_\varphi(\bmu_c,\bmu)\}d\btheta,
\ee
where
\be
\bmu_c \defeq \frac{\btau+\beta\sum_{i\in c} t(\bx_i)}{\nu+\beta|c|}.
\ee
Note that $\bmu = \nabla\psi(\btheta)$ is a function of $\btheta$. The term inside the integral of Eq.~\eqref{eq:integral} has a local minimum
at $\bmu = \bmu_c$, and thus can be approximated by Laplace's method:
\be
\label{eq:integral_approx}
&= \beta^d \exp\bigg\{ (\nu+\beta |c|)\varphi(\bmu_c) - \displaystyle{\sum_{i\in c}} h_{\beta}(\bx_i) - \xi(\nu,\btau)\bigg\}\nn
& \bigg(\dfrac{2\pi}{\nu+\beta|c|}\bigg)^{\frac{d}{2}}\bigg|\dfrac{\partial^2\sfB_\varphi(\bmu_c,\bmu)}{\partial\btheta\partial\btheta\tr}\bigg|^{-\frac{1}{2}}_{\bmu=\bmu_c}\Big\{1+O(\beta^{-1})\Big\}.
\ee
It follows from this result that, as $\beta \rightarrow \infty$,
the asymptotic limit of dissimilarity $\mathsf d(c_0,c_1)$ in \eqref{eq:bhc_dis_2} is given by
{\small
\bee
\label{eq:lim}
\lefteqn{\lim_{\beta\to\infty}\frac{\alpha\Gamma(|c_0|)\Gamma(|c_1|)p(\bX_{c_0}|\calH_{c_0})p(\bX_{c_1}|\calH_{c_1})}{\Gamma(|c|)p(\bX_c|\calH_c)}}\\
&\propto& \hspace*{-.1in} \lim_{\beta\to\infty}\alpha\beta^{\frac{d}{2}}\exp \Bigl\{ \beta(|c_0|\varphi(\bmu_{c_0})
+ |c_1|\varphi(\bmu_{c_1}) - |c|\varphi(\bmu_c)) \Bigr\}\nonumber,
\eee}
Let $\alpha = \beta^{-\frac{d}{2}}\exp(-\beta\lambda)$, then we have
\bee
= \lim_{\beta\to\infty}\exp \Big\{\beta(|c_0|\varphi(\bmu_{c_0}) + |c_1|\varphi(\bmu_{c_1}) - |c|\varphi(\bmu_c) - \lambda) \Big\}.
\eee
As $\beta \to \infty$, the term inside the exponent converges to
\be
|c_0|\varphi(\bar t_{c_0}) + |c_1|\varphi(\bar t_{c_1}) - |c|\varphi(\bar t_{c})-\lambda,
\ee
where
\be
\bar t_c \defeq \frac{1}{|c|} \sum_{i\in c} t(\bx_i),
\ee
and this is the average of sufficient statistics for cluster $c$.
With this result, we define a new dissimilarity $\mathsf d_\star(c_0,c_1)$ as
\be
\label{eq:bhc_dis_star}
\mathsf d_\star(c_0,c_1) \hspace*{-.1in} & \defeq & \hspace*{-.1in} |c_0|\varphi(\bar t_{c_0}) + |c_1|\varphi(\bar t_{c_1}) - |c|\varphi(\bar t_{c}) \nn
\hspace*{-.1in} &=& \hspace*{-.1in} |c_0|\varphi(\bar t_{c_0}) + |c_1|\varphi(\bar t_{c_1})\nn
\hspace*{-.1in} & - & \hspace*{-.1in} (|c_0|+|c_1|)\varphi\bigg(\frac{|c_0|\bar t_{c_0}+|c_1|\bar t_{c_1}}{|c_0|+|c_1|}\bigg).
\ee
Note that $\mathsf d_\star(\cdot,\cdot)$ is always positive since $\varphi$ is convex.
If $\mathsf d_\star(c_0,c_1) \geq \lambda$, the limit Eq.~\eqref{eq:lim} diverges to $\infty$,
and converges to zero otherwise. When $|c|=1$, Eq.~\eqref{eq:lim} is the same as the limit of the dissimilarity $\mathsf d(c_0,c_1)$,
and thus the dissimilarity diverges when $\mathsf d_\star(c_0,c_1) \geq \lambda$ and converges otherwise. When $|c|>1$, assume that the dissimilarities of
children $\mathsf d(c_{00},c_{01})$ and $\mathsf d(c_{10},c_{11})$ converges to zero. From Eq.~\eqref{eq:bhc_dis_2}, we can easily see that
$\mathsf d(c_0,c_1)$ converges only if $\mathsf d_\star(c_0,c_1) < \lambda$. In summary,
\be\label{eq:bhc_asymp_thres}
\lim_{\beta\to\infty} \mathsf d(c_0,c_1) = \bigg\{ \begin{array}{ll} 0 & \textrm{ if } \mathsf d_\star(c_0,c_1) < \lambda,\\ \infty & \textrm { otherwise.}\end{array}.
\ee
In similar way, we can also prove the following:
\be\label{eq:bhc_asymp_compare}
\lim_{\beta\to\infty} \frac{\mathsf d(c_0,c_1)}{\mathsf d(c_2,c_3)} = \bigg\{ \begin{array}{ll} 0 & \textrm{ if } \mathsf d_\star(c_0,c_1) < \mathsf d_\star(c_2,c_3),\\ \infty & \textrm { otherwise.}\end{array},
\ee
which means that comparing two dissimilarities in original BHC is equivalent to comparing the new dissimilarities $\mathsf d_\star(\cdot,\cdot)$,
and we can choose the next pair to merge by comparing $\mathsf d_\star(\cdot,\cdot)$ instead of $\mathsf d(\cdot,\cdot)$.
With Eqs.~\eqref{eq:bhc_asymp_thres} and \eqref{eq:bhc_asymp_compare},
we conclude that when $\beta\to\infty$, BHC reduces to {\textbf{Algorithm~\ref{alg:greedy_ahc}}}
with dissimilarity measure $\mathsf d_\star(\cdot,\cdot)$ and threshold $\lambda$,
where the algorithm terminates when the minimum $\mathsf d_\star(\cdot,\cdot)$ exceeds $\lambda$.
On the other hand, a simple calculation yields
\bee
\mathsf d_\star(c_0,c_1) = |c_0| \sfB_\varphi(\bar t_{c_0}, \bar t_c) + |c_1|\sfB_\varphi(\bar t_{c_1},\bar t_c),
\eee
which is exactly same as the dissimilarity proposed in \citep{TelgarskyM2012icml}.
Due to the close relationship between exponential family and the Bregman divergence,
the dissimilarities derived from two different perspective has the same form.
As an example, assume that $p(\bx|\btheta) = \calN(\bx|\bmu,\sigma^2\bI)$ and $p(\bmu) = \calN(\bmu|0,\rho^2\bI)$. We have $\varphi(\bx) = \norm{\bx}^2/(2\sigma^2)$ and
\be\label{eq:ward}
\mathsf d_\star(c_0,c_1) = \frac{|c_0||c_1|\norm{\bx_{c_0}-\bx_{c_1}}^2}{2\sigma^2(|c_0|+|c_1|)},
\ee
which is same as the Ward's merge cost~\citep{WardJH63jasa}, except for the constant $1/(2\sigma^2)$. Other examples
can be found in \citep{BanerjeeA2005jmlr}.
Note that $\mathsf d_\star(\cdot,\cdot)$ does not need hyperparameter tunings, since
the effect of prior $p(\btheta)$ is ignored as $\beta\to 0$. This provides a great advantage
over BHC which is sensitive to the hyperparameter settings.
\textbf{Smoothing}: In some particular choice of $\varphi$, the singleton clusters may have degenerate values~\citep{TelgarskyM2012icml}.
For example, when $p(\bx|\btheta) = \mathrm{Mult}(\bx|m,q)$, the function $\varphi(\bx) = \sum_{j=1}^d x_j\log(x_j/m)$ has degenerate values
when $x_j = 0$. To handle this, we use the smoothing strategy proposed in \citep{TelgarskyM2012icml}; instead of the
original function $\varphi(\bx)$, we use the smoothed functions $\varphi_0(\bx)$ and $\varphi_1(\bx)$ defined as follows:
\be
&\varphi_0(\bx) \defeq \varphi((1-\alpha)\bx + \alpha \bgamma), \\
&\varphi_1(\bx) \defeq \varphi(\bx + \alpha \bgamma),
\ee
where $\alpha\in(0,1)$ be arbitrary constant and $\bgamma$ must in the relative interior of the domain of $\varphi$.
In general, we use $\varphi_0(\bx)$ as a smoothed function, but we can also use $\varphi_1(\bx)$ when the domain of $\varphi$ is a convex cone.
\textbf{Heuristics for choosing $\lambda$}: As in \citep{KulisB2012icml}, we choose the threshold value $\lambda$.
Fortunately, we found that the clustering accuracy was not extremely sensitive to the choice of $\lambda$; merely
selecting the scale of $\lambda$ could result in reasonable accuracy. There can be many simple heuristics,
and the one we found effective is to use the $k$-means clustering. With the very rough guess on the desired number
of clusters $\widetilde k$, we first run the $k$-means clustering (with Euclidean distance) with $k = a\widetilde k$ (we fixed
$a = 4$ for all experiments). Then, $\lambda$ was set to the average value of dissimilarities $\mathsf d_\star(\cdot,\cdot)$
between the all pair of $k$ centers.
\subsection{Reducibility of $\mathsf d_\star(\cdot,\cdot)$}
\label{subsec:reducible}
The relaxed BHC with small-variance asymptotics still has the same complexities to BHC.
If we can show that $\mathsf d_\star(\cdot,\cdot)$ is reducible, we can reduce the complexities by adapting the nearest neighbor chain algorithm.
Unfortunately, $\mathsf d_\star(\cdot,\cdot)$ is not reducible in general (one can easily find counter-examples for some distributions).
However, we argue that $\mathsf d_\star(\cdot,\cdot)$ is reducible in many cases,
and thus applying the nearest neighbor chain algorithm as if $\mathsf d_\star(\cdot,\cdot)$ is reducible
does not degrades the clustering accuracy. In this section, we show the reason by analyzing $\mathsf d_\star(\cdot,\cdot)$.
At first, we investigate a term inside the dissimilarity:
\be
f(\bar t_c) \defeq |c|\varphi(\bar t_c).
\ee
The second-order Taylor expansion of this function around the mean $\bmu$ yields:
\be
\label{eq:taylor}
&f(\bar t_c) = |c|\varphi(\bmu) + |c| \varphi^{(1)}(\bmu)\tr (\bar t_c-\bmu)\nn
& + \Delta_\varphi(\bar t_c,\bmu) + \epsilon_\varphi(\bar t_c,\bmu),
\ee
where $\varphi^{(k)}$ is the $k$th order derivative of $\varphi$, and
\be
\Delta_\varphi(\bar t_c) \defeq \frac{|c|}{2}(\bar t_c-\bmu)\tr\varphi^{(2)}(\bmu)(\bar t_c-\bmu), \\
\epsilon_\varphi(\bar t_c) \defeq |c|\sum_{|\balpha|=3} \frac{\partial^{\balpha} \varphi(\bnu)}{\balpha!}(\bar t_c-\bmu)^{\balpha}.
\ee
Here, $\balpha$ is the multi-index notation, and $\bnu = \bmu + k(\bar t_c-\bmu)$ for some $k\in(0,1)$. The term $\Delta_\varphi(\bar t_c)$ plays an important role in analyzing the
reducibility of $\mathsf d_\star(\cdot,\cdot)$. To bound the error term $\epsilon_\varphi(\bar t_c)$,
we assume that $|\varphi^{(3)}|\leq M$\footnote{This assumption holds for the most of distributions we will discuss (if properly smoothed), but not holds in general.}. As earlier, assume that $\bar t_c$ is a average of $|c|$ observations generated from the same
scaled-exponential family distribution:
\be
\bx_1,\dots, \bx_n \isim p(\cdot | \beta,\btheta),\quad \bar t_c = \frac{1}{|c|}\sum_{i\in c} t(\bx_i).
\ee
By the property of the log-partition function of the exponential family distribution, we get the following results:
\be\label{eq:expected_error}
\bbe[\epsilon_\varphi(\bar t_c)] = \frac{1}{\beta^2 |c|}\sum_{|\balpha|=3} \frac{\partial^{\balpha}\varphi(\bnu)\partial^{\balpha} \psi(\btheta)}{\balpha!}.
\ee
One can see that the expected error converges to zero as $\beta^2|c|\to\infty$. Also, it can be shown that
the expectation of the ratio of two terms converges to zero as $\beta \to \infty$:
\be
\lim_{\beta\to\infty} \bbe\bigg[\frac{\epsilon_\varphi(\bar t_c)}{\Delta_\varphi(\bar t_c)}\bigg] = 0,
\ee
which means that $\Delta_\varphi(\bar t_c)$ asymptotically dominates $\epsilon_\varphi(\bar t_c)$ (detailed derivations are given in the supplementary material).
Hence, we can safely approximate $f(\bar t_c)$ up to second order term.
Now, let $c_0$ and $c_1$ be clusters belong to the same super-cluster (i.e. $\bX_{c_0}$ and $\bX_{c_1}$ were
generated from the same mean vector $\bmu$). We don't need to investigate the case where the pair belong to a different cluster,
since then they will not be merged anyway $(\mathsf d_\star(\cdot,\cdot) \geq \lambda)$ in our algorithm. By the independence, $\bbe[\bar t_{c_0}] = \bbe[\bar t_{c_1}] = \bbe[\bar t_{c_0\cup c_1}] = \bmu$. Applying the approximation~\eqref{eq:taylor}, we have
\be
\lefteqn{\mathsf d_\star(c_0,c_1) \approx \Delta_\varphi(\bar t_{c_0}) + \Delta_\varphi(\bar t_{c_1}) - \Delta_\varphi(\bar t_{c_0\cup c_1}) }\nn
& = \frac{|c_0||c_1|}{2(|c_0|+|c_1|)}(\bar t_{c_0}-\bar t_{c_1})\tr \varphi^{(2)}(\bmu) (\bar t_{c_0}-\bar t_{c_1}).
\ee
This approximation, which we will denote as $\widetilde\mathsf d_\star(c_0,c_1)$, is a generalization of the Ward's cost~\eqref{eq:ward} from Euclidean distance to Mahalanobis distance with matrix $\varphi^{(2)}(\bmu)$ (note that this approximation is exact for the spherical Gaussian case).
More importantly, $\widetilde\mathsf d_\star(c_0,c_1)$ is reducible.
\begin{thm}
\bee
&\widetilde\mathsf d_\star(c_0,c_1) \leq \min \Big\{\widetilde\mathsf d_\star(c_0,c_2),\widetilde\mathsf d_\star(c_1,c_2) \Big\}\nn
& \Rightarrow \min \Big\{\widetilde\mathsf d_\star(c_0,c_2),\widetilde\mathsf d_\star(c_1,c_2) \Big\} \leq \widetilde\mathsf d_\star(c_0\cup c_1, c_2).
\eee
\end{thm}
\begin{proof}
For the Ward's cost, the following Lance-Williams update formula~\citep{LanceGN67tcj} holds for $\widetilde\mathsf d_\star(\cdot,\cdot)$:
\be
\lefteqn{\widetilde\mathsf d_\star(c_0\cup c_1,c_2)=\frac{(|c_0|+|c_2|)\widetilde\mathsf d_\star(c_0,c_2)}{|c_0|+|c_1|+|c_2|}}\nn
& & + \frac{(|c_1|+|c_2|)\widetilde\mathsf d_\star(c_1,c_2)-|c_2|\widetilde\mathsf d_\star(c_0,c_1)}{|c_0|+|c_1|+|c_2|}.
\ee
Hence, by the assumption, we get
\bee
\widetilde\mathsf d_\star(c_0\cup c_1,c_2) \geq \min \Big\{\widetilde\mathsf d_\star(c_0,c_2),\widetilde\mathsf d_\star(c_1,c_2) \Big\}.
\eee
\end{proof}
\begin{algorithm}[t!]
\small
\caption{\small Nearest neighbor chain algorithm for BHC with small-variance asymptotics}
\label{alg:nnchain_ahc_thres}
\begin{algorithmic}[1]
\REQUIRE $\bX = \{\bx_i\}_{i=1}^n$, $\mathsf d_\star(\cdot,\cdot)$, $\lambda$.
\ENSURE A clustering $C$.
\STATE Set $R = [n]$ and $C = \varnothing$.
\WHILE{$R \neq \varnothing$}
\STATE Pick any $c_0\in R$.
\STATE Build a chain $c_1 = \mathrm{nn}(c_0), c_2 = \mathrm{nn}(c_1),\dots$,
where $\mathrm{nn}(c) \defeq \argmin_{c'} \mathsf d_star(c,c')$.
Extend the chain until $c_i = \mathrm{nn}(c_{i-1})$ and $c_{i-1} = \mathrm{nn}(c_i)$.
\STATE Remove $c_i$ and $c_{i-1}$ from $R$.
\IF{$\mathsf d_\star(c_{i-1},c_i)<\lambda$}
\STATE Add $c = c_i\cup c_{i-1}$ to $R$.
\STATE If $i \geq 2$, go to line 3 and extend the chain from $c_{i-2}$. Otherwise, go to line 2.
\ELSE
\STATE Add $c_i$ and $c_{i-1}$ to $C$.
\ENDIF
\ENDWHILE
\end{algorithmic}
\end{algorithm}
As a result, the dissimilarity $\mathsf d_\star(\cdot,\cdot)$ is reducible provided that the Taylor's approximation \eqref{eq:taylor} is accurate.
In such a case, one can apply the nearest-neighbor chain algorithm with $\mathsf d_\star(\cdot,\cdot)$,
treating
as if it is reducible to build a binary tree in $O(n^2)$ time and $O(n)$ space. Unlike
\textbf{Algorithm~\ref{alg:nnchain_ahc}}, we have a threshold $\lambda$ to determine the number of clusters,
and we present a slightly revised algorithm~({\textbf{Algorithm~\ref{alg:nnchain_ahc_thres}}). Note again
that the revised algorithm generates forests instead of trees.
\begin{table}
\small
\centering
\caption{Average value of the exact dissimilarity $\mathsf d_\star$, approximate dissimilarity $\widetilde\mathsf d_\star$
and relative error, and number of not-reducible case among 100,000 trials.}
\begin{tabular}{|c|c|c|c|c|}
\hline & $\bbe[\mathsf d_\star]$ & $\bbe[\widetilde{\mathsf d}_\star]$ & \parbox{1.2cm}{average relative error ($\%$)} & \parbox{1.2cm}{ not\\ reducible $/ 100k$ } \\
\hline Poisson & 0.501 & 0.500 & 1.577 & 0 \\
Multinomial & 3.594 & 3.642 & 4.675 & 87 \\
Gaussian & 8.739 & 8.717 & 7.716 & 48 \\
\hline
\end{tabular}
\label{table:reducibility}
\end{table}
\begin{figure*}
\centering
\includegraphics[width = 0.4\linewidth]{./exp1_gaussian_relerr_csize.pdf}
\includegraphics[width = 0.4\linewidth]{./exp1_gaussian_relerr_beta.pdf}
\caption{Average relative error vs maximum cluster size and scale factor $\beta$.}
\label{fig:reduciblllity}
\end{figure*}
\begin{table*}
\small
\centering
\caption{Average adjusted Rand index values for randomly generated datasets. Best ones are marked as bold face.}
\begin{tabular}{|c|c|c|c|c|c|c|}
\hline & Poisson ($k$) & Poisson ($2k$) & multinomial ($k$) & multinomial ($2k$) & Gaussian ($k$) & Gaussian ($2k$) \\
\hline single linkage & 0.091 (0.088) & 0.015 (0.030) & 0.000 (0.000) & 0.000 (0.000) & 0.270 (0.280) & 0.057 (0.055) \\
complete linakge & 0.381 (0.091) & 0.263 (0.059) & 0.266 (0.144) & 0.090 (0.025) & 0.565 (0.166) & 0.475 (0.141) \\
Ward's method & 0.465 (0.119) & 0.273 (0.049) & 0.770 (0.067) & 0.564 (0.055) & 0.779 (0.145) & 0.763 (0.122) \\
RBHC-greedy & \textbf{0.469} (0.112) & \textbf{0.290} (0.056) & 0.870 (0.046) & 0.733 (0.028) & \textbf{0.875} (0.087) & \textbf{0.883} (0.067) \\
RBHC-nnca & \textbf{0.469} (0.112) & \textbf{0.290} (0.056) & 0.865 (0.045) & 0.736 (0.040) & \textbf{0.875} (0.087) & \textbf{0.883} (0.067) \\
BHC & 0.265 (0.080) & 0.134 (0.052) & \textbf{0.907} (0.069) & \textbf{0.894} (0.044) & 0.863 (0.109) & 0.860 (0.108) \\
\hline
\end{tabular}
\label{table:synthetic_ari}
\end{table*}
\section{EXPERIMENTS}
\label{sec:experiments}
\subsection{Experiments on Synthetic Data}
\label{subsec:synthetic_experiments}
{\textbf{Testing the reducibility of $\mathsf d_\star(\cdot,\cdot)$}}:
We tested the reducibility of $\mathsf d_\star(\cdot,\cdot)$ empirically.
We repeatedly generated the three clusters $c_0, c_1$ and $c_2$ from
the exponential family distributions, and counted the number of cases where the dissimilarities between those
clusters are not reducible. We also measured the average value of the relative error to support our arguments
in Section~\ref{subsec:reducible}.
We tested three distributions; Poisson, multinomial and Gaussian.
At each iteration, we first sampled the size of the clusters $|c_0|$, $|c_1|$ and $|c_2|$
from $\mathrm{Unif}([20,100])$. Then we sampled the three clusters from one of the three distributions,
and computed $\mathsf d_\star(c_0,c_1), \mathsf d_\star(c_1,c_2)$ and $\mathsf d_\star(c_0, c_2)$. We then first checked
whether these three values satisfy the reducibility condition~\eqref{eq:reducibility} (for example,
if $\mathsf d_\star(c_0,c_2)$ is the smallest, we checked if $\mathsf d_\star(c_0\cup c_2, c_1) \geq \min\{ \mathsf d_\star(c_0,c_1),
\mathsf d_\star(c_2, c_1)\}$). Then, for $\mathsf d_\star(c_0,c_1)$, we computed the approximate value $\widetilde\mathsf d_\star(c_0,c_1)$
and measured the relative error
\be
2\times\left|\frac{ \mathsf d_\star(c_0,c_1)-\widetilde{\mathsf d}_\star(c_0,c_1) }{\mathsf d_\star(c_0,c_1)+\widetilde{\mathsf d}_\star(c_0,c_1)}\right|.
\ee
We repeated this process for $100,000$ times for the three distributions and measured the average values.
For Poisson distribution, we sampled the mean $\rho \sim\mathrm{Gamma}(2, 0.05)$ and sampled the
data from $\mathrm{Poisson}(\rho)$. We smoothed the function $\varphi(\bx) = \bx\log(\bx)-\bx$ as $\varphi(\bx+0.01)$
to prevent degenerate function values. For multinomial distribution, we tested the case where the dimension $d$ is $10$
and the number of trials $m$ is 5. We sampled the parameter $\bq\sim\mathrm{Dir}(5\cdot\mathbf{1}_{d})$ where $\mathbf{1}_d$
is $d$-dimensional one vector, and sampled the data from $\mathrm{Mult}(\bq)$. We smoothed the function $\varphi(\bx) = \sum_{j=1}^d x_j \log(x_j/m)$ as $\varphi(0.9\bx + 0.1 m \mathbf{1}_d / d)$. For Gaussian, we
tested with $d=10$, and sampled the mean and covariance $\bmu,\bSigma \sim \calN(0,(0.08\cdot\bLambda)^{-1})\calW (d+2, \bPsi)$ ($\bPsi = \bA \bA\tr + d\bI$ where $\bA$ was sampled from unit Gaussian). We smoothed
the function $\varphi(\bx,\bX) = - \frac{1}{2}\log\det(\bX-\bx\bx\tr)$ as $-\frac{1}{2}\log\det(\bX-\bx\bx\tr + 0.01\bI)$.
The result is summarized in Table~\ref{table:reducibility}. the generated dissimilarities were reducible in most case, as expected.
The relative error was small, which supports our arguments of the reason why $\mathsf d_\star(\cdot,\cdot)$ is reducible with high probability.
We also measured the change of average relative error by controlling two factors; the maximum cluster size $n_{\mathrm{max}}$
and variance scale factor $\beta$. We plotted the average relative error of Gaussian distribution by changing those two factors,
and the relative error decreased as predicted (Figure~\ref{fig:reduciblllity}).
\textbf{Clustering synthetic data}: We evaluated the clustering accuracies of original BHC (BHC),
BHC in small-variance limit with greedy algorithm (RBHC-greedy), BHC in small-variance limit with nearest neighbor chain
method (RBHC-nnca), single linkage, complete linkage, and Ward's method. We generated the datasets from Poisson,
multinomial and Gaussian distribution. We tested two types of data; 1,000 elements with 6 clusters and 2,000 elements with 12 clusters.
For Poisson distribution, each mixture component was generated from $\mathrm{Poisson}(\rho)$ with $\rho\sim\mathrm{Gamma}(2,0.05)$
for both datasets. For multinomial distribution, we set $d=20$ and $m=10$
for 1,000 elements dataset, and set $d=40$ and $m=10$ for 2,000 elements datasets. For both dataset, we sampled
the parameter $\bq\sim\mathrm{Dir}(0.5\cdot\mathbf{1}_d)$. For Gaussian case, we set $d=3$ for 1,000 elements dataset
and set $d=6$ for 2,000 elements dataset. We sampled the parameters from $\calN(0,(0.08\cdot\Lambda)^{-1})\calW(\bPsi,6)$
for 1,000 elements, and from $\calN(0,(0.2\cdot\Lambda)^{-1})\calW(\bPsi,9)$ for 2,000 elements. For each distribution
and type (1,000 or 2,000), we generated 10 datasets for each type and measured the average clustering accuracies.
We evaluated the clustering accuracy using the adjusted Rand index~\citep{HubertL85joc}. For traditional
agglomerative clustering algorithms, we assumed that we know the true number of clusters $k$ and cut the tree at corresponding level.
For RBHC-greedy and RBHC-nnca, we selected the threshold $\lambda$ with the heuristics described in Section~\ref{subsec:sva_bhc}.
For original BHC, we have to carefully tune the hyperparemters, and the accuracy was very sensitive to this setting.
In the case of Poisson distribution where $\rho \sim\mathrm{Gamma}(a, b)$, we have to tune two
hyperparameters $\{a, b\}$. For multinomial case where $\bq \sim \mathrm{Dir}(\balpha)$, we set
$\balpha = \alpha \mathbf{1}_d$ and tuned $\alpha$.
For Gaussian case where $(\bmu,\bSigma)\sim \calN(\bm, (r\bLambda)^{-1})\calW(\nu,\bPsi^{-1})$,
we have four hyperparameters $\{\bm, r, \nu, \bPsi\}$. We set $\bm$ to be the empirical mean of $\bX$
and fixed $r=0.1$ and $\nu = d + 6$. We set $\bPsi = k \bS$ where $\bS$ is the empirical covariance of $\bX$
and controlled $k$ according to the dimension and the size of the data. The result is summarized in Table~\ref{table:synthetic_ari}.
The accuracies of RBHC-greedy and RBHC-nnca were best for most of the cases, and the accuracies of the two methods
were almost identical expect for the multinomial distribution. BHC was best for the multinomial case
where the hyperparameter tuning was relatively easy, but showed poor performance in Poisson case (we failed
to find the best hyperparameter setting in that case). Hence, it would be a good choice to use
RBHC-greedy or RBHC-nnca which do not need careful hyperparameter tuning, and RBHC-nnca may be the best
choice considering its space and time complexity compared to RBHC-greedy.
\subsection{Experiments on Real Data}
\label{subsec:real_experiments}
We tested the agglomerative clustering algorithms on two types of real-world data. The first one
was a subset of MNIST digit database~\citep{LeCunY98procieee}. We scaled down the original $28\times 28$
to $7\times 7$ and vectorized each image to be $\bbr^{49}$ vector. Then we sampled 3,000 images
from the classes $0, 3, 7$ and $9$. We clustered this dataset with Gaussian asssumption.
The second one was visual-word data extracted from Caltech101 database~\citep{FeiFeiL2004cvprw}.
We sampled 2,033 images from "Airplane", "Mortorbikes" and "Faces-easy" classes, and extracted
SIFT features for image patches. Then we quantized those features into 1,000 visual words. We clustered
the data with multinomial assumption. Table~\ref{table:real_ari} shows the ARI values of agglomerative
clustering algorithms. As in the synthetic experiments, the accuracy RBHC-greedy and RBHC-nnca were identical,
and outperformed the traditional agglomerative clustering algorithms. BHC was best for Caltech101, where
the multinomial distribution with easy hyperparameter tuning was assumed. However, BHC was even worse
than Ward's method for MNIST case, where we failed to tune $49\times 49$ matrix $\bPsi$.
\begin{table}[ht!]
\small
\centering
\caption{Average adjusted Rand index values for MNIST and Caltech101 datasets. Best ones are marked as bold face.}
\begin{tabular}{|c|c|c|c|c|c|c|}
\hline & MNIST & Caltech101\\
\hline single linkage & 0.000 & 0.000 \\
complete linakge & 0.187 & 0.000 \\
Ward's method & 0.485 & 0.465 \\
RBHC-greedy & \textbf{0.637} & 0.560 \\
RBHC-nnca & \textbf{0.637} & 0.560 \\
BHC & 0.253 & \textbf{0.646} \\
\hline
\end{tabular}
\label{table:real_ari}
\end{table}
\section{CONCLUSIONS}
\label{sec:conclusions}
In this paper we have presented a non-probabilistic counterpart of BHC, referred to as RBHC, using a small variance relaxation
when underlying likelihoods are assumed to be conjugate exponential families.
In contrast to the original BHC, RBHC does not requires careful tuning of hyperparameters.
We have also shown that the dissimilarity measure emerged in RBHC is
reducible with high probability, so that the nearest neighbor chain algorithm was used to
speed up the RBHC and to reduce the space complexity, leading to RBHC-nnca.
Experiments on both synthetic and real-world datasets demonstrated the validity of RBHC.
\medskip
{\bf Acknowledgements}:
This work was supported by National Research Foundation (NRF) of Korea (NRF-2013R1A2A2A01067464)
and the IT R\&D Program of MSIP/IITP (14-824-09-014, Machine Learning Center).
\bibliographystyle{abbrvnat}
|
\section{Rolf Hagedorn}}}
The development of physics is the achievement of physicists, of humans,
persisting against often considerable odds. Even in physics, fashion rather
than fact is frequently what determines the judgement and recognition.
\medskip
When Rolf Hagedorn carried out his main work, now quite generally recognized
as truly pioneering, much of the theoretical community not only ignored it,
but even
considered it to be nonsense. ``Hagedorn ist ein Narr'', he is a fool, was
a summary of many leading German theorists of his time. When in the 1990's
the question was brought up whether he could be proposed for the Max Planck
Medal, the highest honor of the German physics community, even then, when
his achievements were already known world-wide, the answer was still
``proposed, yes...'' At the time Hagedorn carried out his seminal research,
much of theoretical physics was ideologically fixed on ``causality, unitarity,
Poincar\'e invariance'': from these three concepts, from axiomatic quantum
field theory, all that is relevant to physics must arise. Those who thought
that science should progress instead by comparison to
experiment were derogated as ``fitters and plotters''. Galileo was almost
forgotten... Nevertheless, one of the great Austrian theorists of the time,
Walter Thirring, himself probably closer to the fundamentalists, noted:
``If you want to do something really {\sl new}, you first have to have a
{\sl new idea}''. Hagedorn did.
\medskip
He had a number of odds to overcome. He had studied physics in G\"ottingen
under Richard Becker, where he developed a life-long love for thermodynamics.
When he took a position at CERN, shortly after completing his doctorate, it
was to perform calulations for the planning and construction of the proton
synchrotron. When that was finished, he shifted to the study of multihadron
production in proton-proton collisions and to modelling the results of these
reactions. It took a while before various members of the community, including
some of the CERN Theory Division, were willing to accept the significance of
his work. This was not made easier
by Hagedorn's strongly focussed region of interest, but eventually it became
generally recognized that here was someone who, in this perhaps a little
similar to John Bell, was developing truly novel ideas which at first sight
seemed quite specific, but which eventually turned out to have a lasting
impact also on physics well outside its regions of origin.
\medskip
We find that Rolf Hagedorn's work centers on two themes:
\begin{itemize}
\vskip0.2cm
\item{the statistical bootstrap model, a self-similar scheme for the
composition and decay of hadrons and their resonances; for Hagedorn,
these were the ``fireballs''.}
\item{the application of the resulting resonance spectrum in an ideal gas
containing all possible hadrons and hadron resonances, and to the
construction of
hadron production models based on such a thermal input.}
\end{itemize}
We will address these topics in the first two sections, and then turn to
their role both in the thermodynamics of strongly interacting matter and in
the description of hadron production in elementary as well as nuclear
collisions. Our aim here is to provide a general overview of Hagedorn's
scientific achievements; other aspects will be covered in other chapters of
this book. Some of what we will say transcends Hagedorn's life. But then,
to paraphrase Shakespeare, we have come to praise Hagedorn, not to bury
him; we want to show that his ideas are still important and very much alive.
{\large{\section{The Statistical Bootstrap}}}
Around 1950, the world still seemed in order for those looking for the
ultimate constituents of matter in the universe. Dalton's atoms had been
found to be not really {\sl atomos}, indivisible; Rutherford's model of
the atom had made them little planetary systems, with the nucleus as the
sun and the electrons as encircling planets. The nuclei in turn consisted
of positively charged protons and neutral neutrons as the essential mass
carriers. With an equal number of protons and electrons, the resulting
atoms were electrically neutral, and the states obtained by considering
the different possible nucleus compositions reproduced the periodic table
of elements. So for a short time, the Greek dream of obtaining the entire
complex world by combining three simple {\sl elementary} particles
in different ways seemed finally feasible: protons,
neutrons and electrons were the building blocks of our universe.
\medskip
But there were those who rediscovered an old problem, first formulated
by the Roman philosopher Lucretius: if your elementary particles, in our
case the protons and neutrons, have a size and a mass, as both evidently
did, it was natural to ask what they are made of. An obvious way to
find out is to hit them against each other and look at the pieces. And it
turned out that there were lots of fragments, the more the harder the
collision.
But they were not really pieces, since the debris found after a proton-proton
collision still also contained the two initial protons. Moreover, the
additional fragments, mesons and baryons, were in almost all ways as
elementary as protons and neutrons.
The study of such collisions was taken up by more and more laboratories
and at ever higher collision energies. As a consequence, the number of
different ``elementary'' particles grew by leaps and bounds, from tens to
twenties to hundreds. The latest compilation of the Particle Data Group
contains over a thousand.
\medskip
What to do? One approach was in principle obvious: just as Dalton's atoms
could be
constructed from simpler, more elementary constituents, so one had to find
a way of reproducing all the hadrons, the particles formed in strong
interaction collisions, in terms of fewer and more elementary building
blocks. This conceptually straight-forward problem was, however, far from
easy: a simply additive composition was not possible. Nevertheless, in the
late 1960's the quark model appeared, in which three quarks and their
antiquarks were found to produce in a non-Abelian composition all the
observed states and more, predicted and found. Not much later,
quantum chromodynamics (QCD) appeared as the quantum field theory governing
strong interactions; moreover, it kept the quarks inherently confined,
without individual existence: they occurred only as quark-antiquark pair
(a meson) or as a three-quark state (a baryon). And wealth of subsequent
experiments confirmed QCD as the basic theory of strong interactions.
The conventional reductionist approach had triumphed once more.
\medskip
Let us, however, return to the time when physics was confronted by all those
elementary particles, challenging its practioners to find a way out. At this
point, in the mid 1960's, Rolf Hagedorn came up with a truly novel idea
\cite{rh2} - \cite{Hagedorn3}.
He was not so much worried about the specific properties of the particles.
He just imagined that a heavy particle was somehow composed out of lighter
ones, and these again in turn of still lighter ones, and so on, until one
reached the pion as the lightest hadron. And by combining heavy ones, you
would get still heavier ones, again: and so on. The crucial input was that
the composition law should be the same at each stage. Today we call that
self-similarity, and it had been around in various forms for many years.
A particularly elegant formulation was written a hundred years before
Hagedorn by the English mathematician Augustus de Morgan, the first president
of the London Mathematical Society:
\bigskip
{\sl
\centerline{Great fleas have little fleas upon their backs to bite'em,}
\centerline{and little fleas have lesser still, and so ad infinitum.}
\centerline{And the great fleas themselves, in turn, have greater fleas to
go on,}
\centerline{while these again have greater still, and greater still,
and so on.}
}
\bigskip
Hagedorn proposed that ``a fireball consists of fireballs, which in turn
consist of fireballs, and so on...'' The concept later reappeared in various
forms in geometry; in 1915, it led to the celebrated triangle devised by
the Polish mathematician Wac{\l}aw Sierpinski: `` a triangle
consists of triangles, which in turn consist of triangles, and so on...'',
in the words of Hagedorn. Still later, shortly after Hagedorn's proposal,
the French mathematician Benoit Mandelbroit initiated the study
of such {\sl fractal behaviour} as a new field of mathematics.
\begin{figure}[htb]
\hfill{\epsfig{file=sierpinski_dreieck.ps,width=5cm,angle=-90}~
~~~~~~~~~~~~~~~~~~~~~~~~}
\end{figure}
\vskip-3.5cm
~~~~~~~~~~~~~~~~~~~The Sierpinski Triangle
\vskip3cm
Hagedorn had recalled a similar problem in number theory: how many ways are
there of decomposing an integer into integers? This was something already
adressed in 1753 by Leonhard Euler, and more than a century later by
the mathematician E.\ Schr\"oder in Germany. Finally G.\ H.\ Hardy and
S.\ Ramanujan in England provided an asymptotic solution. Let us here,
however, consider a simplified, easily solvable version of the problem
\cite{Blanchard}, in
which we count all possible different ordered arrangements $p(n)$ of an
integer $n$. So we have
\medskip
1=1 \hfill $p(1)=1=2^{n-1}$
2=2 ,1+1 \hfill $p(2)=2 = 2^{n-1}$
3= 3, 2+1, 1+2, 1+1+1 \hfill $p(3) = 4 = 2^{n-1}$
4= 4, 3+1, 1+3, 2+2, 2+1+1, 1+2+1, 1+1+2, 1+1+1+1 \hfill $p(4)=8 = 2^{n-1}$
\medskip
and so on. In other words, there are
\begin{equation}
p(n) = 2^{n-1} = {1 \over 2}~ e^{n \ln 2}
\label{1}
\end{equation}
ways of partitioning an integer $n$ into ordered partitions: $p(n)$ grows
exponentially in $n$. In this particular case, the solution could be found
simply by induction. But there is another way of getting it, more in line
with Hagedorn's thinking: ``large integers consist of smaller integers,
which in turn consist of still smaller integers, and so on...''
This can be formulated as an equation,
\begin{equation}
\rho(n) = \delta(n\!-\!1) + \sum_{k=2}^n~ {1 \over k!}~ \prod_{i=1}^k
~\rho(n_i)~ \delta(\Sigma_i n_i\! -\!n).
\label{2}
\end{equation}
It is quite evident here that the form of the partition number $\rho(n)$
is determined by a convolution of many similar partitions of smaller $n$.
The solution of the equation is in fact just the number of partitions of
$n$ that we had obtained above,
\begin{equation}
\rho(n)= z~p(n)
\label{3}
\end{equation}
up to a normalization constant of order unity (for the present case, it
turns out that $z \simeq 1.25$). For Hagedorn, eq.\ (\ref{2}) expressed
the idea that the structure of $\rho(n)$ was determined by the structure
of $\rho(n)$ -- we now call this self-similar. He instead thought of the
legendary Baron von M\"unchhausen, who had extracted himself from a swamp
by pulling on his own bootstraps. So for him, eq.\ (\ref{2}) became a
{\sl bootstrap equation}.
\medskip
The problem Hagedorn had in mind was, of course, considerably more complex.
His heavy resonance was not simply a sum of lighter ones at rest, but it
was a system of lighter resonances in motion, with the requirement that the
total energy of this system added up to the mass of the heavy one. And
similarly, the masses of the lighter ones were the result of still ligher
ones in motion. The bootstrap equation for such a situation becomes
\begin{equation}
\rho(m,V_0) = \delta(m\!-\!m_0) ~+
\sum_N {1\over N!} \left[ {V_0 \over (2\pi)^3} \right]^{N-1}
\hskip-0.2cm \int \prod_{i=1}^N ~[dm_i~ \rho(m_i)~ d^3p_i]
~\delta^4(\Sigma_i p_i - p),
\label{4}
\end{equation}
where the first term corresponds to the case of just one lightest possible
particle, a ``pion''. The factor $V_0$, the so-called composition volume,
specified the size of the overall system, an intrinsic fireball size.
Since the mass of any resonance in the composition chain is thus determined
by the sum over phase spaces containing lighter ones, whose mass is
specified in the same way, Hagedorn called this form of bootstrap
``statistical''.
\medskip
After a number of numerical attempts by others, W.\ Nahm \cite{Nahm}
solved the statistical
bootstrap equation analytically, obtaining
\begin{equation}
\rho(m,V_0) = {\rm const.}~m^{-3} \exp\{m/T_H\}.
\label{5}
\end{equation}
So even though the partitioning now was not just additive in masses,
but included the kinetic energy of the moving constituents, the increase
was again exponential in mass. The coefficient of the increase,
$T_H^{-1}$, is determined by the equation
\begin{equation}
{V_0 T_H^3 \over 2 \pi^2} (m_0/T_H)^2 K_2(m_0/T_H) = 2 \ln 2 - 1,
\label{6}
\end{equation}
in terms of two parameters $V_0$ and $m_0$. Hagedorn assumed that
the composition volume $V_0$, specifying the intrinsic range of
strong interactions, was determined by the inverse pion mass as
scale, $V_0 \simeq (4 \pi/3) m_{\pi}^{-3}$. This leads to a scale factor
$T_H \simeq 150$ MeV. It should be emphasized, however, that this is
just one possible way to proceed. In the limit $m_0 \to 0$, eq.\ (\ref{6})
gives
\begin{equation}
T_H = [\pi^2 (2\ln 2 -1)]^{1/3}~ V_0^{-1/3} \simeq 1/r_h,
\label{7}
\end{equation}
where $V_0 = (4\pi /3) r_h^3$ and $r_h$ denotes the range of
strong interactions. With $r_h \simeq 1$ fm, we thus have
$T_H \simeq$ 200 MeV. From this it is evident that the exponential
increase persists also in the chiral limit $m_{\pi} \to 0$ and is in fact
only weakly dependent on $m_0$, provided the strong interaction
scale $V_0$ is kept fixed.
\medskip
The weights $\rho(m)$ determine the composition as well as the decay of
``resonances'', of fireballs. The basis of the entire formalism, the
self-similarity postulate -- here in the form of the statistical bootstrap
condition -- results in an unending sequence of ever-heavier fireballs and in
an exponentially growing number of different states of a given mass $m$.
\medskip
Before we turn to the implications of such a pattern in thermodynamics,
we note that not long after Hagedorn's seminal paper, it was found that
a rather different approach, the dual resonance model \cite{DR1,DR2,DR3}
led to very much the
same exponential increase in the number of states. In this model, any
scattering amplitude, from an initial two to a final n hadrons, was assumed
to be determined by the resonance poles in the different kinematic channels.
This resulted structurally again in a partition problem of the same type, and
again the solution was that the number of possible resonance states of mass
$m$ must grow exponentially in $m$, with an inverse scale factor of the
same size as obtained above, some 200 MeV. Needless to say, this unexpected
support from the forefront of theoretical hadron dynamics considerably
enhanced the interest in Hagedorn's work.
{\large{\section{The Limiting Temperature of Hadronic Matter}}}
Consider a relativistic ideal gas of identical neutral scalar particles of
mass $m_0$ contained in a box of volume $V$, assuming Boltzmann statistics.
The grand canonical partition function of this system is given by
\begin{equation}
{\cal Z}(T,V) = \sum_N {1 \over N!} \left[ {V \over (2\pi)^3} \int d^3p~
\exp\{-\sqrt{p^2+m_0^2}~/T\} \right]^N,
\end{equation}
\noindent
leading to
\begin{equation}
\ln {\cal Z}(T,V) = {VTm_0^2 \over 2\pi^2}~ K_2({m_0\over T}).
\end{equation}
\medskip
\noindent
For temperatures $T\gg m_0$, the energy density of the system becomes
\begin{equation}
\rm e(T) = - {1 \over V}~ {\partial~\!\ln~\!{\cal Z}(T,V) \over \partial~\!(1/T)}
\simeq {3 \over \pi^2}~ T^4,
\end{equation}
\medskip
\noindent
the particle density
\begin{equation}
n(T) = {\partial~\!\ln~\!{\cal Z}(T,V) \over \partial~\!V}
\simeq {1 \over \pi^2}~ T^3,
\end{equation}
\bigskip
\noindent
and so the average energy per particle is given by
\begin{equation}
\omega \simeq 3~T.
\end{equation}
\noindent
The important feature to learn from these relations is that, in the case
of an ideal gas of one species of elementary particles,
an increase of the energy of the system has three consequences:
it leads to
\vskip0.5cm
\begin{itemize}
\vspace*{-0.3cm}
\item{a higher temperature,}
\vspace*{-0.3cm}
\item{more constituents, and}
\vspace*{-0.3cm}
\item{more energetic constituents.}
\vspace*{-0.3cm}
\end{itemize}
\medskip
If we now consider an {\sl interacting} gas of such basic
hadrons and postulate that the essential form of the interaction
is resonance formation, then we can approximate the interacting medium as a
non-interacting gas of all possible resonance species \cite{B-U,DMB}.
The partition function of this resonance gas is
\begin{equation}
\ln {\cal Z}(T,V) = \sum_i
{VTm_i^2 \over 2\pi^2}~\rho(m_i)~ K_2({m_i\over T})
\label{resgas}
\end{equation}
\medskip
where the sum begins with the stable ground state $m_0$ and
then includes the possible resonances $m_i, i=1,2,...$ with
weights $\rho(m_i)$ relative to $m_0$.
Clearly the crucial question here is how to specify $\rho(m_i)$,
how many states there are of mass $m_i$. It is only at this point
that hadron dynamics enters, and it is here that Hagedorn introduced
the result obtained in his statistical bootstrap model.
\medskip
As we had seen above in eq.\ (\ref{5}), the density of states then
increases exponentially in $m$, with a coefficient $T_H^{-1}$ determined
by eq.\ (\ref{6}) in terms of two parameters $V_0$ and $m_0$.
If we replace the sum in the resonance gas partition function
(\ref{resgas}) by an integral and insert the exponentially growing mass
spectrum (\ref{5}), eq.\ (\ref{resgas}) becomes
$$
\ln {\cal Z}(T,V) \simeq
{V T \over 2\pi^2} \int dm~m^2 \rho(m_i)~ K_2({m_i\over T})
$$
\begin{equation}
\sim V\left[{T \over 2\pi}\right]^{3/2} \int dm~m^{-{3/2}}
\exp\{-m \left[{1\over T} - {1\over T_H}\right]\}.
\label{div}
\end{equation}
\medskip
\noindent
Evidently, the result is divergent for all $T > T_H$: in other words,
$T_H$ is the highest posssible temperature of hadronic matter. Moreover,
if we compare such a system with the ideal gas of only basic particles
( a ``pion'' gas), we find
\medskip
\hspace*{2.9cm}
pion gas \hskip 6cm resonance gas
\medskip
\hskip 3cm
$ n_{\pi} \sim \rm e^{3/4} \hskip 6cm n_{res} \sim \rm e$
\hskip 3cm
$ \omega_{\pi} \sim \rm e^{1/4} \hskip 6cm \omega_{res} \sim {\rm const.}$
\medskip
\noindent
Here $n$ denotes the average number density of constituents, $\omega$ the
average energy of a constituent. In contrast to to the pion gas,
an increase of energy now leads to
\begin{itemize}
\vspace*{-0.2cm}
\item{a fixed temperature limit, $T \to T_H$,}
\vspace*{-0.3cm}
\item{the momenta of the constituents do not continue to increase, and}
\vspace*{-0.3cm}
\item{more and more species of ever heavier particles appear.}
\vspace*{-0.2cm}
\end{itemize}
We thus obtain a new, non-kinetic way to use energy, increasing the
number of species and their masses, not the momentum per particle.
Temperature is a measure of the momentum of the constituents, and if
that cannot continue to increase, there is a highest possible, a ``limiting''
temperature for hadronic systems.
\medskip
Hagedorn originally interpreted $T_H$ as the ultimate
temperature of strongly interacting matter. It is clear today that $T_H$
signals the transition from hadronic matter to a quark-gluon plasma.
Hadron physics alone can only specify its inherent limit; to go beyond
this limit, we need more information: we need QCD.
\medskip
As seen in eq.\ (\ref{5}), the solution of the statistical bootstrap
equation has the general form
\begin{equation}
\rho(m,V_0) \sim m^{-a} \exp\{m/T_H\},
\label{boot}
\end{equation}
with some constant $a$; the exact solution of eq.\ (\ref{4}) by Nahm gave
$a=3$. It is possible, however, to consider variations of the bootstrap
model which lead to different $a$, but always retain the exponential
increase in $m$. While the exponential form makes $T_H$ the upper limit
of permissible temperatures, the power law coefficient $a$ determines
the behavior of the system at $T=T_H$. For $a=3$, the partition function
(\ref{div}) itself exists at that point, while the energy density as first
derivative in temperature diverges there. This is what made Hagedorn
conclude that $T_H$ is indeed the highest possible temperature of matter:
it would require an infinite energy to reach it.
\medskip
Only a few years later it was, however, pointed out by N.\ Cabibbo and
G.\ Parisi \cite{C-P} that larger $a$ shifted the divergence at $T=T_H$ to
ever higher derivatives. In particular, for $4> a > 3$, the energy
density would remain finite at that point, shifting the divergence
to the specific heat as next higher derivative. Such critical behavior was
in fact
quite conventional in thermodynamics: it signalled a phase transition leading
to the onset of a new state of matter. By that time, the quark model and
quantum chromodynamics as fundamental theory of strong interactions
had appeared and suggested the existence of a quark-gluon plasma as the
relevant state of matter at extreme temperature or density. It was
therefore natural to interpret the Hagedorn temperature $T_H$ as the
critical transition temperature from hadronic matter to such a plasma. This
interpretation is moreover corroborated by a calculation of the
critical exponents \cite{HScrit} governing the singular behavior of the
resonance gas thermodynamics based on a spectrum of the form (\ref{boot}).
\medskip
It should be noted, however, that in some sense $T_H$ did remain the
highest possible temperature of matter as we know it. Our matter exists in
the physical vacuum and is constructed out of fundamental building blocks
which in turn have an independent existence in this vacuum. Our matter
ultimately consists of and can be broken up into nucleons; we can
isolate and study a single
nucleon.
The quark-gluon plasma, on the other hand, has its own ground state, distinct
from the physical vacuum, and its constituents can exist only in a dense
medium of other quarks -- we can never isolate and study a single quark.
\medskip
That does not mean, however, that quarks are eternally confined to a given
part of space. Let us start with atomic matter and compress that to form
nuclear matter, as it exists in heavy nuclei. At this stage, we have nucleons
existing in the physical vacuum. Each nucleon consists of three quarks, and
they are confined to remain close to each other; there is no way to break
up a given nucleon into its quark constituents. But if we now continue to
compress, then eventually the nucleons will then penetrate each other,
until we reach a dense medium of quarks. Now each quark finds in its
immediate neighborhood
many other quarks besides those which were with it in the nucleon stage.
It is therefore no longer possible to partition quarks into nucleons;
the medium consists of unbound quarks, whose interaction becomes ever weaker
with increasing density, approaching the limit of asymptotic freedom
predicted by QCD. Any quark can now move freely throughout the medium:
we have quark liberation through swarm formation. Wherever a quark
goes, there are many other quarks nearby. The transition from
atomic to quark matter is schematically illustrated in Fig.\ \ref{states}.
\begin{figure}[h]
\centerline{\psfig{file=states.eps,width=11cm}}
\caption{Schematic view of matter for increasing density, from atomic (a) to
nuclear (b) and then to quark matter (c).}
\label{states}
\end{figure}
\medskip
We have here considered quark matter formation through the compression of
cold nuclear matter. A similar effect is obtained if we heat a meson gas;
with increasing temperature, collisions and pair production lead to an
ever denser medium of mesons. And according to Hagedorn, also of ever
heavier mesons of an increasing degeneracy. For Hagedorn, the fireballs
where pointlike, so that the overlap we had just noted simply does not
occur. In the real world, however, they do have hadronic size, so that
they will in fact interpenetrate and overlap before the divergence of
the Hagedorn resonance gas occurs \cite{Zinov}. Hence now again there
will be a transition from resonance gas to a quark-gluon plasma, now formed
by the liberation of the quarks and gluons making up the resonances.
\medskip
At this point, it seems worthwhile to note an even earlier approach leading
to a limiting temperature for hadronic matter. More than a decade before
Hagedorn, I.\ Ya.\ Pomeranchuk \cite{Pom} had pointed out that a crucial
feature of hadrons is their size, and hence the density of any hadronic
medium is limited by volume restriction: each hadron must have its own volume
to exist, and once the density reaches the dense packing limit, it's the end
for hadronic matter. This simply led to a temperature limit, and for an ideal
gas of pions of 1 fm radius, the resulting temperature was again around
200 MeV. Nevertheless, these early results remained largely unnoticed until
the work of Hagedorn.
\medskip
Such geometric considerations do, however, lead even further. If hadrons are
allowed to interpenetrate, to overlap, then percolation theory predicts
two different states of matter \cite{Perc1,Perc2}: hadronic matter, consisting
of isolated hadrons or finite hadronic clusters, and a medium formed as an
infinite sized cluster of overlapping hadrons. The transition from one to
the other now becomes a genuine critical phenomenon, occurring at a critical
value of the hadron density.
\medskip
We thus conclude that the pioneering work of Rolf Hagedorn opened up the
field of critical behavior in strong interaction physics, a field in which
still today much is determined by his ideas. On a more theoretical level,
the continuation of such studies was provided by finite temperature lattice
QCD, and on the more experimental side, by resonance gas analyses of
the hadron abundances in high energy collisions. In both cases, it was
found that the observed behavior was essentially that predicted by
Hagedorn's ideas.
{\large{\section{Resonance gas and QCD thermodynamics}}}
With the formulation of Quantum Chromodynamics (QCD) as a
theoretical framework for the strong interaction force among
elementary particles it became clear that the appearance of the ultimate
Hagedorn temperature $T_H$ signals indeed the transition from hadronic phase
to a new phase of strongly interacting matter, the quark-gluon plasma
(QGP) \cite{HS}.
As QCD is an asymptotically free theory, the interaction between quarks and
gluons vanishes
logarithmically with increasing temperature, thus at very high
temperatures the QGP effectively
behaves like an ideal gas of quarks and gluons.
\medskip
Today we have detailed information, obtained from numerical calculations
in the framework of finite temperature lattice Quantum Chromodynamics
\cite{first1,first2,first3,review,fodor}, about the thermodynamics of
hot and dense
matter. We know the transition temperature to the QGP and the temperature
dependence of basic bulk thermodynamic observables such as the
energy density and the pressure \cite{fk,fodor}. We also begin to have results
on fluctuations and correlations of conserved charges
\cite{C1,C2,C3}.
\medskip
The recent increase in numerical accuracy of lattice QCD calculations
and their extrapolation to the continuum limit make it possible to
confront the fundamental results of QCD with Hagedorn's
concepts \cite{rh2, Hagedorn3}, which provide a theoretical scenario
for the thermodynamics of strongly interacting
hadronic matter
\cite{C3,taw1,taw2,frit}.
\medskip
In particular, the
equation of state calculated on the lattice at vanishing and finite
chemical potential, and restricted to the confined hadronic phase, can be
directly compared to that obtained from the partition function (\ref{div})
of the hadron resonance gas, using the form (\ref{boot}) introduced by
Hagedorn for a continuum mass spectrum.
Alternatively, as first approximation, one can also consider a discrete
mass spectrum which accounts for all experimentally known hadrons and
resonances. In this case the continuum partition function of the Hagedorn
model is expressed by Eq.\ (\ref{resgas}) with $\rho(m_i)$ replaced by the
spin
degeneracy factor of the $i^{th}$ hadron, and with the summation taken over
all known resonance species listed by
the Particle Data Group \cite{pdg}.
\medskip
With the above assumption on the dynamics and the mass spectrum,
the Hagedorn resonance gas partition function \cite{rh2, Hagedorn3}
can be calculated exactly and expressed as a sum of
one--particle partition functions $Z^1_i$ of all hadrons and
resonances,
\begin{equation}
\ln Z(T,V)=\sum_i Z_i^1(T,V).
\label{qq1}
\end{equation}
For particles of mass $m_i$ and spin
degeneracy factor $g_i$, the one--particle partition function
$Z^1_i$, in the Boltzmann approximation, reads
\begin{equation} Z^1_i(T,V)= g_i{{VTm_i^2}\over {2\pi^2}} K_2({{m_i}\over T}). \label{eeq2} \end{equation}
Due to the factorization of the partition function in
Eq.~(\ref{qq1}), the energy density and the pressure of the Hagedorn
resonance gas with a discrete mass spectrum,
can also be expressed as a sum over single particle contributions
\begin{equation}
\epsilon=\sum_i\epsilon_i^1~~,~~P=\sum_i P_i^1, \label{qq3}
\end{equation}
with
\begin{eqnarray}
{{\epsilon_i^1}\over {T^4}} &=& \frac{g_i}{2\pi^2}\;
(\frac{ m_i}{T})^3
\;\left[\frac{3\;K_2(\beta m_i)}{\beta m} +
\;K_1(\beta m_i)\right]\label{eqq4} \label{eqq5}
\end{eqnarray}
and
\begin{eqnarray}
\frac{P_i^1}{T^4} &=& \frac{g_i}{2
\pi^2}\;(\frac{m_i}{T})^3 \; K_2(\beta m_i) \label{eqq6},
\end{eqnarray}
where $\beta =1/T$ and $K_1$ and $K_2$ are modified Bessel functions.
At vanishing chemical potentials and at finite temperature, the energy density
$\epsilon$, the entropy density $s$ and the pressure $P$,
are connected through the thermodynamic relation,
\begin{eqnarray}
\epsilon=-P+sT.\label{eq6}
\end{eqnarray}
Summing up in Eq.~(\ref{qq3}) the contributions from
experimentally known hadronic states, constitutes the resonance gas
\cite{rh2, Hagedorn3,redb} for the
thermodynamics of the hadronic phase of QCD.
Taking into account contributions of all mesonic and baryonic resonances
with masses up to 1.8 GeV and 2.0 GeV, respectively, amounts to 1026
resonances.
\medskip
\begin{figure}[!t]
\center\includegraphics[width=3.50in,height=2.3in]{hotqcd_eos.eps}\hskip 0.9cm
\vskip -0.1cm \caption{The
normalized pressure $P(T)$, the energy density $\epsilon(T)$ and the entropy
density $s(T)$ obtained in lattice QCD calculations as a function of
temperature. The dark lines show the prediction of the Hagedorn resonance gas
for discrete mass spectrum, eqs.~(\ref{qq3}) - (\ref{eq6}). The
lattice results are from Ref.\ \cite{fk}.}\label{fig1}
\end{figure}
\medskip
The crucial question thus is if the equation of state
introduced by Hagedorn can describe the
corresponding results obtained from QCD within lattice approach. In
Fig.~\ref{fig1} we show the temperature dependence of the energy
density, pressure and the entropy density obtained recently in
lattice QCD studies with physical masses of up, down and strange
quarks \cite{fk}. The bands in lattice QCD results indicate error bars due
to extrapolation to the continuum limit.
The vertical band marks the temperature,
$T_c = (154 \pm 9)$ MeV, which within error,
is the crossover temperature from a hadronic phase to a quark-gluon
plasma \cite{tc}.
These QCD results are compared in Fig.\ \ref{fig1} to
Hagedorn's resonance gas model (HRG)
formulated for a discrete mass spectrum in
Eqs.~(\ref{qq3}) and (\ref{eq6}).
There is excellent agreement between the Hagedorn model results for the
equation of states and the corresponding lattice data.
All bulk thermodynamical observables are very strongly
changing with temperature when approaching the deconfinement
transition temperature. In Hagedorn's formulation, this behavior is
well understood in terms of increasing resonance contributions.
Although the HRG formulated for discrete mass spectrum does not
exhibit any critical
behavior, it nevertheless reproduces remarkably well the lattice
results in the hadronic phase. This agreement has now been extended to
an analysis of
fluctuations and correlations of conserved charges as well.
\medskip
The excellent description of the lattice QCD results by
Hagedorn's model
justifies the claim {\sl that resonances are indeed the
essential degrees of freedom near deconfinement}. Thus,
on the thermodynamical level, modeling hadronic interactions
by formation and excitation of resonances, as introduced by Hagedorn,
is an excellent approximation of strong interactions.
\bigskip\bigskip
{\large{\section{Resonance Gas and Heavy Ion Collisions}}}
Long before lattice QCD could provide a direct evidence that strong
interaction thermodynamics can be quantified by the resonance
gas partition function, Hagedorn's concept was verified phenomenologically
by considering particle production in elementary and heavy ion collisions
\cite{jc,redf,heppe,beca,redb}.
In a strongly interacting medium, one includes the conservation of electric
charge, baryon number and strangeness. In this case, the partition function
of Hagedorn's thermal model depends not only on temperature but also on
chemical
potential $\vec\mu$, which guarantees, that charges are conserved on an
average.
For a non vanishing $\vec\mu$, the partition function Eq. (\ref{qq1})
is replaced by
\begin{equation}
\ln Z(T,V,\vec\mu )=\sum_i Z_i^1(T,V,\vec\mu),
\label{par}
\end{equation}
with $\vec \mu =(\mu_B,\mu_S,\mu_Q)$, where $\mu_i$ are the chemical
potentials
related to the baryon number, strangeness and electric charge conservation,
respectively.
\medskip
For particle $i$ carrying stran\-geness $S_i$, the baryon number $B_i$, the
electric
charge $Q_i$ and the spin--isospin degeneracy factor $g_i$, the one particle
partition function, reads
\begin{equation}
Z_i^1(T,V,\vec\mu) =\frac{ Vg_iTm_i^2 }{ 2\pi^2 }
K_2({{m_i}/T}) \exp \left(
\frac{B_i\mu_B+S_i\mu_S+Q_i\mu_Q}{T} \right).
\label{parp}
\end{equation}
For $\vec\mu=0$ one recovers the result from Eq. (\ref{eeq2}).
\medskip
The calculation of a density $n_i$ of particle $i$ from the partition
function Eq. (\ref{par} )
is rather straightforward \cite{hagedornred}. It amounts to the replacement
$Z_i^1\to \gamma_i Z_i^1$ in Eq. (\ref{par}) and taking a derivative with
respect
to
the particle fugacity $\gamma_i$, as
\begin{equation}
n_i=\frac{ \langle N_i\rangle^\mathrm{th} }{V}=\left.\frac{\partial \ln Z}
{\partial \gamma_i}\right|_{\gamma_i=1},\label{denj}
\end{equation}
consequently, $n_i=Z_i^1/V$ with $Z_i^1$ as in Eq. (\ref{parp}).
\medskip
The Hagedorn model, formulated in Eq. (\ref{par}), describes bulk
thermodynamic properties
and particle composition of a thermal fireball at finite temperature and at
non
vanishing charge densities. If such a fireball is created in high energy
heavy
ion collisions, then yields of different hadron species are fully quantified
by thermal parameters. However, following Hagedorn's idea, the
contribution
of resonances decaying into lighter particles, must be included~
\cite{rh2,Hagedorn3}.
\medskip
In Hagedorn's thermal model, the average number $\langle N_i\rangle$
of particles $i$ in volume $V$ and at temperature $T$ that carries
strangeness
$S_i$, the baryon number $B_i$, and the electric charge $Q_i$, is obtained
from
Eq. (\ref{par}), see~\cite{rh2,Hagedorn3}
\begin{equation}
\langle N_i^{}\rangle (T,\vec\mu)~=~\langle N_i\rangle^\mathrm{th} (T,\vec\mu)
+{\sum}_j\Gamma_{j\to i}
\langle N_j\rangle^{th,R}(T,\vec\mu)\label{denn} .
\end{equation}
The first term in Eq. (\ref{denn}) describes the thermal average number of
particles
of species $i$ from Eq. (\ref{denj}) and the second term describes overall
contribution
from resonances. This term is taken as a sum of all resonances that decay
into
particle $i$. The $\Gamma_{j\to i}$ is the corresponding decay branching ratio
of
$j\to i$. The multiplicities of resonances $\langle N_j\rangle^{th,R}$ in
Eq. (\ref{denn}), are obtained from Eq. (\ref{denj}).
\medskip
The importance of resonance contributions to the total particle yield in
Eq. (\ref{denn})
is illustrated in Fig. (\ref{fig2}) for charge pions. In Fig. (\ref{fig2})
we show the
ratio of the
total number of charge pions from Eq. (\ref{denn}) and the number of
prompt
pions from
Eq. (\ref{denj}). The ratio is strongly increasing with temperature and
chemical
potential.
This is due to an increasing contribution of mesonic and baryonic
resonances. From
Eq. (\ref{fig2}) it is clear, that at high temperature and/or density,
the
overall
multiplicity of pions is mostly due to resonance decays.
\medskip
The particle yields in Hagedorn's model Eq. (\ref{denn}) depend, in general,
on five parameters. However, in high energy heavy ion collisions, only three
parameters are independent. The isospin asymmetry, in the initial state fixes
the charge chemical potential and the strangeness neutrality condition
eliminates
the strange chemical potential. Thus, on the level of particle multiplicity, we
are
left with temperature $T$ and the baryon chemical potential $\mu_B$ as
independent
parameters, as well as, with fireball volume as an overall normalization
factor.
\medskip
\begin{figure
\centerline{\psfig{file=RES.ps, width=3.50in, height=2.9in, angle=180}}
\vspace*{-1.3cm}
\caption{The ratio of the total density of positively charged pions,
$n_{\pi^+}^R$ from Eq. (\ref{denn}), and the density of thermal pions,
$n_{\pi^+}^\mathrm{th}$ from Eq. (\ref{denj}). The calculations are done in the
Hagedorn
resonance gas model for $\mu_B=$250 MeV and $\mu_B=$550 MeV at different
temperatures. }\label{fig2}
\end{figure}
Hagedorn's thermal model introduced in Eq. (\ref{denn}) was successfully
applied to describe particle yields measured in heavy ion collisions.
The model was compared with available experimental data obtained in a broad
energy range from AGS up to LHC. Hadron multiplicities ranging from pions
to omega baryons and their ratios, as well as composite objects like e.g.
deuteron or alpha particle, were used to verify if there was a set of
thermal parameters $(T,\mu_B)$ and $V$, which simultaneously reproduces
all measured yields.
\medskip
The systematic studies of particle production extended over more than
two decades, using experimental results at different beam energies,
have revealed a clear justification, that in central heavy ion collisions
particle yields are indeed consistent with the expectation of the Hagedorn
thermal model. There is also a clear pattern of the energy,
$\sqrt s$-dependence of thermal parameters. The temperature is
increasing with $\sqrt s$, and at the SPS energy essentially saturates
at the value, which corresponds to the transition temperature from a
hadronic phase to a QGP, as obtained in LQCD. The chemical potential,
on the other hand, is gradually decreasing with $\sqrt s$ and almost
vanishes at the LHC.
\medskip
\begin{figure
\centerline{\psfig{file=yield.ps, width=3.50in, height=2.2in}}
\caption{Yields of several different particle species per unit rapidity
normalized to spin degeneracy factor as a function of their mass.
Data are from ALICE collaboration taken at the LHC in central Pb-Pb
collisions. The line is the Hagedorn thermal model result, Eq.
(\ref{raf}), see Ref.\cite{andronic1,andronic2}.}\label{fig3}
\end{figure}
In Fig. (\ref{fig3}) we show, as an illustration, a comparison of
Hagedorn's thermal model and recent data on selected particle yields,
obtained by ALICE collaboration in central Pb-Pb collisions at midrapidity
at the LHC energy~\cite{andronic1,andronic2}. At such high collision energy,
particle yields from Eq. (\ref{denn}) are quantified entirely by the
temperature and the fireball volume.\footnote{The chemical potential
$\vec\mu$ in Eq. (\ref{denn}) vanishes, since at the LHC and at
midrapidity particles and their antiparticles are produced symmetrically.}
Thus, there is transparent prediction of Hagedorn's model
Eq. (\ref{denn}), that yields of heavier particles $\langle N_{i}\rangle$
with no resonance decay contributions, normalized to their spin degeneracy
factor $g_i=(2J+1)$, should be quantified by
\begin{equation}
\frac{ \langle N_{i}\rangle}{ 2J+1 }\simeq VT^3 \left(\frac{{m_i}}{ 2\pi T}
\right)^{3/2}
\exp( -{{m_i}/ T}),
\label{raf}
\end{equation}
where we have used Eq. (\ref{denj}) and the asymptotic expansion of the
Bessel function, $K_2(x)\sim x^{-1/2}\exp(-x)$, valid for large $x$.
\medskip
In Fig. (\ref{fig3}) we show the yields of particles with no resonance
contribution, like $\phi$, $\Omega$, the deuteron `d', $^3$He and the
hypertriton $_\Lambda^3$He, normalized to their spin degeneracy factor,
as a function of particle mass. Also shown in this figure is the prediction
from Eq. (\ref{raf}) at $T\simeq 156$ and for volume
$V\simeq 5000$ fm$^3$~\cite{andronic1,andronic2}. There is a clear coincidence
of data taken in Pb-Pb collisions at the LHC and predictions of the
Hagedorn model Eq. (\ref{raf}). Particles with no resonance contribution
measured by ALICE collaboration follow the Hagedorn's expectations that
they are produced from a thermal fireball at common temperature.
A similar agreement of Hagedorn's thermal concept and experimental
data taken in central heavy ion collisions has been found for different
yields of measured particles and collision energies from AGS, SPS, RHIC
and LHC~\cite{redb}.
\bigskip\bigskip
{\large{\section{Particle Yields and Canonical Charge Conservation}}}
The Hagedorn thermodynamical model for particle production,
was originally applied to quantify and understand particle
yields and spectra measured in elementary collisions --
there were no data available from heavy ion collisions.
\medskip
Initial work on particle production by Hagedorn begins in 1957
in collaboration with F. Cerulus when they apply the Fermi
phase space model. In this
microcanonical approach, conservation laws of baryon number
or electric charge are implemented exactly. Almost 15 years
later the production of complex light antinuclei, such as \medskip
anti-He$^3$, preoccupied Hagedorn~\cite{rh2,Hagedorn3}.
He realized and discussed clearly the need to find a path
to enforce exact conservation of baryon number to describe
the anti-He$^3$ production correctly within the canonical
statistical formulation.
\medskip
Indeed, applying in \cal pp ~reactions the thermal model without
concern for conservation of baryon number overestimates the
production of anti-He$^3$ in proton-proton collisions by
seven orders of magnitude \cite{rh2,rh4,Hagedorn3}. The reason
was that when the number of particles in the interaction volume
is small, one has to take into account the fact that the
production of anti-He$^3$ must be accompanied by the production
of another three nucleons with energy $E_N$, in order to exactly
conserve the baryon number. Thus, in case that the production
of anti-He$^3$ is not originating from reservoir of many
antiquarks or antinucleons already present in a large volume,
but is rather originating from some small volume $V_{pp}$
that is present in pp~ collisions, the abundance of anti-He$^3$
will not be proportional to the single standard Boltzmann factor,
as in Eq. (\ref{raf})
\begin{equation}
n_{\overline{\,\mathrm{He}}^3} \sim
\exp \left( -{m_{\,\overline{\mathrm{He}}^3}/T} \right),
\end{equation}
but is accompanied by additional Boltzmann factors that characterize
the production of the associated nucleons, needed in order to conserve
baryon number~\cite{rh2,Hagedorn3}
\begin{equation}
n_{\,\overline{\,\mathrm{He}}^3} \sim
\exp \left( -{m_{\overline{\,\mathrm{He}}^3}/ T} \right)
\left[ V_{\rm pp} \int \frac{d^3p}{ (2\pi)^3}
\exp \left( -\frac{E_N}{ T}\right)\right]^3.
\end{equation}
This suppresses the rate and introduces a strong power-law
dependence on volume $V_{pp}$ for the anti-He$^3$ yield.
\medskip
The problem of exact conservation of discrete quantum numbers in a
thermal model formulated in early '70s by Hagedorn in the context of
baryon number conservation remained unsolved for a decade. When the
heavy ion QGP research program was approaching and strangeness emerged
as a potential QGP signature, Hagedorn pointed out the need to consider
exact conservation of strangeness \cite{privat}. This is the reason
that the old problem of baryon number conservation was solved in the new
context of strangeness conservation ~\cite{rafeldan,tounsi,oeschler}.
A more general solution, applicable to {\em all} discrete conserved
charges, abelian and non-abelian, was also introduced in
Ref. ~\cite{turko1} and expanded in ~\cite{turko2,petrov,rafm,
hagedornred,Derreth,Elze}.
Recently, it has become clear that a similar treatment should be
followed not only for strangeness but also for charm abundance
study in high energy $e^+e^-$ collisions \cite{s1,b1}.
\medskip
To summarize this section, we note that the usual form of the
statistical model, based on a grand canonical formulation of the
conservation laws, cannot be used when either the temperature or the
volume or both are small. As a rough estimate, one needs $V T^3 > 1$
for a grand canonical description to hold \cite{rafeldan,hagedornred}.
In the opposite limit, a path was found within the canonical ensemble
to enforce charge conservations exactly.
\medskip
The canonical approach has been shown to provide a consistent
description of particle production in high energy hadron-hadron,
$e^+e^-$ and peripheral heavy ion collisions \cite{redb,rafm,s1,b1}.
In the context of developing strangeness as signature of QGP, such a model also provides, within
the realm of assumed strangeness chemical equilibrium, a description
of an observed increase of single- and multi-strange particle yields
from \cal pp,~ \cal pA ~to \cal AA ~collisions and its energy dependence \cite{tounsi}.
\bigskip\bigskip
{\large{\section{Concluding Remarks}}}
Rolf Hagedorn's work, introducing concepts from statistical mechanics
and from the mathematics of self-similarity into the analysis of high
energy multiparticle production, started a new field of research, alive
and active until today. On the theory side, the limiting temperature of
hadronic matter and the behavior of the Hagedorn resonance gas approaching
that limit were subsequently verified by first principle calculations
in finite temperature QCD. On the experimental side, particle yields as
well as, more recently, fluctuations of conserved quantities, were also
found to follow the pattern predicted by the Hagedorn resonance gas.
Rarely has an idea in physics risen from such humble and little
appreciated beginnings to such a striking vindication. So perhaps
it is appropriate to close with a poetic summary one of us (HS)
formulated some twenty years ago for a Hagedorn-Fest, with a slight update.
\vskip0.5cm
\begin{center}
\bigskip
{\bf Hot Hadronic Matter}
\medskip
{\sl (A Poetic Summary)}
\bigskip
In days of old
a tale was told
of hadrons ever fatter.
Behold, my friends, said Hagedorn,
the ultimate of matter.
\bigskip
Then Muster Mark
called in the quarks,
to hadrons they were mated.
Of colors three, and never free,
all to confinement fated.
\bigskip
But in dense matter,
their bonds can shatter
and they can freely move around.
Above $T_H$, their colors shine,
as the QGP is found.
\bigskip
Said Hagedorn,
when quarks were born
they had different advances.
Today they form, as we can see,
a gas of all their chances.
\end{center}
\vskip1cm
\centerline{\bf Acknowledgement:}
\medskip
K.R. acknowledges
support by the Polish Science Foundation (NCN), under
Maestro grant DEC-2013/10/A/ST2/00106.
|
\section{\label{sec:level1}Introduction}
Since the discovery of M\"ossbauer effect spectroscopy more than half a century ago \cite{ADD1_ZP1958}, superconductivity has been one of the states that has been investigated by this technique \cite{ADD3_NC1961}. Although M\"ossbauer spectroscopy is widely accepted as one of the most sensitive techniques in terms of energy resolution, it has not contributed significant insight to studies of conventional superconductors. After the discovery of cuprate high temperature superconductors (HTSC), M\"ossbauer spectroscopy was widely used for studies of these materials and reports of observation of some anomalies in the spectral parameters in the vicinity of the superconducting critical temperature (T$_c$) \cite{1PRL1991,2PRB1992,3PRB1988,4JPCM1993,5JPCM1994} were published. Due to the absence of commonly available M\"ossbauer nuclides in the cuprates, most studies were accomplished either by partial substitution of copper atoms by $^{57}$Fe and/or $^{119}$Sn, or by using resonant isotopes of the rare earth metals, like $^{151}$Eu, which increases the degree of difficulty of the measurements and reduces the clarity of the results \cite{1PRL1991,3PRB1988,6SSC1987,7PC1999}. Recently, the discovery of iron-based superconductors, that naturally contain the common M\"ossbauer nuclide, $^{57}$Fe, has triggered intense M\"ossbauer studies of these superconductors \cite{8PC2009,9PRB2011,10EPL2014,11PRB2011,12HI2013,13HI2012}. Superconductivity in iron-pnictides is usually achieved by doping a magnetic parent compound with electrons or holes, or by application of chemical or physical pressure, and thereby, suppressing the magnetic order, suggesting that superconductivity and magnetism are closely related in this system. Although there are some studies on iron-based superconductors, which state that M\"ossbauer spectral parameters show anomalies near T$_c$, in these materials it is hard to attribute the variation of the hyperfine parameters observed by M\"ossbauer spectroscopy to purely magnetic or purely superconducting origins \cite{10EPL2014,12HI2013,13HI2012,14PRB2013}.
In order to study possible variations of hyperfine parameters caused by the transition from the normal to the superconducting state in a conventional superconductor, we revisited lutetium-iron-silicide, Lu$_2$Fe$_3$Si$_5$. Lu$_2$Fe$_3$Si$_5$ is a stoichiometric, Fe-containing, superconductor with relatively high T$_c \approx$ 6 K \cite{15PL1980}. In a previous M\"ossbauer study \cite{16JMMM1981}, the non-magnetic nature of Fe was already confirmed. Hence, Lu$_2$Fe$_3$Si$_5$ can be considered as an ideal compound to investigate the variation of hyperfine parameters caused only by the superconducting transition without any complications associated by the absence of M\"ossbauer nucleus and/or the presence of magnetism. To the best of our knowledge no detailed, temperature dependent, $^{57}$Fe M\"ossbauer spectroscopy measurements were performed on this material so far, and our goal is to shed some light on the applicability of M\"ossbauer spectroscopy for studies of conventional, albeit multigap superconductors \cite{ADD2_PC2012}.
\section{Experimental details}
Polycrystalline samples of Lu$_2$Fe$_3$Si$_5$ were prepared by arc melting constituent elements with the nominal composition of Lu$_{2}$Fe$_{3.32}$Si$_{5.26}$ (corresponding to Lu$_2$Fe$_3$Si$_5$ + Fe$_{0.32}$Si$_{0.26}$) in Zr-gettered Ar atmosphere. Extra iron was added to suppress the formation of a Lu$_2$FeSi$_4$ second phase and extra silicon was added to compensate apparent loss during the arc melting. To ensure the homogeneity of the sample, the arc melting was repeated iteratively after flipping of the melted and resolidified ingot, for more than ten times. The weight loss was about 0.26\%. The arc-melted ingot was then sealed in an amorphous silica tube, under a partial pressure of argon, and annealed at 1050 \textcelsius \ for 12 days.
Powder X-ray diffraction (XRD) was performed using a Rigaku Miniflex diffractometer with Cu K$\alpha$ radiation at room temperature (RT). The powder X-ray spectra of the samples were refined by Rietveld analysis using the EXPGUI software \cite{17EXPGUI}. Magnetic measurements were performed using a Quantum Design Magnetic Property Measurement System SQUID magnetometer, specific heat capacity was measured in a Quantum Design Physical Property Measurement System.
M\"ossbauer spectroscopy measurements were performed using a SEE Co. conventional constant acceleration type spectrometer in transmission geometry with an $^{57}$Co(Rh) source, which had an initial intensity 50 mCi, kept at RT. The absorber was prepared in a powder form (10 mg of natural Fe/cm$^2$) by grinding of approximately 175 mg piece of the arc-melted and annealed button. The absorber holder comprised two nested white Delrin cups. The powder was placed uniformly on the bottom of the larger cup and was held in place by a smaller cup. The absorber holder was locked in a thermal contact with a copper block with a temperature sensor and a heater, and aligned with the $\gamma$ - source and detector. The absorber was cooled to a desired temperature using Janis model SHI-850-5 closed cycle refrigerator (with vibrations damping) that has long-term temperature stability better than 0.1 K at low temperature. The driver velocity was calibrated by $\alpha$-Fe foil and all isomer shifts (IS) are quoted relative to the $\alpha$-Fe foil at RT. At first, spectra with maximum velocity 6 mm/s and 3 mm/s were both measured at RT to check that no iron-containing impurity can be seen in the M\"ossbauer spectra. Then, three rounds of measurements, progressively focusing in on temperatures near T$_c = 6.1$ K, were carried out: collecting 24 h with maximum velocity 2 mm/s from 4.3 K to 293.8 K (S1); collecting 48 h with maximum velocity 3 mm/s from 4.4 K to 10 K (S2); collecting 48 h with maximum velocity 3 mm/s from 4.7 K to 6.4 K (S3). All the M\"ossbauer spectra were fitted by the commercial software package MossWinn \cite{18MossWinn}, in which the standard error of parameters can be estimated either by calculating and inverting the curvature matrix of the $\chi^2$ with respect to the fit parameters, or by the Monte Carlo method. The standard error of IS, quadroupole splitting (QS) and line width ($\Gamma$) in this work were obtained from the curvature matrix, while the error the area under the spectra here were obtained by Monte Carlo method by iterating 100 times.
\begin{figure}[htp]
\centering
\includegraphics[width=0.5\textwidth]{fig1}
\centering
\caption{\label{fig:epsart} (Color online) Rietveld refinement of the powder XRD spectrum of Lu$_2$Fe$_3$Si$_5$. Measured (black cross), calculated intensities (red line) and background (green line) and difference curve (blue line) are shown. Vertical bars at the bottom indicate the positions of the Bragg reflections. The peaks of unknown phase are marked by the blue arrows.}
\end{figure}
\begin{figure}[htp]
\includegraphics[width=0.5\textwidth]{fig2
\caption{\label{fig:epsart}(Color online) Expanded view of the background part of the RT Lu$_2$Fe$_3$Si$_5$ M\"ossbauer spectrum in large velocity scale. Red arrows show the expected peaks positions for $\alpha$-Fe.The inset is the full view of the spectrum.}
\end{figure}
\section{Results and discussion}
\begin{figure}[htp]
\centering
\includegraphics[width=0.48\textwidth]{fig3}
\centering
\caption{\label{fig:epsart} (Color online) (a) Temperature dependent of zero field cooled magnetic susceptibility of Lu$_2$Fe$_3$Si$_5$ (H = 10 Oe); (b) the temperature dependent low temperature specific heat capacity. Vertical dashed line marks T$_c = 6.1$ K.}
\end{figure}
\begin{figure}[htp]
\centering
\includegraphics[width=0.5\textwidth]{fig4}
\centering
\caption{\label{fig:epsart} (Color online) The specific heat capacity divided by temperature, C$_p$/T, as a function of T$^2$ for Lu$_2$Fe$_3$Si$_5$. The dashed line represents the linear fit to the data in the normal state as described in the text.}
\end{figure}
\subsection{Structure and superconductivity}
The XRD pattern of the Lu$_2$Fe$_3$Si$_5$ polycrystalline sample is presented in Fig. 1. The majority of the peaks match to the tetragonal structure with the $P4/mnc$ space group. The Rietveld refinement results in an estimate of $\sim 1.8(1)$ wt.\% $\alpha$-Fe impurity and a trace amount of unknown phase. A 1.8 wt.\% of $\alpha$-Fe corresponds to 6.7 \% of Fe atoms in the $\alpha$-Fe form in the ground sample, which, for un-enriched Fe, is below the resolution of $^{57}$Fe M\"ossbauer spectroscopy to detect $\alpha$-Fe. As can be seen in Fig. 2, there are no peaks associated with the $\alpha$-Fe or with the other impurity in the M\"ossbauer spectrum, which means the unknown phase is either iron-free or, if contains iron, is below the resolution limit. Consequently, the M\"ossbauer spectra can be analyzed as a single - phase (Lu$_2$Fe$_3$Si$_5$) spectra.
The superconductivity of the sample is confirmed by the dc susceptibility measurement in a magnetic field of 10 Oe. As shown in Fig. 3 (a), the susceptibility data shows diamagnetic signal below $\sim 6.2$ K. The transition is sharp with a width of less than 0.4 K. Fig. 3 (b) shows the temperature dependent specific heat capacity (C$_p$). A sudden jump caused by superconducting transition can be observed below 6.3 K on cooling, the transition temperature is close to that obtained from the susceptibility data. Based on these two thermodynamic measurements we take T$_c$ = 6.1$\pm$0.1 K.
C$_p$/T is also plotted as a function of T$^2$ in Fig. 4. A linear fit above the superconducting transition yields the values of $\gamma_n$ ($\gamma_n$T is the electronic contribution to specific heat capacity) and $\beta_n$ ($\beta_n$$T^3$ represents the phonon contribution) of $\gamma_n$ = 24.6(2) mJ/mol K$^2$, $\beta_n$ = 0.287(2) mJ/mol K$^4$, which are very close to the previously reported values \cite{19PRL2008}. From the $\beta$ value, we can estimate the value of the Debye temperature ($\Theta$$_D$) using the relation: $\Theta$$_D$ = (12$\pi$$^4$$N$$r$k$_B$/5$\beta$)$^{1/3}$, where $N$ is Avogadro's number, $r$ is the number of atoms per formula unit, and k$_B$ is the Boltzmann's constant. We further obtained $\Theta$$_D$ is 408 K. The value of the normalized specific-heat jump at T$_c$, $\Delta$C/$\gamma_n$T$_c$, is $\approx 1.06$, a value that is smaller than the BCS value of 1.43, but consistent with the previously reported value 1.05 \cite{19PRL2008}. From the above characterizations, we can conclude that our sample is a bulk superconductor and can be considered as single phase for M\"ossbauer measurements and analysis.
\subsection{M\"ossbauer results and discussion}
\subsubsection{Symmetry of Fe sites and choice of the model}
In the Lu$_2$Fe$_3$Si$_5$ crystal structure, there are two, nonequivalent, Fe positions, Fe$_{\rm{I}}$ and Fe$_{\rm{II}}$ of 1 : 2 occupation. Fe$_{\rm{I}}$ atoms are located at the 4d sites, which form 1D chains along the $c$ axis. Fe$_{\rm{II}}$ atoms are located at 8h sites, which form squares with planes perpendicular to the $c$ axis. In each Fe position, the Fe atom is located in a polyhedron formed by Si atoms. Fe$_{\rm{I}}$ has four Si atoms at a distance of 2.31 \AA\ which form an irregular tetrahedron and two Si atoms at a distance of 2.54 \AA\ with the nearest Fe-Fe distance is 2.67 \AA. The Fe$_{\rm{II}}$ has four Si atoms at a distance of 2.34 \AA\, which are in the same face and form a quadrangle. On each side of the face, there are two Si atoms with the 2.35 \AA\ Fe-Si distances. The nearest Fe-Fe distance for Fe$_{\rm{II}}$ is 2.71 \AA.
The anisotropic environments of the Fe atoms ensure nonzero electric field gradient (EFG) tensor at both sites. Hence, in analyzing the data, two doublets are expected. All the $^{57}$Fe M\"ossbauer spectra of Lu$_2$Fe$_3$Si$_5$ over the whole observed temperature range share similar spectral shapes with a clear quadrupole splitting as shown in Fig. 5 (a). A very small asymmetry and small shoulders were observed, which suggest that at least two subspectra are needed to resolve the spectra. However, due to the small splitting and consequently poor resolution of the spectra, the data can be analyzed with more than one set of parameters. In the previous M\"ossbauer studies of R$_2$Fe$_3$Si$_5$ (R = rare earth), both (i) one doublet, and (ii) two doublets with fixed area ratio were employed to fit the spectra \cite{16JMMM1981,20JAP1981}. In our approach to the fitting, two doublets without any restriction are used. In model 1, the two subspectra have close values of QS but obviously different values of IS, which is similar to the reported Sc$_2$Fe$_3$Si$_5$ fit result \cite{21PL1980}. In model 2, the two subspectra have similar values of IS, but distinct QS values. Two sets of parameters yielding fits of acceptable quality can be obtained, the corresponding parameters of the RT spectrum fits are listed in the Table 1. The RT spectrum fitted using our two models is shown in Fig. 5 (b) and (c). The relative area of two subspectra in model 2 are closer to the theoretical value 1 : 2 and $\chi^2$ values are closer to 1, so we have chosen the model 2 to fit all the collected spectra. As an example, the fit using model 2 of the spectrum measured at 4.4 K from the S2 set is shown in Fig. 5 (d) and the corresponding parameters are listed in Table 1.
\begin{figure}[htp]
\includegraphics[width=0.48\textwidth]{fig5
\caption{\label{fig:wide}(Color online) $^{57}$Fe M\"ossbauer spectra of Lu$_2$Fe$_3$Si$_5$. (a)the spectrum at RT; the results of the fits using model 1 and model 2 are shown in (b) and (c), respectively; (d) is the spectra collected at 4.4 K and fitted using the model 2. The model 1 and model 2 are described in detail in the text.}
\end{figure}
\begin{table}
\caption{\label{tab:table4}
Hyperfine parameters obtained by fitting using different models as discussed in the text. IS is the isomer shift, QS is the quadrupole splitting and I is the relative area of the two subspectra. }
\begin{ruledtabular}
\begin{tabular}{ccccccc}
T &model &site &IS &QS &I &$\chi^2$ \\
\mbox{K} & & &\mbox{mm/s}&\mbox{mm/s}&\mbox{\%} & \\
\hline
RT &1 &4d & 0.129(3) &0.311(1) &39.4 &1.63 \\
& &8h & 0.278(3) &0.296(1) &60.6 & \\
RT &2 &4d & 0.2157(9) &0.513(6) &36.4 &0.98 \\
& &8h & 0.2206(6) &0.248(8) &63.6 & \\
4.4 &2 &4d & 0.320(2) &0.523(4) &52.8 &1.29 \\
& &8h & 0.330(1) &0.202(3) &47.2 & \\
\end{tabular}
\end{ruledtabular}
\end{table}
\begin{figure*}[htp]
\includegraphics[width=0.7\textwidth]{fig6
\caption{\label{fig:epsart}(Color online) Temperature dependence of parameters derived from fitting the of Lu$_2$Fe$_3$Si$_5$ M\"ossbauer spectra with model 2 as described in the text. (a) Isomer shift, (c) quadrupole splitting, (e) line width and (g) relative area of two subspectra. S1, S2, S3 - mark three measurement sets, the parameters are given for two Fe sites, 4d and 8h (see the text for more details). The lines in panel (a) are fits to the Eq. (1). The figures in the right column are the expansion one of the low temperature region of the corresponding left figure. The temperature stability for each of the measurements in the right column is better than that presented by the size of the symbols.}
\end{figure*}
\subsubsection{Hyperfine parameters}
Fig. 6 summarizes the variation of hyperfine parameters of the spectra with temperature. The plots in the right column present an expanded view of the low temperature range.
The IS of Lu$_2$Fe$_3$Si$_5$ plotted as a function of temperature is shown in Fig. 6 (a) and (b). The IS values of the 4d and 8h sites at 294 K are 0.225(2) mm/s and 0.227(1) mm/s, respectively, which are slightly larger than the typical values for iron-silicon compounds in which Fe carries no moment \cite{22JMMM1983,23JPCSSC1973}, but are about 0.2 mm/s smaller than that in the iron-pnictide compounds \cite{8PC2009,9PRB2011,10EPL2014}. The temperature dependencies of the IS corresponding to the two sites are very similar, and no anomalies can be observed around T$_c$ (Fig. 6(b)). The IS values obtained from the fits includes contributions from both the chemical shift and the second-order Doppler shift, which is known to increase convexly upon decreasing temperature, due to gradual depopulation of the excited phonon states. However, it should be constant at low temperature, because of the quantum mechanical zero-point motion. The chemical shift should not depend on temperature. The main contribution to this variation is from the second-order Doppler shift, which is usually described by Debye model:
\begin{equation}
IS(T)=IS(0)-\frac{9}{2}\frac{k_BT}{Mc}(\frac{T}{\Theta_D})^3\int_0^{\Theta_D/T}\frac{x^3dx}{e^x-1},
\end{equation}
where c is the velocity of light, M is the mass of the $^{57}$Fe nucleus, and $IS(0)$ is is the temperature-independent part, i.e. the chemical shift. A fit with Eq. (1) to the data of S1 shown in Fig. 6 (a) yields $\Theta$$_D$ = 517(18) K and 545(16) K for 4d and 8h sites, respectively.
\begin{figure}[htp]
\includegraphics[width=0.5\textwidth]{fig7
\caption{\label{fig:epsart}(Color online) (a) Temperature dependence of the spectral area for S1 set of measurements on a semi-log scale. The solid line is a fit to Eq.(2), as explained in the text. (b) Normalized to the values at $\approx 6$ K, temperature dependent spectral area data for all three sets of measurements. Inset: enlarged low temperature part.}
\end{figure}
The quadrupole splittings in Lu$_2$Fe$_3$Si$_5$ at the 4d and 8h sites are 0.50(1) mm/s and 0.19(1) mm/s at 294 K, respectively. The magnitude of the QS is proportional to the $z$ component V$_{zz}$ of the EFG tensor, which is composed of two contributions: (V$_{zz}$)$_{lig}$, from the ligand charges around the M\"ossbauer nucleus, and (V$_{zz}$)$_{val}$, from the valence electrons of M\"ossbauer nucleus. Usually, (V$_{zz}$)$_{lig}$ is small and weakly dependent on temperature, whereas (V$_{zz}$)$_{val}$ is strongly temperature dependent. As can be seen from Fig. 5 (c). the QS of two sites are both almost temperature independent, which indicates the QS is mainly determined by the contribution from the ligand charge distribution around the $^{57}$Fe sites.
In order to get a better understanding of the electronic origin of EFG at the Fe sites in Lu$_2$Fe$_3$Si$_5$, a first-principles calculation was performed using the full-potential linearized augmented plane wave method as embodied in the WIEN2K \cite{ADD4_CPC1990,ADD5_WIEN2K}. The generalized gradient approximation (GGA) suggested by Perdew, Burke, and Ernzerhof (PBE GGA) \cite{ADD6_PRL1996} was employed for the exchange-correlation effects. Once the electron densities are calculated self-consistently and with high accuracy, the EFG tensor can be obtained from an integral over the non-spherical charge density. The principal component V$_{zz}$ for the 4d site is 2.25$\times$10$^{21}$V/m$^2$ and asymmetry parameter $\eta = (V_{xx} - V_{yy} )/V_{zz} $ = 0.457; the V$_{zz}$ for the 8h site is -1.16$\times$10$^{21}$V/m$^2$ and $\eta$ = 0.307, which qualitatively agree with experimental results. It is found that the $p$-$p$ and $d$-$d$ interactions mainly contribute to the EFG of Lu$_2$Fe$_3$Si$_5$ and $p$- electrons is play a significant role for states far from the Fermi energy whereas the $d$-$d$ interaction dominates around the Fermi energy, which is similar to what is found for the iron-pnictides superconductors \cite{ADD7_PRB2011,ADD8_HI2011}.
The relative areas of the subspectra are determined by the proportion of the Fe atoms on different lattice sites. As mentioned above, for Lu$_2$Fe$_3$Si$_5$, the theoretical relative area of the two subspectra representing iron atoms at the 4d and 8h sites should be 33.3\% and 66.7\%. At RT the relative areas are close to expected value. However, as shown in Fig. 6 (e) as temperature decreases, the relative area values deviate from theoretical values gradually, with the relative area for the 4d site becoming even larger than that for 8h site at low temperature. This phenomenon may be related to the variation of the linewidth, $\Gamma$, for the doublets corresponding to the two sites with reducing temperature. As shown in Fig. 6 (g), the $\Gamma$ of 4d site increases slightly at lower temperature, but the $\Gamma$ of 8h site is almost constant during the whole temperature range. The slightly increase of $\Gamma$ of 4d site suggests the existence of instability at this lattice site at low temperature.
\subsubsection{The spectral area}
In M\"ossbauer studies of superconductivity, the variation of the total spectral area has also been the focus of discussion. There are number of reports showing a decrease of the spectral area near T$_c$ due to the softening of lattice with the opening of superconducting gap in cuprates \cite{1PRL1991,2PRB1992,3PRB1988} but to the best of our knowledge only one report on iron-pnictide superconductors \cite{24SSC2011}. There are two kinds of behavior reported: a rapid decrease near T$_c$ \cite{1PRL1991,4JPCM1993,5JPCM1994}; and a pit-like decrease either around T$_c$ or at higher temperature, serving as a precursor to T$_c$ \cite{2PRB1992}. At the same time, some publications noted a poor reproducibility of those observations in the cuprates \cite{6SSC1987,25PRB1992}. In our measurements, as can be seen in the Fig. 7, for the S1, we did not observe any abnormal variation around 6.1 K. To make sure we didn't miss any minor variation around T$_c$, we remeasured the M\"ossbauer spectra of the same sample between 4.4 K to 10 K with higher density of the data points around 6.1 K for 48 h. Surprisingly, for data set S2, there is a sharp, 6\%, decrease around 5.5 K. Nevertheless, when we repeated the same measurement again, data set S3, this phenomenon disappeared. The sharp spectral area change seen in data set S2 is, in our opinion, most likely an artifact. The feature in data set S2 occurs resolvably below T$_c$ = 6.1 K with the very sharp and rather large jump occuring between spectra taken at 5.3 K and 5.7 K (with spectra at 5.7 K, 6.0 K, and 6.3 K all having normalized areas near 1.0). If this feature were associated with T$_c$, it should have occured either between 5.7 K and 6.0 K or between 6.0 K and 6.3 K. In addition, there are no corresponding anomalies in data set S2's isomer shift, quadrupole splitting or relative areas (as shown in Fig. 6). A simple explanation for such behavior could be, among others, some mechanical shift/rearrangement of the powder composing the absorber. This observation gives a warning that even in the measurement of iron-containing stoichiometric material, irreproducibilities/artifacts might exist and should be addressed appropriately.
Finally, we also fitted the temperature dependence of area under the two doublets line of S1 measurements with Debye model:
\begin{equation}
f=exp[\frac{-3E_{\gamma}^2}{k_B\Theta_DMc^2}\{\frac{1}{4}+(\frac{T}{\Theta_D})^2\int_0^{\Theta_D/T}\frac{xdx}{e^x-1}\}],
\end{equation}
where $f$ is the recoilless fraction, which is proportional to the area for thin sample and E$_{\gamma}$ is the $\gamma$-ray energy. This expression also allows to estimate the value of $\Theta_D$. We obtained the $\Theta_D$ = 385(8) K which is very close to the 408 K obtained from the analysis of the low temperature specific heat capacity data and about 140 K less than the value estimated by temperature dependence of IS. A similar difference was found earlier in studies of e.g. FeSe$_{0.5}$Te$_{0.5}$ and $^{57}$Fe- doped YBa$_2$Cu$_3$O$_{6.8}$ compounds \cite{24SSC2011,26SSC1995}. This discrepancy may be explained by the fact the area reflects the average mean-square displacements, while IS related to the mean-square velocity of the M\"ossbauer atom. Both quantities may respond in different ways to the lattice anharmonicities.
\section{Conclusions}
In summary, we performed detailed $^{57}$Fe M\"ossbauer measurements on Lu$_2$Fe$_3$Si$_5$ in the temperature range of 4.4 K to RT. The contributions from two Fe crystallographic sites can be well distinguished by M\"ossbauer spectra. The main contribution of EFG in this compound comes from the lattice anisotropy and the first principles calculations yield the values of EFG that qualitatively agree with the experiment. The Debye temperature was estimated by the temperature dependence of specific heat capacity, spectral area and IS. The $\Theta_D$ obtained from temperature dependence of spectral area and heat capacity are very similar, but about 140 K smaller than the value estimated by the IS variation with temperature. Additionally, we didn't observe any obvious, abnormal variation of hyperfine parameters around T$_c$. Two possibilities could lead to this result: the opening of the superconducting gap doesn't bring variation of the environment at Fe site at all; the T$_c$ of this system is too low, and M\"ossbauer spectroscopy is not sensitive enough to detect the minute change.
\begin{acknowledgments}
X. M. was supported in part by the China Scholarship Council. The authors (H. P. and F. L.) gratefully acknowledge the financial support from the National Natural Science Foundation of China under grant No. 11275086. Work at the Ames Laboratory (X. M., S. R., P. C. C. and S. L. B. ) was supported by the US Department of Energy, Basic Energy Sciences, Division of Materials Sciences and Engineering under Contract No. DE-AC02-07CH11358.
\end{acknowledgments}
\nocite{*}
|
\section*{Introduction}
Modeling biological or physical systems often requires handling one-dimensional diffusion processes. The marginal probability distribution of such processes, at a fixed time, permits
a quite precise description of the model. Nevertheless, in many applications, this information is insufficient and the description of the whole path becomes crucial. This is namely the case for a variety of problems related to neuronal sciences, financial derivatives with barriers, ruin probability of an insurance fund, optimal stopping problems,... In these frameworks, the main task is the description of the first passage time densities for time-dependent boundaries. Let us just mention some references in engineering reliability \cite{Ebrahimi-2005}, epidemiology \cite{Tuckwell-Wan-2000}, biology \cite{Ricciardi-al-1999}, mathematical finance \cite{Garrido-1989, Roberts-al-1997, Novikov-al-2003} and references concerning the framework of level-crossing problems \cite{Abrahams, Blake}.
For instance, let us focus our attention on a simple interpretation of neural transmission. When a neuron is stimulated by pressure, heat, light, or chemical information, its membrane voltage changes as time elapses and, as soon as it reaches a constant threshold, the depolarization phenomenon occurs and the voltage is reset to a resting potential. The family of integrate-and-fire spiking neuron models is based on this simple interpretation. The firing time therefore corresponds to the first-passage time of the membrane potential, represented by a stochastic mean-reverting process (usually the Ornstein-Uhlenbeck process) to the neural threshold (Giorno et al. \cite{Giorno-al-1988}, Lansky et al. \cite{Lansky-1995}, Wan and Tuckwell \cite{Wan-Tuckwell-1982}, for an introduction to noise in the nervous system see Part I Chapter 5 in \cite{Gerstner}, for the integrate-and-fire model see Chapter 10 in \cite{Ermentrout}).
Our main motivation is to emphasize an algorithmic approach in order to approximate the first-passage time of the Brownian motion to curved boundaries. The field of application of such an algorithm at a first glance may appear as quite restrictive since it concerns the Brownian motion but in fact a lot of families of diffusion processes are concerned. Indeed it is possible to express various stochastic paths as functions of Brownian paths in the spirit of Wang and P\"{o}tzelberger \cite{Wang-Potzelberger-2007}.
Hence using simple time transformations, we are going to present an application of the results to the Ornstein-Uhlenbeck process (see Section~\ref{sec:examples}).
In order to describe approximations of the first-passage time of the Brownian motion,
we assume that this stopping time is almost surely finite.
In this way, we introduce particular conditions for this property to be satisfied.
Let us consider a continuous function $\varphi:\mathbb{R}_+\to\mathbb{R}$
satisfying the following hypothesis:
\begin{equation}
\label{hypo}
\varphi(0)>0\quad\mbox{and}\quad \limsup_{t\to\infty}\frac{\varphi(t)}{\sqrt{2t\log\log t}} <1.\tag{H1}
\end{equation}
We then define the hitting time
\begin{equation}
\label{def:tau}
\tau_\varphi=\inf\{t>0:B_t=\varphi(t)\}
\end{equation}
where $(B_t,\,t\ge 0)$ stands for a standard one-dimensional Brownian motion. Under \eqref{hypo}, the a.s. finiteness of $\tau_\varphi$ is an obvious consequence of the law of the iterated logarithm (see e.g.\cite[Th.9.23 p.112]{karatzas}). It is quite difficult to obtain precise information about this stopping time in general situations.
The study of the approximation of the hitting times for Brownian motion and general Gaussian Markov processes is an active area of research. Several alternatives for dealing with the characterization of hitting times exist.
\subsubsection*{Approximation of the probability density function}
For particular cases, the probability density function of the Brownian passage time can be computed explicitly. Lerche \cite{Lerche} used the method of images in order to obtain explicit expressions of the p.d.f. $p$ defined by $p(t)\,dt=\mathbb{P}(\tau_\varphi\in dt)$. However only few cases are concerned by such a study.
Durbin \cite{Durbin85, Durbin92} proposed to approximate the first-passage distribution $p(t)$ of the Brownian motion as follows: $p$ can be represented by an expansion
\[
p(t)=\sum_{j=1}^k (-1)^{j-1}q_j(t)+(-1)^k r_k(t),\quad k\ge 1,
\]
where $q_j$ for $1\le j\le k$ and $r_k$ are defined by multiple integrals depending on the boundary $\varphi$. The approximation simply consists in truncating the expansion.
Let us note that the first term corresponds in fact to the tangent approximation of Strassen \cite{Strassen} and Daniels \cite{Daniels}, see also \cite{Ferebee-1982}. The convergence of the series and the error bounds
can be made precise if the curved boundary is wholly concave or wholly convex.
Many studies concern such a series expansion, let us mention a few: Ferebee \cite{Ferebee-1983}, Ricciardi et al. \cite{Ricciardi-al-1984}; Giorno et al. \cite{Giorno-al-1989},
Sacerdote and Tomassetti \cite{Sacerdote-Tomassetti-1996} to deal with more general diffusion processes. The numerical approach proposed in \cite{Buonocore-1987} seems to be particularly efficient. In \cite{Nardo-al-2001}, the authors proposed a comparison between the approximation developed by Durbin and an other numerical resolution of the Volterra equation for Gaussian processes.
One method in approximating the passage time of a Brownian motion, or even of a quite general diffusion, through a curved boundary is to replace the initial boundary by an other one which is close and which leads to an explicit expression of the hitting time probability. Such method permits to obtain some bounds. It was first introduced for the Brownian motion in \cite{Borovkov-Novikov-2005},
and applied for instance to piecewise continuous boundaries \cite{Wang-Potzelberger-2007}.
Finally an other method consists in writing the p.d.f of the hitting time as the expectation of a particular functional of a three-dimensional Brownian bridge. It suffices then to approximate this expectation through a Monte Carlo method \cite{Ichiba-2011}.
\subsubsection*{Approximation of the first passage time.}
All methods described so far concern the approximation of the pdf. It can be of particular interest to simulate directly the first passage time $\tau_\varphi$ or to compute the probability for the hitting time to be smaller than some given $T>0$, without computing the pdf. The solution consists in using a time discretization of the Brownian motion on $[0,T]$. The time interval is then split into $n$ small intervals of the kind $[(k-1)T/n,kT/n]$, with $1\le k\le n$.
It is therefore possible just to simulate the hitting time of the corresponding Euler scheme. Of course this should upper-bound the stopping time. One solution to overcome this problem is to improve the algorithm by shifting the boundary to reach: we stop the Euler scheme as soon as it exits from a suitable smaller domain. Let us note that this general procedure can also be applied to diffusion processes. It has been first introduced for geometrical Brownian motion (finance) in \cite{Broadie-Glasserman-Kou-1997} and then extended to general diffusions with \emph{nice} coefficients in \cite{Gobet-Menozzi-10}.
An other method in order to improve the approximation of the hitting time consists in testing, at each endpoint $kT/n$, if the event $B_{kT/n}<\varphi(kT/n)$ is satisfied and if the Brownian path on the small intervals,
conditionally on its value at the end point, hits the curved boundary. This method can also be applied to diffusions and needs therefore precise asymptotics of hitting probabilities for pinned diffusions. A first important study in that direction is \cite{Gobet-2000} where the coefficients of the diffusion are frozen at the starting point on each small interval leading to asymptotics of the probabilities. Nevertheless the method can become onerous if the observed time interval $[0,T]$ is large and sometimes gives incorrect asymptotics: it has been pointed out, by the numerical treatment of some precise examples, that the approximations produced by this method can be far from the true ones. See for this point Giraudo and Sacerdote \cite{Giraudo-Sacerdote-1999} (O.U. process and Feller model),
who also suggest some formulas for the computation of the crossing probability, see also \cite{Giraudo-al-2001}. Baldi and Caramellino \cite{Baldi-Caramellino-2002} presents precise asymptotics for general pinned diffusions which permits to improve the approximation of hitting times. Such results can be developed further in the particular gaussian framework for any dimension \cite{Caramellino-Pacchiarotti-Salvadei-2015}.
\subsubsection*{A new algorithm.}
The aim of this study is to present a new method of approximation of $\tau_\varphi$. Let us explain intuitively the simulation procedure. If $\varphi$ is an increasing curve with $\varphi(0)>0$, then the Brownian motion needs to successively cross a sequence of imaginary horizontal lines before hitting the boundary. The first line to cross corresponds to the value $\varphi(0)$ and needs a random time denoted by $\mathcal{T}_1$. At that time, the value of the curved boundary is $\varphi(\mathcal{T}_1)$. The Brownian motion therefore needs to cross this second horizontal line, it shall happen at time $\mathcal{T}_2$ and the new horizontal line to cross becomes $\varphi(\mathcal{T}_2)$ and so on... Figure \ref{fig:description} (left) illustrates this procedure. The sequence of stopping times $(\mathcal{T}_n)$ converges towards $\tau_\varphi$ and will be used in order to obtain an approximation. We shall introduce a stopping procedure in this sequence of random times which depends on a small parameter $\epsilon$ associated to the error size of the approximation: the sequence is stopped as soon as the distance between two successive horizontal lines is smaller than $\epsilon$. The outcome of the algorithm corresponds therefore to a random variable $\tau_\varphi^\epsilon$ which can be exactly simulated
and such that $\tau_\varphi^{\epsilon}$ converges toward $\tau_\varphi$ in distribution as $\epsilon$ tends to $0$.
The algorithm can be modified when the curved boundary $\varphi$ does not satisfy the monotonic property anymore. In such a slightly different context, it suffices to tilt the successive imaginary horizontal lines in such a way that the common slope corresponds to $\inf_{t\ge 0}\varphi'(t) $, see Figure \ref{fig:description} (right).
\begin{figure}[h]
\centerline{\includegraphics[width=0.48\textwidth]{algo-descript-2.pdf}
\includegraphics[width=0.5\textwidth]{algo-descript-3.pdf}}
\caption{Illustration of the algorithm for an increasing boundary $\varphi$ with its associated successive horizontal lines (left) and for a general boundary (right).}
\label{fig:description}
\end{figure}
To sum up, two different families of sequences will be developed and the associated
convergence rates are estimated. The first algorithm developed in Section \ref{sec:incre} concerns increasing curved boundaries and the second one, Section \ref{sec:gen}, permits
us to deal with quite general boundaries provided that its derivative is bounded. In the last section, we present different examples in order to illustrate the algorithm efficiency.
\section{First-passage time to non-decreasing boundaries}\label{sec:incre}
Let us assume that the boundary $\varphi$ satisfies \eqref{hypo} and that the following additional conditions hold
\begin{equation}
\label{hypo+}
\varphi:\mathbb{R}_+\to\mathbb{R}\quad\mbox{is a non-decreasing }\ \mathcal{C}^1\mbox{-continuous function,}\tag{H2}
\end{equation}
\begin{equation}
\label{hypo++}
2\varphi'(t)\sqrt{1+t}\le 1,\quad \forall t\ge 0.\tag{H3}
\end{equation}
We introduce the algorithm associated to the hitting time $\tau_\varphi$ defined by \eqref{def:tau}.
\begin{Algo}
\label{algo1}
Let $\epsilon>0$ be a small parameter and $(G_n)_{n\ge 0}$ a sequence of independent standard Gaussian distributed random variables.\\ Initialization: $T_0=0$, $T_1=(\varphi(0)/G_0)^2$ and $\mathcal{N}_\epsilon=1$.\\
While $\varphi(T_1)-\varphi(T_0)>\epsilon$ do:
\begin{eqnarray}\label{eq:algo1}
\left\{\begin{array}{l}
(T_0,T_1)\leftarrow \Big(T_1,T_1+(\varphi(T_1)-\varphi(T_0))^2/G_{\mathcal{N}_\epsilon}^2\Big)\\
\mathcal{N}_\epsilon\leftarrow \mathcal{N}_\epsilon+1.
\end{array}\right.
\end{eqnarray}
Outcome: $\tau_{\varphi}^\epsilon\leftarrow T_1$ and $\mathcal{N}_\epsilon$.
\end{Algo}
Let us just note that Algorithm \ref{algo1} is very simple to use since each step only requires one Gaussian distributed random variable. Moreover it is a approximation of the first-passage time:
\begin{thm}\label{thm1}\begin{enumerate}
\item Let us assume that the boundary function $\varphi$ satisfies \eqref{hypo}, \eqref{hypo+} and \eqref{hypo++} then the random variable $\tau_\varphi^\epsilon$ defined in Algorithm \ref{algo1} converges in distribution towards $\tau_\varphi$ defined by \eqref{def:tau} as $\epsilon$ tends to zero. More precisely
\begin{equation}\label{eq:thm:1}
F_\epsilon(t-\epsilon)-\frac{3\sqrt{\epsilon}}{\sqrt{2\pi}}\le F(t)\le F_\epsilon(t),\quad\mbox{for any}\quad t\ge \epsilon,
\end{equation}
where $F$ (resp. $F_\epsilon$) is the cumulative distribution function of $\tau_\varphi$ (resp. $\tau_\varphi^\epsilon$).
\item There exists a constant $C>0$ such that the random number of iterations $\mathcal{N}_\epsilon$ defined in Algorithm \ref{algo1} satisfies:
\begin{equation}
\label{eq:thm:conv1}
\mathbb{E}[\mathcal{N}_\varepsilon]\le C\sqrt{|\log \epsilon|}.
\end{equation}
\end{enumerate}
\end{thm}
The parameter $\epsilon$ describes the precision of the approximation. The number of steps in the Algorithm \ref{algo1} is very small (even smaller than usual results obtained for algorithms based on random walks on spheres, which are close to Algorithm \ref{algo1}, see \cite{muller}) : in fact the constant appearing in \eqref{eq:thm:conv1} can be explicitly computed: for any constant $0<\kappa<1/2$, there exists $\epsilon_0(\kappa)>0$ such that \eqref{eq:thm:conv1} is satisfied as soon as
$\epsilon<\epsilon_0$, with the particular constant
\[
C=\frac{1}{m\kappa},\quad m=\log(4)+\frac{2\sqrt{2}}{\sqrt{\pi}}\, \mu
\]
and
\begin{equation}
\label{eq:def:mu}
\mu=\int_0^\infty(\log|x|)\,e^{-x^2}\,dx.
\end{equation}
The proof of Theorem \ref{thm1} is based on a main argument developed in the following proposition: each step of Algorithm \ref{algo1} has to be related to a particular part of the Brownian paths before hitting the boundary.
\begin{prop}
\label{prop:temps} Let $(B_t,\, t\ge 0)$ be a standard one-dimensional Brownian motion. We define the following sequence of stopping times: $s_0=\mathcal{T}_0=0$ and for any $n\ge 1$:
\begin{equation}\label{eq:prop:temps}
s_n:=\inf\Big\{t\ge 0:\ B_{t+\mathcal{T}_{n-1}}=\varphi(\mathcal{T}_{n-1})\Big\}\quad \mbox{and} \quad \mathcal{T}_n:=s_1+\ldots+s_n,
\end{equation}
where the function $\varphi$ satisfies \eqref{hypo}, \eqref{hypo+} and \eqref{hypo++}.
Then the following properties hold:
\begin{enumerate}
\item $(\mathcal{T}_n)_{n\ge 0}$ is a non-decreasing sequence which almost surely converges towards $\tau_\varphi$.
\item Let $n\ge 1$, then the probability distribution of $s_{n+1}$ given the $\sigma$-algebra $\mathcal{F}_n:=\sigma\{ s_1\ldots,s_n\}$ is identical as $( \varphi(\mathcal{T}_{n})-\varphi(\mathcal{T}_{n-1}) )^2/G_n^2$ where $(G_n)_{n\ge 0}$ is a sequence of independent standard Gaussian random variables. Moreover $s_1\stackrel{(d)}{=}(\varphi(0)/G_0)^2$.
\item Let $\mathcal{M}_\epsilon:=\inf\{n\ge 1:\ \varphi(\mathcal{T}_n)-B_{\mathcal{T}_n}\le \epsilon\}$, then $\mathcal{T}_{\mathcal{M}_\epsilon}$ and $\tau_\varphi^\epsilon$, defined in Algorithm \ref{algo1}, are identically distributed, so are $\mathcal{M}_\epsilon$ and $\mathcal{N}_\epsilon$.
\end{enumerate}
\end{prop}
Let us note that the mean of each random variable $s_n$ defined by \eqref{eq:prop:temps} is infinite since $\mathbb{E}[G^{-2}]=+\infty$ where $G$ is a standard Gaussian variable. Proposition \ref{prop:temps} suggests that the first-passage time can be obtained as a sum of positive random variables of infinite average, we easily deduce $\mathbb{E}[\tau_\varphi]=+\infty$. In the particular case of increasing boundaries $\varphi$, the sum has infinitely many terms.
\begin{proof}[Proof of Proposition \ref{prop:temps}] ~\\\textbf{Step 1.}
By construction, the sequence $(\mathcal{T}_n)_{n\ge 0}$ is non-decreasing and non-negative: it converges almost surely to $\mathcal{T}_\infty$. Since $\varphi$ is a non-decreasing boundary, $\mathcal{T}_n\le \tau_\varphi$ for any $n\ge 0$. In particular $\mathcal{T}_\infty$ is less than $\tau_\varphi$ which is a finite stopping time due to the law of the iterated logarithm, see \eqref{hypo} followed by discussion. Consequently, the random variable $B_{\mathcal{T}_\infty}$
is well defined.
Since $\varphi$
is non-decreasing, we get $B_{\mathcal{T}_n}=\varphi(\mathcal{T}_{n-1})
$ for any $n\ge 1$.
Taking the large $n$ limit leads to $B_{\mathcal{T}_\infty} = \varphi(\mathcal{T}_\infty)$, the Brownian paths and the function \(\varphi\) being continuous. We deduce that $\mathcal{T}_\infty=\tau_\varphi$.
\\
\textbf{Step 2.} Let us first consider the stopping time $s_1$. Using the reflection principle of the Brownian paths and a scaling property, we obtain:
\begin{align*}
\mathbb{P}(s_1> t)&=\mathbb{P}\Big( \sup_{0\le u\le t}B_u < \varphi(0) \Big)=\mathbb{P}(|B_t|<\varphi(0))\\
&=\mathbb{P}(B_1^2< \varphi(0)^2/t)=\mathbb{P}(\varphi(0)^2/G_0^2> t),\quad t\ge 0.
\end{align*}
The general $n$-th case can be proven using similar arguments combined with the Markov property of the Brownian motion:
\begin{align*}
\mathbb{P}(s_{n+1}> t\vert \mathcal{F}_n)&=\mathbb{P}\Big( \sup_{\mathcal{T}_n\le u\le \mathcal{T}_n+ t}B_u < \varphi(\mathcal{T}_n) \Big| \mathcal{F}_n\Big)\\
&=\mathbb{P}\Big( \sup_{0\le u\le t}B_{u+\mathcal{T}_n}-B_{\mathcal{T}_n} < \varphi(\mathcal{T}_n)-\varphi(\mathcal{T}_{n-1}) \Big| \mathcal{F}_n\Big)\\
&=\mathbb{P}\Big( \sup_{0\le u\le t}\tilde{B}_u < \varphi(\mathcal{T}_n)-\varphi(\mathcal{T}_{n-1}) \Big| \mathcal{F}_n\Big),
\end{align*}
where $\tilde{B}$ is a Brownian motion independent of $\mathcal{F}_n$.\\
\textbf{Step 3.} Using the results developed in Step 2, we observe that $(s_n)_{n\wedge \mathcal{M}_\epsilon}$ and the sequence of values $T_1$, defined in Algorithm \ref{algo1}, have the same distribution. It is therefore obvious that $\mathcal{T}_{\mathcal{M}_\epsilon}$ and $\tau_\varphi^\epsilon$ are identically distributed. Indeed the stopping time can be rewritten as follows:
\begin{equation}\label{eq:reecr}
\mathcal{M}_\epsilon=\inf\{ n\ge 1: \varphi(\mathcal{T}_n)-\varphi(\mathcal{T}_{n-1})\le \epsilon \}.
\end{equation}
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm1}] ~\\
\textbf{Step 1.} Let us recall that $\mathcal{T}_n$ is defined by \eqref{eq:prop:temps}. By Proposition \ref{prop:temps}, $\mathcal{T}_n\le \tau_\varphi$ for any $n\ge 0$ and in particular $\mathcal{T}_{\mathcal{M}_\epsilon}\le \tau_\varphi$. Hence
\[
\mathbb{P}(\mathcal{T}_{\mathcal{M}_\epsilon}\le t)\ge \mathbb{P}(\tau_\varphi\le t),\quad \forall t\ge 0.
\]
Since $\tau_\varphi^\epsilon$ has the same distribution as $\mathcal{T}_{\mathcal{M}_\epsilon}$, we obtain
\begin{equation}
\label{eq:repar-facil}
F_\epsilon(t)\ge F(t),\quad\forall t\ge 0,
\end{equation}
where $F_\epsilon$ and $F$ are the associated cumulative distribution functions. Let us now prove the second bound in \eqref{eq:thm:1}. For $t\ge \epsilon$,
\begin{align}\label{eq:etap1}
F_\epsilon(t-\epsilon)&=\mathbb{P}(\tau_\varphi^\epsilon\le t-\epsilon)=\mathbb{P}(\mathcal{T}_{\mathcal{M}_\epsilon}\le t-\epsilon)\nonumber\\
&\le\mathbb{P}(\mathcal{T}_{\mathcal{M}_\epsilon}\le t-\epsilon,\tau_\varphi>t)+\mathbb{P}(\tau_\varphi\le t)\nonumber\\
&\le \mathbb{P}(|\mathcal{T}_{\mathcal{M}_\epsilon}-\tau_\varphi|>\epsilon)+F(t).
\end{align}
Combining the Markov property of the Brownian motion and the reflection principle leads to
\begin{align*}
P_\epsilon &:=\mathbb{P}(|\mathcal{T}_{\mathcal{M}_\epsilon}-\tau_\varphi|>\epsilon)\le 1-\mathbb{P}\Big(\sup_{0\le u \le \epsilon} B_{\mathcal{T}_{\mathcal{M}_\epsilon}+u}\ge
\sup_{0\le u \le \epsilon} \varphi(\mathcal{T}_{\mathcal{M}_\epsilon}+u)\Big)\\
&\le 1-\mathbb{P}\Big( \sup_{0\le u \le \epsilon} B_{\mathcal{T}_{\mathcal{M}_\epsilon}+u}
- B_{\mathcal{T}_{\mathcal{M}_\epsilon}}\ge \sup_{0\le u \le \epsilon} \varphi(\mathcal{T}_{\mathcal{M}_\epsilon}+u)- \varphi(\mathcal{T}_{\mathcal{M}_\epsilon}) + \epsilon\Big) \\
&\le 1-\mathbb{P}\Big( \sup_{0\le u \le \epsilon} B_{\mathcal{T}_{\mathcal{M}_\epsilon}+u}
- B_{\mathcal{T}_{\mathcal{M}_\epsilon}}\ge \varphi(\mathcal{T}_{\mathcal{M}_\epsilon}+\epsilon)- \varphi(\mathcal{T}_{\mathcal{M}_\epsilon}) + \epsilon\Big) \\
&\le 1-\mathbb{P}\Big( \sup_{0\le u \le \epsilon}\tilde{B}_u\ge \sup_{\mathcal{T}_{\mathcal{M}_\epsilon}\le \theta \le \mathcal{T}_{\mathcal{M}_\epsilon}+\epsilon} \varphi'(\theta)\epsilon+\epsilon\Big)\\
&\le 1-\mathbb{P}\Big( |\tilde{B}_\epsilon|\ge \sup_{\mathcal{T}_{\mathcal{M}_\epsilon}\le \theta \le \mathcal{T}_{\mathcal{M}_\epsilon}+\epsilon} \varphi'(\theta)\epsilon+\epsilon\Big).
\end{align*}
Using Hypothesis \eqref{hypo++} and straightforward computations permits us to obtain
\begin{equation}\label{eq:etap2}
\mathbb{P}(|\mathcal{T}_{\mathcal{M}_\epsilon}-\tau_\varphi|>\epsilon)\le 1-\mathbb{P}(|\tilde{B}_\epsilon|\ge 3\epsilon/2)\le 3\sqrt{\frac{\epsilon}{2\pi}}.
\end{equation}
The lower bound in \eqref{eq:thm:1} holds due to both \eqref{eq:etap1} and \eqref{eq:etap2}. \\
\textbf{Step 2.} Let us now focus our attention to the efficiency of this algorithm.
We need to estimate the number of steps which depends on the small parameter $\epsilon$. Using the third result presented in Proposition \ref{prop:temps} on one hand and \eqref{eq:reecr} on the other hand, we obtain
\[
\mathbb{P}(\mathcal{N}_\epsilon> n)=\mathbb{P}(\mathcal{M}_\epsilon> n)=\mathbb{P}(\varphi(\mathcal{T}_1)-\varphi(\mathcal{T}_0)>\epsilon,\ldots,\varphi(\mathcal{T}_n)-\varphi(\mathcal{T}_{n-1})>\epsilon ).
\]
Hypothesis \eqref{hypo++} implies
\begin{equation}\label{eq:lien}
\mathbb{P}(\mathcal{N}_\epsilon> n)\le \mathbb{P}(s_1>2\epsilon,\ldots, s_n>2\epsilon).
\end{equation}
\textbf{Step 2.1.} Let us first estimate the previous upper-bound. We introduce a sequence of independent standard Gaussian random variables $(G_n)_{n\ge 0}$ and define
\begin{equation}
\label{eq:def:Z}
X_n=\log(4G_n^2),\quad \Xi_n=\sum_{k=0}^n X_k\quad\mbox{and}\quad Z_n=\sum_{k=0}^n \Xi_k.
\end{equation}
Let us define $\Pi(n,\epsilon):=\mathbb{P}(s_n>2\epsilon)$.
By Proposition \ref{prop:temps}, we know that the random variables $s_{n+1}$
are related to $G_n$ and therefore
\begin{align*}
\Pi(1,\epsilon)&=\mathbb{P}(2\epsilon G_0^2<\varphi(0)^2)=\mathbb{P}\Big(\log(4G_0^2)<-\log(\epsilon)+\log(2)+2\log\varphi(0)\Big)\\
&=\mathbb{P}\Big( Z_0< -\log(\epsilon)+\log(2)+2\log\varphi(0)\Big).
\end{align*}
Let us prove that, for $n\ge 1$, we have the general formula:
\begin{equation}
\label{eq:genform}
\Pi(n,\epsilon)\le \mathbb{P}\Big(Z_{n-1}<-\log(\epsilon)+(2n-1)\log(2)+(2n)\log \varphi(0)\Big).
\end{equation}
By Proposition \ref{prop:temps}, we have for $n\ge 2$,
\begin{align}\label{eq:1}
\Pi(n,\epsilon)=\mathbb{P}\Big( ( \varphi(\mathcal{T}_{n-1})-\varphi(\mathcal{T}_{n-2}) )^2>2\epsilon G_{n-1}^2 \Big).
\end{align}
Since $\varphi$ is a non decreasing function satisfying Hypothesis \eqref{hypo++}, the following upper-bound holds for $n\ge 2$:
\begin{equation}\label{eq:2}
\varphi(\mathcal{T}_{n-1})-\varphi(\mathcal{T}_{n-2}) \le \frac{\mathcal{T}_{n-1}-\mathcal{T}_{n-2} }{2\sqrt{1+\mathcal{T}_{n-2}}}\le\frac{s_{n-1}}{2\sqrt{1+s_{n-2}}}.
\end{equation}
Hence for $n=2$, \eqref{eq:1} and \eqref{eq:2} imply
\begin{align*}
\Pi(2,\epsilon)&\le \mathbb{P}\Big(\frac{s_{1}^2}{2^2}>2\epsilon G_1^2\Big)=\mathbb{P}\Big(\epsilon (2 G_1^2)(2G_0^2)^2<\varphi(0)^4\Big)\\
&=\mathbb{P}(2X_0+X_1<-\log(\epsilon)+3\log(2)+4\log\varphi(0))\\
&=\mathbb{P}(Z_1<-\log(\epsilon)+3\log(2)+4\log\varphi(0)).
\end{align*}
Using the lower-bound $1+s_{n-1}\ge s_{n-1}$ and similar arguments as those developed previously, the general case is expressed as follows:
\begin{align*}
\Pi(n,\epsilon)&\le \mathbb{P}\Big( \frac{s_{n-1}^2}{2^2(1+s_{n-2})}>2\epsilon G_{n-1}^2 \Big)\le \mathbb{P}\Big( \frac{s_{n-2}^3}{2^22^4s_{n-3}^2}>2\epsilon G_{n-1}^2G_{n-2}^4 \Big)\\
&\le \mathbb{P}\Big(\frac{s_{2}^{n-1}}{s_1^{n-2}}>\epsilon 2 2^2 2^4\ldots 2^{2(n-2)}G_{n-1}^2G_{n-2}^4\ldots G_{2}^{2(n-2)}\Big)\\
&\le\mathbb{P}\Big(\Big(\frac{s_{1}^2}{2^2G_1^2}\Big)^{n-1}\frac{1}{s_1^{n-2}}>\epsilon 2 2^2 2^4\ldots 2^{2(n-2)}G_{n-1}^2G_{n-2}^4\ldots G_{2}^{2(n-2)}\Big)\\
&\le\mathbb{P}\Big(\varphi(0)^{2n}>\epsilon 2 2^2 2^4\ldots 2^{2(n-1)}G_{n-1}^2G_{n-2}^4\ldots G_{0}^{2n}\Big)\\
&\le \mathbb{P}\Big(Z_{n-1}<-\log(\epsilon)+(2n-1)\log(2)+(2n)\log\varphi(0)\Big).
\end{align*}
\textbf{Step 2.2.} By \eqref{eq:lien} and the arguments developed in Step 2.1, we obtain
\[
\mathbb{P}(\mathcal{N}_\epsilon>n)\le \mathbb{P}(s_n>2\epsilon)\le \mathbb{P}(Z_{n-1} - \mathbb{E}Z_{n-1}<\eta(\epsilon,n)- \mathbb{E}Z_{n-1}),
\]
where \[
\eta(\epsilon,n):=-\log(\epsilon)+(2n-1)\log(2)+(2n)\log\varphi(0).
\]
Let us observe that, for any $n\ge 0$,
$m:=\mathbb{E}[X_n]=\log(4)+\frac{2\sqrt{2}}{\sqrt{\pi}}\mu>0$ where $\mu$ is defined by \eqref{eq:def:mu}. Hence
\[
\mathbb{E}[Z_n]=\sum_{k=0}^n\mathbb{E}[\Xi_n]=\sum_{k=0}^n\sum_{j=0}^k\mathbb{E}[X_j]=m \sum_{k=0}^n(k+1)=\frac{m(n+1)(n+2)}{2}.
\]
Thus, for \(n\) large enough, \(\eta(\epsilon,n)- \mathbb{E}Z_{n-1} < 0\).
Introducing $d_n:=|mn(n+1)/2-\eta(\epsilon,n)|$, we observe that, for any $0<\kappa<1/2$ there exists $\aleph(\kappa,\epsilon)\in\mathbb{N}$ such that $d_n>mn^2(1/2-\kappa)$ for $n$ sufficiently large that is $n\ge \aleph(\kappa,\epsilon)$. After straightforward computations, we can choose
\begin{equation}\label{eq:eta}
\aleph(\kappa, \epsilon):=\Big\lfloor\sqrt{\frac{|\log(2\epsilon)|}{m\kappa}}+\Big| \frac{1}{2\kappa}-\frac{\log(2\varphi(0))}{m\kappa} \Big|\Big\rfloor+1.
\end{equation}
Markov's inequality leads to
\begin{equation}\label{eq:mark}
\mathbb{P}(\mathcal{N}_\epsilon>n)\le \mathbb{P}(|Z_{n-1}-\mathbb{E}[Z_{n-1}]|>d_n)\le \frac{\mathbb{E}[(Z_{n-1}-\mathbb{E}[Z_{n-1}])^4]}{d_n^4}.
\end{equation}
Let us note that $\overline{X}_j:=X_j-m$ are i.i.d. random variables with finite moments of any order. We denote $m_k:=\mathbb{E}[\overline{X}_j^k]$. Therefore we obtain
\begin{align}
\label{eq:mom}
\mathcal{Z}_{n-1}&:=\mathbb{E}[(Z_{n-1}-\mathbb{E}[Z_{n-1}])^4]=\mathbb{E}\Big[ \Big(\sum_{k=0}^{n-1}\sum_{j=0}^k \overline{X}_j\Big)^4 \Big]=\mathbb{E}\Big[ \Big(\sum_{j=0}^{n-1}(n-j) \overline{X}_j\Big)^4 \Big]\nonumber\\
&=\sum_{j=0}^{n-1}(n-j)^4m_4+2\sum_{0\le j<k\le n-1}(n-j)^2(n-k)^2m_2^2\nonumber\\
&\le \frac{m_4}{30}\, n(n+1)(6n^3+9n^2+n-1)+\frac{m_2^2}{36}\, n^2(n+1)^2(2n+1)^2.
\end{align}
Hence, there exist a constant $C_0>0$ such that $\mathbb{E}[(Z_{n-1}-\mathbb{E}[Z_{n-1}])^4]\le C_0 n^6$. Combining the previous inequality with \eqref{eq:eta} and \eqref{eq:mark} leads to
\[
\mathbb{P}(\mathcal{N}_\epsilon>n)\le \frac{C_0}{m^4(1/2-\kappa)^4}\,\frac{1}{n^2},\quad \mbox{for}\quad n\ge \aleph(\kappa,\epsilon).
\]
Consequently, the following upper-bound holds
\[
\mathbb{E}[\mathcal{N}_\epsilon]=\sum_{n\ge 0}\mathbb{P}(\mathcal{N}_\epsilon>n)\le \aleph(\kappa,\epsilon)+\frac{C_0}{m^4(1/2-\kappa)^4}\sum_{n\ge \aleph(\kappa,\epsilon)}\frac{1}{n^2}.
\]
In order to conclude, it suffices to note that $\aleph(\kappa,\epsilon)\to \infty$ as $\epsilon\to 0$, the second term in the previous inequality therefore becomes small as $\epsilon\to 0$: the leading term is finally $\aleph(\kappa,\epsilon)$ which is equivalent to $\sqrt{|\log(2\epsilon)|/(m\kappa)}$ by \eqref{eq:eta}.
\end{proof}
\section{First-passage time to boundaries with bounded derivative}
\label{sec:gen}
The algorithm presented in Section \ref{sec:incre} is simple to achieve (it only requires independent Gaussian random variables)
and efficient: the averaged number of steps is of the order $\sqrt{|\log\epsilon|}$ where $\epsilon$ stands for the small parameter
appearing in the rejection sampling (see Theorem \ref{thm1}). In order to apply Algorithm \ref{algo1} the curved boundary, the
Brownian motion is going to hit, has to satisfies suitable conditions: \eqref{hypo}, \eqref{hypo+} and \eqref{hypo++}.
Asking for the monotonicity of the function $\varphi$ is quite restrictive, that's why we present an extension of the algorithm
which is of course less efficient (even if the average number of steps is still very small) but which permits us to deal with more general boundaries.
Let us introduce the following assumption: there exist two constants $\rho_+>0$ and $\rho_->0$ such that
\begin{align}
\label{hypo-new}
\varphi:\mathbb{R}_+\to \mathbb{R}\ &\mbox{is a}\ \mathcal{C}^1\mbox{-continuous function satisfying}\nonumber\\ &\sup_{t\ge 0}\varphi'(t)\le \rho_+\quad\mbox{and}\quad \inf_{t\ge 0}\varphi'(t)\ge -\rho_-.
\tag{H4}
\end{align}
For such boundaries, we present an algorithm which permits us,
for any \(K\in\mathbb{R}^+\), to approximate the hitting time
$\tau_\varphi^K = \tau_\varphi\wedge K$, where \(\tau_\varphi\) is defined
in \eqref{def:tau}.
Let us introduce some notations: the \emph{inverse Gaussian distribution} of parameters $\mu>0$ and $\lambda>0$ will be denoted by $I(\mu,\lambda)$ and \ is defined by its
the probability distribution function:
\[
f(x)=\sqrt{\frac{\lambda}{2\pi x^3}}\,\exp-\Big\{\frac{\lambda(x-\mu)^2}{2\mu^2x}\Big\}\ \mathbbm{1}_{\{x\ge 0\}}.
\]
\begin{Algo}
\label{algo2}
Let $\epsilon>0$ be a small parameter and $r>\rho_-$ where $\rho_-$ is defined in \eqref{hypo-new}.\\
Initialization: $(T,H)=(0,\varphi(0))$ and $\mathcal{N}_{\epsilon,K}=0$.\\
While $H>\epsilon$ and \(T < K\), simulate $\hat{G}$ an inverse Gaussian random variable with distribution $I(H/r,H^2)$ and do:
\begin{align}\label{eq:algo2}
\left\{\begin{array}{l}
H\leftarrow\varphi(T+\hat{G})-\varphi(T)+
r\,
\hat{G},\\
T\leftarrow \hat{G}+T,\\
\mathcal{N}_{\epsilon,K}\leftarrow
\mathcal{N}_{\epsilon,K}+1.
\end{array}\right.
\end{align}
Outcome: $\tau^{\epsilon,K}_\varphi\leftarrow T\wedge K$ and $\mathcal{N}_{\epsilon,K}$.
\end{Algo}
Algorithm \ref{algo2} is quite simple, it only requires the simulation of inverse Gaussian distributed random variables. Let us recall the following scaling property: if $\hat{G}\sim I(H/r,H^2)$ then $H\hat{G}/r\sim I(1,r H)$. Moreover $\frac{(r\hat{G}-H)^2}{\hat{G}}$ is Chi-squared distributed with one degree of freedom (the square of a standard Gaussian random variable). In order to simulate an inverse Gaussian random variable, we suggest to use the algorithm introduced by Michael, Schucany and Haas (see \cite{MSH} or \cite[p.~149]{devroye}). Let us now state the efficiency of Algorithm \ref{algo2}. The inverse Gaussian distribution does not permit us to argue in a similar way as in Section \ref{sec:incre}. That's why we are going to use the general potential theory in order to upper-bound the averaged number of steps. This kind of arguments was already introduced in convergence results associated to the Random Walk on Spheres algorithm which permits
the approximation of the solution of the Dirichlet problem, see for instance \cite{muller}.
\begin{thm}\label{thm2}\begin{enumerate}
\item Let us assume that the boundary function $\varphi$ satisfies \eqref{hypo-new} then the random variable $\tau_\varphi^{\epsilon,K}$ defined in Algorithm \ref{algo2} converges in distribution towards $\tau_\varphi^K=\tau_\varphi\wedge K$ where $\tau_\varphi$ is defined by \eqref{def:tau} as $\epsilon$ tends to zero. More precisely
\begin{equation}\label{eq:thm:21}
F_{\epsilon,K}(t-\epsilon)-(1+\rho)
\sqrt{\frac{2\epsilon}{\pi}}\le F_K(t)\le F_{\epsilon,K}(t),\quad\mbox{for any}\quad t\ge \epsilon,
\end{equation}
where $F_K$ (resp. $F_{\epsilon,K}$) is the cumulative distribution function of $\tau_\varphi^K$ (resp. $\tau_\varphi^{\epsilon,K}$).
\item There exist positive constants $a$, $b$, $\kappa_0$, $\kappa_1$ and $\epsilon_0$ such that: for any $\rho_+\le \kappa_0$ and any $(K,r)$ satisfying $(r+\kappa_0)K\le \kappa_1$, the random number of iterations $\mathcal{N}_{\epsilon,K}$ defined in Algorithm \ref{algo2} satisfies the following upper bound
\begin{equation}
\label{eq:thm:conv12}
\mathbb{E}[\mathcal{N}_{\varepsilon,K}]\le (a+br)|\log \epsilon|,\quad \forall \epsilon\le \epsilon_0.
\end{equation}
\item For non increasing functions $\varphi$: there exists two positive constants $a$ and $\epsilon_0$ such that
\begin{equation}
\label{eq:thm:conv13}
\mathbb{E}[\mathcal{N}_{\varepsilon,K}]\le ar^2K|\log \epsilon|,\quad \forall \epsilon\le \epsilon_0.
\end{equation}
\end{enumerate}
\end{thm}
This theorem is based on the following intermediate statement which is a modification of Proposition \ref{prop:temps}.
\begin{prop}
\label{prop:2}
Let $(B_t,\ t\ge 0)$ be a standard one-dimensional Brownian motion. We introduce the following stopping times: $s_0=\mathcal{T}_0^K=0$ and for any $n\ge 1$:
\begin{equation}
\label{eq:prop:temps2}
s_n:=\inf\Big\{ t\ge 0:\ B_{t+\mathcal{T}_{n-1}^K}=
\varphi(\mathcal{T}_{n-1}^K)-r t \Big\}\quad\mbox{and}\quad \mathcal{T}_n^K:=(s_1+\ldots+s_n)\wedge K,
\end{equation}
where the boundary $\varphi$ satisfies \eqref{hypo-new}. Then the following properties hold:
\begin{enumerate}
\item $(\mathcal{T}_n^K)_{n\ge 0}$ is a non-decreasing sequence which almost surely converges towards $\tau_\varphi^K$.
\item On the event \(\{s_1 + \cdots + s_n < K\}\), the probability distribution of $s_{n+1}$ given the $\sigma$-algebra $\mathcal{F}_n:=\sigma\{ \mathcal{T}_1^K,\ldots,\mathcal{T}_n^K\}$ is the inverse Gaussian distribution $I(\mathcal{H}_n/r,\mathcal{H}_n^2)$ with
\begin{equation}
\label{eq:def:hn}
\mathcal{H}_n:=\varphi(\mathcal{T}_{n}^K)
-\varphi(\mathcal{T}_{n-1}^K)+r s_n.
\end{equation}
\item Let $\mathcal{M}_\epsilon:=\inf\{n\ge 1:\ \varphi(\mathcal{T}_n^K)-B_{\mathcal{T}_n^K}\le \epsilon\}$,
\(\mathcal{M}^K:=\inf\{n\ge 1,\ \mathcal{T}_n^K = K\}\),
\(\mathcal{M}_{\epsilon}^K=\mathcal{M}_\epsilon \wedge \mathcal{M}^K\).
then $\mathcal{T}_{\mathcal{M}_\epsilon^K}$ and $\tau_\varphi^{\epsilon,K}$, defined in Algorithm \ref{algo2}, are identically distributed, so are $\mathcal{M}_\epsilon^K$ and $\mathcal{N}_{\epsilon,K}$.
\end{enumerate}
\end{prop}
\begin{proof}[Proof of Proposition \ref{prop:2}] The first and the third part of the proof are left to the reader. They need similar arguments as those presented in Proposition \ref{prop:temps}.
Here the monotonicity property is just replaced by \eqref{hypo-new} which permits us easily to prove that $\mathcal{T}_n^K\le \tau_\varphi^K$.\\
Let us now focus our attention to the second part of the statement. Due to the definition of $s_{n+1}$ and since $\{\mathcal{T}_n^K<K\}$, we get $B_{\mathcal{T}_n^K}=\varphi(\mathcal{T}_{n-1}^K)-rs_n$. Hence, we have
\begin{align*}
s_{n+1}&=\inf\{t\ge 0:\ B_{t+\mathcal{T}_n^K}-B_{\mathcal{T}_n^K}=\varphi(\mathcal{T}_n^K)-B_{\mathcal{T}_n^K}-r t\}\\
&=\inf\{t\ge 0:\ W_t=\mathcal{H}_n -r t \},
\end{align*}
where $W_t=B_{t+\mathcal{T}_n^K}-B_{\mathcal{T}_n^K}$ is a standard Brownian motion independent of
$\mathcal{F}_n$
and the $\mathcal{F}_n$ adapted r.v. $\mathcal{H}_n$
is defined by \eqref{eq:def:hn}. The distribution of $s_{n+1}$ corresponds
to the distribution of the first passage time of the standard Brownian motion
with drift at the constant level $\mathcal{H}_n$. The probability distribution is well known (see, for instance
\cite[p.~197]{karatzas}):
\[
\mathbb{P}(s_{n+1}\in dt|\mathcal{F}_n)=\frac{\mathcal{H}_n}{\sqrt{2\pi t^3}}\,\exp -\Big\{ \frac{(\mathcal{H}_n-r t)^2}{2t} \Big\}\, dt,
\]
we can consequently identify the inverse Gaussian distribution $I(\mathcal{H}_n/r,\mathcal{H}_n^2)$.
\end{proof}
\begin{proof}[Proof of Theorem
\ref{thm2}] ~\\
\textbf{Step 1.} We can prove the convergence in distribution of $\tau_\varphi^{\epsilon,K}$ towards $\tau_\varphi^K$ using similar arguments as those presented in the proof of Theorem \ref{thm1}. The upper-bound in \eqref{eq:thm:21} is an adaptation of \eqref{eq:repar-facil} which requires that $\mathcal{T}_n^K\le \tau_\varphi^K$ and that $\tau_\varphi^{\epsilon,K}$ and $\mathcal{T}_{\mathcal{M}_\epsilon^K}$ are identically distributed. These conditions are satisfied, see Proposition~\ref{prop:2}. For the lower-bound in \eqref{eq:thm:21}, we obtain
\[
F_{\epsilon,K}(t-\epsilon)\le \mathbb{P}(|\mathcal{T}_{\mathcal{M}_\epsilon^K}
-\tau_\varphi^K|>\epsilon)+F_{\epsilon,K}(t),
\]
see \eqref{eq:etap1} for the details. Let us note that $|\tau_{\mathcal{M}_\epsilon^K}-\tau_\varphi^K|>\epsilon$ leads to the condition $\tau_{\mathcal{M}_\epsilon^K}<K$. Hence $\mathcal{M}_\epsilon^K=\mathcal{M}_\epsilon$. Using the Markov property, the following bound holds:
\[
\mathbb{P}(|\mathcal{T}_{\mathcal{M}_\epsilon^K}
-\tau_\varphi^K|>\epsilon)\le 1-\mathbb{P}\Big( \sup_{0\le u\le \epsilon}\tilde{B}_u\ge \epsilon+\sup_{0\le u\le \epsilon}\varphi(\mathcal{T}_{
\mathcal{M}_\epsilon}+u)-\varphi(\mathcal{T}_{
\mathcal{M}_\epsilon}) \Big).
\]
Here $(\tilde{B}_t,\, t\ge 0)$ stands for a standard Brownian motion independent of $\mathcal{T}_{\mathcal{M}_\epsilon}$.
Combining Hypothesis \eqref{hypo-new} and the reflection principle of the Brownian motion leads to
\begin{align*}
\mathbb{P}(|\mathcal{T}_{\mathcal{M}_\epsilon}
-\tau_\varphi^K|>\epsilon)&\le 1-\mathbb{P}(|\tilde{B}_\epsilon|\ge \epsilon(1+\rho_+))\le (1+\rho_+)\sqrt{\frac{2\epsilon}{\pi}},
\end{align*}
and consequently to the lower bound \eqref{eq:thm:21}.\\
\textbf{Step 2.} Let us now focus our attention to the averaged number of steps in Algorithm \ref{algo2}, denoted by $\mathcal{N}_{\epsilon,K}$. A rough description of the method: we aim to construct a Markov chain and to describe the associated potential. The classical potential theory then permits us to obtain the announced bound. We introduce the Markov chain $R_n:=(\mathcal{T}_n,\mathcal{H}_n)$ for $n\ge 0$.
We recall that $\mathcal{T}_n=s_1+\ldots+s_n$ is defined by \eqref{eq:prop:temps2} and $\mathcal{H}_n$ by \eqref{eq:def:hn}. The stopping time $\mathcal{M}^K_\epsilon$ defined in Proposition \ref{prop:2} can also be interpreted as the first time the Markov chain $(R_n,\ n\ge 0)$ goes out of the domain $E:=[0,K]\times ]\epsilon,+\infty]$. \\
Let us consider the function $f(x,y)=\log(y)$, defined on $E$, and denote by $P$ the infinitesimal generator associated to the Markov chain $(R_n)_{n\ge 0}$. By Proposition~\ref{prop:2} and for any $(t,h)\in E$, we obtain
\begin{align*}
Pf(t,h)=\mathbb{E}\Big[\log(\varphi(t
+\hat{G})-\varphi(t)+r \,\hat{G})\Big],
\end{align*}
where $\hat{G}$ is an inverse Gaussian distributed random variable with the following density function:
\[
p(x)=\frac{h}{\sqrt{2\pi x^3}}\exp\left\{ -\frac{(h-r x)^2}{2x} \right\},\quad x\ge 0.
\]
By \eqref{hypo-new}, $\varphi(t+\hat{G})-\varphi(t)\le \rho_+\, \hat{G}$, we get
\begin{align}\label{eq:prem}
Pf(t,h)-f(t,h)&\le \log\Big(1+\frac{\rho_+}{r}\Big)+\mathbb{E}\Big[\log\Big(\frac{
r \hat{G}}{h}\Big)\Big].
\end{align}
Let us find now an explicit upper bound of $Pf-f$.
Using first the change of variables $u=r x/h$ and secondly $u\mapsto 1/u$, we get
\begin{align}\label{eq:calc}
\mathbb{E}\Big[\log\Big(\frac{r \hat{G}}{h}\Big)\Big]&=\int_0^\infty \log\Big( \frac{r x}{h} \Big)\frac{h}{\sqrt{2\pi x^3}}\exp-\frac{(h-r x)^2}{2x}\, dx \nonumber\\
&=\sqrt{\frac{hr}{2\pi}}\int_0^\infty \frac{\log(u)}{u^{3/2}}\, \exp-\frac{hr(1-u)^2}{2u}\, du\nonumber\\
&=\sqrt{\frac{h r}{2\pi}}\int_1^\infty \frac{(1-u)\log(u)}{u^{3/2}}\exp-\frac{h r(1-u)^2}{2u}\, du.
\end{align}
It is then obvious that $\mathbb{E}\Big[\log\Big(\frac{r\hat{G}}{h}\Big)\Big]< 0$. Let us now give a more precise upper-bound.
We set $\alpha=h r$, then \eqref{eq:calc} emphasizes that $\mathbb{E}\Big[\log\Big(\frac{r\hat{G}}{h}\Big)\Big]$ only depends
on the parameter $\alpha$, this dependence being continuous. Let us therefore denote this function $\psi(\alpha)$
(see Figure~\ref{fig:MCpsi} below representing $\psi$ obtained with the Monte-Carlo method sample size: $10\,000$).\\
\begin{figure}[t]
\begin{center}
\includegraphics[scale=0.4]{psi.jpg}
\end{center}
\caption{Monte Carlo approximation of the function \(\psi\)}\label{fig:MCpsi}
\end{figure}
Simple computations lead to
\begin{align}\label{eq:alpha}
\psi(\alpha)&:=\mathbb{E}\Big[\log\Big(\frac{r \hat{G}}{h}\Big)\Big]=-\sqrt{\frac{\alpha}{2\pi}}\int_0^\infty \frac{u\log(1+u)}{(1+u)^{3/2}}\exp-\frac{\alpha u^2}{2(1+u)}\, du\\
&\le -\sqrt{\frac{\alpha}{2\pi}}\int_0^\infty \frac{u\log(1+u)}{(1+u)^{3/2}}\exp-\frac{\alpha u}{2}\, du
\nonumber\\
&\le
-\frac{1}{\sqrt{2\pi}}\int_0^\infty \frac{w\log(1+w/\alpha)}{(\alpha+w)^{3/2}}\, \exp-\frac{w}{2}\,dw\nonumber\\
&\le
-\frac{1}{\sqrt{2\pi}}\int_{1/2}^\infty \frac{w\log(1+w/\alpha)}{(\alpha+w)^{3/2}}\, \exp-\frac{w}{2}\,dw.\nonumber\end{align}Using the inequality $(\alpha+w)\le (1+2\alpha)w$, we get
\begin{align*}
\psi(\alpha)
& \le -\frac{\log(1+(2\alpha)^{-1})}{(1+2\alpha)^{3/2}\sqrt{2\pi}}\int_{1/2}^\infty \frac{1}{\sqrt{w}}\exp-\frac{w}{2}\,dw
\\
&\le -\frac{\log(1+(2\alpha)^{-1})}{(1+2\alpha)^{3/2}}\,\mathbb{P}(G\ge 1/2),
\end{align*}
where $G$ is a standard gaussian r.v. and so $\mathbb{P}(G\ge 1/2)\approx 0.3085$\\
We deduce from the previous upper-bound that $\lim_{\alpha\to 0^+}\psi(\alpha)=-\infty$. Moreover the right hand side is a non decreasing function with respect to the variable $\alpha$. Hence
\begin{equation}\label{eq:num1}
\psi(\alpha)\le -\frac{\log(3/2)}{3\sqrt{3}}\,\mathbb{P}(G\ge 1/2)\approx -0.0241,\quad\mbox{for}\ \alpha\le 1.
\end{equation}
Let us observe what happens for large values of the variable $\alpha$. The Laplace method implies that
\[
\psi(\alpha)\sim -\frac{1}{2\alpha}\quad \mbox{as}\ \alpha\to \infty.
\]
Let us prove now that there exists a constant $c>0$ such that
\begin{equation}\label{eq:num2}
\psi(\alpha)\le -\frac{c}{\alpha},\quad \mbox{for any}\ \alpha\ge 1.
\end{equation}
For $\alpha\ge 1$, we get
\begin{align*}
\psi(\alpha)&\le-\sqrt{\frac{\alpha}{2\pi}}\int_0^\infty \frac{u\log(1+u)}{(1+u)^{3/2}}\,\exp-\frac{\alpha u^2}{2}\,du\\
&\le -\sqrt{\frac{\alpha}{2\pi}}\int_0^1 \frac{u\log(1+u)}{(1+u)^{3/2}}\,\exp-\frac{\alpha u^2}{2}\,du.
\end{align*}
Due to the convexity property of the logarithm function ($\log (1+u)\ge \log(2) u$) and the Cauchy-Schwarz inequality, we obtain
\begin{align*}
\psi(\alpha)&\le -\frac{\log(2)}{\alpha 2^{3/2}}\Big(\frac{1}{2}\mathbb{E}[G^2]-\mathbb{E}[G^21_{\{G\ge \sqrt{\alpha}\}}] \Big)\\
&\le -\frac{\log(2)}{\alpha 2^{3/2}}\Big( \frac{1}{2}-\sqrt{\mathbb{E}[G^4]}\sqrt{\mathbb{P}(G\ge \sqrt{\alpha})} \Big)\\
&\le -\frac{\log(2)}{\alpha 2^{3/2}}\Big( \frac{1}{2}-\frac{\sqrt{3}}{2}\,e^{-\alpha} \Big)\le -\frac{\log(2)}{\alpha 2^{5/2}}(1-\sqrt{3}e^{-1}),\quad\mbox{for}\ \alpha\ge 1.
\end{align*}
We deduce that $\psi(\alpha)\le -c/\alpha$ with $c\approx 0.0445$ when $\alpha\ge 1$. Combining both inequalities \eqref{eq:num1} and \eqref{eq:num2} leads to the existence of a constant $c>0$ such that
\begin{equation}
\label{eq:upper}
\psi(\alpha)\le -c\ \Big( \frac{1}{\alpha}\wedge 1\Big).
\end{equation}
By \eqref{eq:prem}, the following upper-bound holds: for $f(x,y)=\log(y)$,
\begin{align}
\label{eq:bou}
Pf(t,h)-f(t,h)&\le \log\Big( 1+\frac{\rho_+}{r} \Big)-c\Big(\frac{1}{hr}\wedge 1\Big)\nonumber\\
&\le \frac{\rho_+}{r}-c\Big(\frac{1}{hr}\wedge 1\Big),\quad h\ge 0,\ t\ge 0.
\end{align}
Due to the definition of $\rho_+$, we know that
\[
h\le \varphi(0)\vee(r+\rho_+)t\le \varphi(0)\vee(r+\rho_+)K,
\]
where $\varphi$ is the boundary the process has to hit.
In other words, there exist two constants $\kappa_0>0$ and $\kappa_1>0$ such that for any $\rho_+\le \kappa_0$ and any $(K,r)$ satisfying $(r+\kappa_0)K\le \kappa_1$ the following bound holds
\(
\rho_+\le \frac{c}{2}\,\Big(\frac{1}{h}\wedge r\Big).
\)
Hence:
\[
Pf(t,h)-f(t,h)\le -\frac{c}{2r}\Big(\frac{1}{\varphi(0)\wedge \kappa_1}\wedge r\Big)=:-\mathcal{R}^{-1}(r).
\]
We deduce that the function $g(t,h)$ defined by $g(t,h)=\mathcal{R}(r)\,(f(t,h)-\log\epsilon)$ satisfies $g(t,h)\ge 0$ for any $(t,h)\in E$ and $Pg(t,h)-g(t,h)\le -1$ on $E$. The potential theory therefore implies:
\[
\mathbb{E}[\mathcal{N}_{\epsilon,K}]\le g(0,\varphi(0))\le \mathcal{R}(r)(\log(\varphi(0))-\log(\epsilon)).
\]
We finally deduce the existence of $a>0$ and $b>0$ such that $\mathbb{E}[\mathcal{N}_{\varepsilon,K}]\le (a+br)|\log\epsilon|$ for $\epsilon$ small enough.\\
For the particular case of a non increasing boundary function it suffices to vanish $\rho_+$ in \eqref{eq:bou} and to apply the same arguments of the potential theory in order to get \eqref{eq:thm:conv13}
\end{proof}
\section{Examples and numerics.}\label{sec:examples}
In this section, we present three different examples which nicely illustrate the efficiency of these new algorithms \ref{algo1} and \ref{algo2}.
\mathversion{bold}
\subsection{Brownian hitting time of $\varphi(t)=\sqrt{1+\alpha t}$ }
\mathversion{normal}
Let us first consider an application of Theorem \ref{thm1}. We observe that $\varphi(t)=\sqrt{1+\alpha t}$ is an increasing function satisfying \eqref{hypo}, \eqref{hypo+} and \eqref{hypo++} for $\alpha\in [0,1]$. Consequently Algorithm \ref{algo1} converges and permits us to obtain an approximation of the hitting time $\tau_\varphi$. In the figures, we present the link between the averaged number of steps and $\epsilon$ which characterizes the approximation error size.\\
The first figure (resp. the second one) concerns: $\alpha=1$ (resp. $\alpha=0.01$), $\epsilon=0.5^n$ ($n$ is represented on the horizontal axis) and the number of simulation in order to estimate the averaged number of steps is $10\,000$.
\begin{figure}[ht]
\begin{center}
\subfloat[{$\alpha = 1$}]{
\includegraphics[width=0.48\textwidth]{mean_steps_alpha1}
}
\subfloat[{$\alpha = 0.01$}]{
\includegraphics[width=0.48\textwidth]{mean_steps_alpha001}
}
\end{center}
\caption{\(\mathbb{E}(\mathcal{N}_\epsilon)\): mean number of steps for \(\epsilon = 0.5^n\) as a function of \(n\). The boundary is
$\varphi(t)=\sqrt{1+\alpha t}$.}\label{fig:nombredetapetmoyen}
\end{figure}
Let us now present the approximate distribution of the hitting time.
\begin{figure}[ht]
\begin{center}
\subfloat[{$\alpha = 1$}]{
\includegraphics[width=0.48\textwidth]{distrib_tau_alpha1}
}
\subfloat[{$\alpha = 0.01$}]{
\includegraphics[width=0.48\textwidth]{distrib_tau_alpha001}
}
\end{center}
\caption{Empirical distribution of the approximate first hitting time of the boundary
$\varphi(t)=\sqrt{1+\alpha t}$.}\label{fig:distribfht}
\end{figure}
\mathversion{bold}
\subsection{Brownian hitting time of $\varphi(t)=\alpha+\beta\cos(\omega t)$ }
\mathversion{normal}
Let us now consider the first time the Brownian motion hits the periodic boundary $\varphi(t)=\alpha+\beta\cos(\omega t)$. Since the boundary is not an increasing function,
we shall use Algorithm \ref{algo2}. Theorem \ref{thm2} ensures that the algorithm converges. Let us therefore use the Monte-Carlo method in order to estimate precisely
the average number of steps.
As explained in the previous section, the simulation procedure permits to approximation of the stopping time $\tau_\varphi\wedge K$ for some given fixed time $K$.
Figure~\ref{fig:mean_number_cos} illustrates the approximation \(\tau_\varphi\) by \(\tau^{\epsilon,K}_\varphi\), where the parameters are
fixed at $\alpha=3.5$, $\beta=3$ and $\omega=\pi/2$.
The maximal time are $K=20$ on one hand and \(K=100\) on the other hand and the error rate is given by $\epsilon=0.5^n$, for \(1\leq n\leq 10\).
A sample of $10E8$ paths has been simulated to approximate the mean.
\begin{figure}[h]
\begin{center}
\subfloat[\(\mathbb{E}(\mathcal{N}_{1/2^n,K})\) versus \(n\)]{
\includegraphics[width=0.48\textwidth]{meannbstepscos_tmax20_100.jpg}}
\subfloat[Distribution of \(\tau^{1/2^{n},K}_{\varphi}\). \(n=10\), \(K=20\).]{
\includegraphics[width=0.48\textwidth]{distrib_cos_tmax20.jpg}}
\end{center}
\caption{Approximation of \(\tau_{\varphi}\) with $\varphi(t)=3.5+3\cos(\pi t/2)$ }\label{fig:mean_number_cos}
\end{figure}
We know that the mean number of steps is a decreasing function of \(\epsilon\) and an increasing function of \(K\).
Figure~\ref{fig:effetoftruncation} gives the evolution of the mean number of steps as a function of the truncation \(K\).
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.6\textwidth]{meannbstepscos_fndek_full.jpg}
\caption{\(\mathbb{E}(\mathcal{N}^{\epsilon,K}_\varphi)\) as a function of \(K\).}\label{fig:effetoftruncation}
\end{center}
\end{figure}
In practice, we obtained easily an impressively accurate approximation of \(\tau_\varphi\).
\subsection{The first time the Ornstein Uhlenbeck process hits the boundary $\varphi(t)=\alpha+\beta\cos(\omega t)$ }
\mathversion{normal}
This last section concerns a particular framework where passage times play a crucial role: the spiking neuron analysis. Let us roughly explain how neuronal firing activities has been modeled. The potential difference that exists across the cell membrane is modeled as an Ornstein-Uhlenbeck process. As soon as this membrane potential exceeds a given threshold, the neuron releases a rapid electrical signal called a spike and the membrane potential is directly reset to an initial voltage. Hence the interspike interval is identify with the first hitting time of an OU process whereas the spike train forms a renewal process. Such a stochastic leaky integrate and fire (LIF) neuronal model is a good compromise between realism and mathematical tractability \cite{Burkitt-2006, Gerstner, Tuckwell-1989, Tuckwell-Wan-2000}.
The standard OU process can be adapted when the inputs are time-dependent. It especially concerns many situations where the sensory stimuli, like sound, contain an oscillatory component. We observe then oscillating membrane potentials in the neuron, generating rhythmic spiking patterns (see Iolov, Ditlevsen and Longtin \cite{Iolov-Ditlevsen-Longtin-2014} and references therein).
In such a model with time-dependent forcing, the membrane voltage denoted by $(V_t,\,t\ge 0)$ satisfies the following stochastic differential equation:
\[
dV_t=\Big( \chi(t)-\frac{V_t}{\tau}\Big)\, dt+\sigma\, dB_t,
\]
until it reaches the voltage threshold $V_{\rm th}$. Here $\chi$ represents a current acting on the cell, $\tau$ is the membrane time constant, $\sigma$ is
the strength of the stochastic fluctuations and finally $(B_t,\, t\ge 0)$ stands for the standard Brownian motion. The length of the interspike interval is therefore directly related to the passage time of the stochastic process $(V_t)$ through the threshold $V_{\rm th}$.
Let us introduce a simple change of variable, given by $X_t=V_t-\varphi(t)$ with $\varphi$ the deterministic function satisfying:
\[
\varphi'(t)=\chi(t)-\frac{\varphi(t)}{\tau},\quad \varphi(0)=V_0.
\]
By straightforward computations, we can prove that $X_t$ is a classical OU process. In other words, the first passage time of the voltage $V_t$ through the given threshold $V_{\rm th}$ is almost surely equal to the first passage time of the OU process $(X_t)$ through the curved boundary $\varphi$. Simulating hitting times to curved boundaries for OU processes is therefore a main task.
That's why, we focus our attention to the last example which concerns the one-dimensional Ornstein-Uhlenbeck process defined by:
\begin{equation}\label{eq:orn}
dX_t=dB_t-\lambda X_t\,dt,\quad X_0=x_0.
\end{equation}
The aim is to approximate the first passage time through the particular simple curved boundary $\varphi(t)=\alpha+\beta\cos(\omega t)$ where $\varphi(0)>x_0$.
Since the Ornstein-Uhlenbeck process can be represented as a time-changed Brownian motion, the question is directly related to the main results of this study. Indeed the solution of \eqref{eq:orn} is given by
\[
X_t=e^{-\lambda t}\left( x_0+\int_0^t e^{\lambda s}dB_s \right),\quad t\ge 0.
\]
Using Levy's theorem, $(X_t,\, t\ge 0)$ has the same distribution as $(Y_t,\, t\ge 0)$ defined by
\[
Y_t:=e^{-\lambda t}\left(x_0+W_{u(t)} \right),\quad t\ge 0,
\]
with $u(t):=\frac{1}{2\lambda}\,(e^{2\lambda t}-1)$ and $W$ a standard Brownian motion. We deduce that
\[
\mathcal{T}_\varphi:=\inf\{t\ge 0:\ X_t=\varphi(t)\}
\]
has the same distribution as
\begin{align*}
\hat{\mathcal{T}}_\varphi & :=\inf\Big\{ t\ge 0:\ e^{-\lambda t}\Big(x_0+W_{u(t)}\Big)=\varphi(t) \Big\}\\
&=\inf\Big\{u^{-1}(s)\ge 0:\ W_s=\varphi(u^{-1}(s)) e^{\lambda u^{-1}(s)}-x_0\Big\}\\
&=u^{-1}(\tau_\psi),
\end{align*}
where
\[
\tau_\psi:=\inf\{t\ge 0:\, W_t=\psi(t)\},\quad \psi(t):=\sqrt{1+2\lambda t}\ \varphi\Big(\frac{\log(1+2\lambda t)}{2\lambda}\Big)-x_0.
\]
Consequently, in order to simulate the Ornstein-Uhlenbeck hitting time $\mathcal{T}_\varphi\wedge K$ for some $K$, we simply use Algorithm \ref{algo2} and propose an approximation of the Brownian hitting time $\tau_\psi\wedge \tilde{K}$ with $\tilde{K}:=u(K)=(e^{2\lambda K}-1)/(2\lambda)$.\\
Let us note that a straightforward computation leads to the following upper-bound:
\[
|\psi'(t)|\le \frac{\lambda\alpha+\lambda\beta+\omega\beta}{\sqrt{1+2\lambda t}}\le \lambda\alpha+\lambda\beta+\omega\beta,\quad t\ge 0.
\]
In other words, the continuous curve $\psi$ satisfies Hypothesis \eqref{hypo-new}: Algorithm \ref{algo2} therefore converges and Theorem \ref{thm2} can be applied.\\
In the following numerical experiences, we will choose $r=0.5+\lambda\alpha+\lambda\beta+\omega\beta$.
Figures~\ref{fig:premiereou} and \ref{fig:oudeuxieme} concern the following choice of parameters: $x_0=0$, $\alpha=2$, $\beta=1$, $\omega=\pi/5$, $\lambda=0.5$.
We have chosen $K=5$ for Figure~\ref{fig:premiereou} and $K=10$ for Figure~\ref{fig:oudeuxieme}.
In both cases, the first figure represents the average number of steps as a function of \(n\) where the approximation parameter \(\epsilon\) is chosen as
\(0.5^n\), for \(n=1,\cdots,10\).
The average has been estimated using $5.10E6$ simulations.
The second figure represents the distribution of $\mathcal{T}_\varphi\wedge K$ for \(n=10\).
We observe that the \textit{change of time} \(\tilde{K} = (e^{2\lambda K}-1)/(2\lambda)\) increases very fast with \(K\) and the number becomes quite large when \(K\) increases.
Note however that the number of random variables we have to simulate keeps relatively small in comparaison with the use of
a classical stopped Euler scheme usually used to approximate \(\mathcal{T}_\varphi\).
Each Euler scheme introduces a bias, which goes to \(0\) as the time discretization length goes to \(0\). We have plotted the error on
Figure~\ref{fig:compare_euler_schemes} for the approximation of \(\mathbb{E}(\mathcal{T}_\varphi^K)\). We can observe that a simple
Euler scheme has an order of convergence \(1/2\). We illustrate the improvement of this rate of convergence using two
particular modifications of the scheme: the first one from \cite{Gobet-2000} (i.e. we take into account the first order term of
the probability to hit the boundary between two time successive time-steps), the second-one from \cite{Gobet-Menozzi-10} (i.e. an adapted modification of the boundary).
Both modifications yield to a scheme of order \(1\).
The numerical cost of our algorithm increases very slowly as the parameter \(\epsilon\) goes to \(0\).
The numerical comparisons are done with \(\epsilon=2^{-20}\), such that the error is almost negligible.
The time we need is similar to the time for an Euler scheme with step \(0.01\) and the Brownian bridge modification with time step \(0.02\).
Empirically, we conclude that our scheme over performs previous ones if one needs an accuracy larger than those obtained with
an Euler scheme with time step \(0.01\).
\begin{figure}[h]
\begin{center}
\subfloat[\(\mathbb{E}(\mathcal{N}_{1/2^n,K})\) versus \(n\)]{
\includegraphics[width=0.48\textwidth]{meannbstepsOUcos_t5.jpg}}
\subfloat[Distribution of \(\mathcal{T}^{K,\epsilon}_\varphi\) (\(\epsilon = 1/2^{10}\)).]{
\includegraphics[width=0.48\textwidth]{distribOUcost5.jpg}}
\end{center}
\caption{First hitting time of \(\varphi(t)=\alpha+\beta\cos(\omega t)\) by an Ornstein Uhlenbeck process solution of \eqref{eq:orn} ($\alpha=2$, $\beta=1$, $\omega=2\pi$, $\lambda=0.5$, \(K=5\).)}\label{fig:premiereou}
\end{figure}
\begin{figure}[h]
\begin{center}
\subfloat[\(\mathbb{E}(\mathcal{N}_{1/2^n,K})\) versus \(n\)]{
\includegraphics[width=0.48\textwidth]{meannbstepsOUcos.jpg}}
\subfloat[Distribution of \(\mathcal{T}^{K,\epsilon}_\varphi\) (\(\epsilon = 1/2^{10}\)).]{
\includegraphics[width=0.48\textwidth]{distribOUcost10.jpg}}
\end{center}
\caption{First hitting time of \(\varphi(t)=\alpha+\beta\cos(\omega t)\) by an Ornstein Uhlenbeck process solution of \eqref{eq:orn} ($\alpha=2$, $\beta=1$, $\omega=2\pi$, $\lambda=0.5$, \(K=10\)).}\label{fig:oudeuxieme}
\end{figure}
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.48\textwidth]{comparison_3_methods.pdf}
\end{center}
\caption{Bias for Euler schemes to approximate \(\mathbb{E}(\mathcal{T}_\varphi^K)\), the mean first hitting time of \(\varphi(t)=\alpha+\beta\cos(\omega t)\) by an Ornstein Uhlenbeck process solution of \eqref{eq:orn} ($\alpha=2$, $\beta=1$, $\omega=2\pi$, $\lambda=0.5$, \(K=5\)).}\label{fig:compare_euler_schemes}
\end{figure}
\clearpage
\bibliographystyle{abbrv}
|
\section{Overview}
We consider the geometric flow equations
\begin{align}\label{eq:flow}
\frac{d}{dt} X = -F\nu.
\end{align}
and ask whether closed strictly convex $2$-dimensional surfaces $M_t$ in ${\mathbb{R}}^3$
converge to round points or to spheres at $\infty$.
The answer is affirmative for many normal velocities $F$, including certain powers of
the Gauss curvature, $K$, the mean curvature, $H$,
and the norm of the second fundamental form, $|A|$.
Here, authors like B. Andrews \cite{ba:gauss,ac:surfaces}, O. Schn\"urer \cite{os:surfacesA2,os:surfaces}, and F. Schulze \cite{fs:convexity},
use functions of the principal curvatures $w$
to show convergence to a round point or to spheres at $\infty$.
In \cite{os:surfaces}, \cite{os:surfacesA2},
O. Schn\"urer proposes a characterization of
these functions.
And in this paper, we extend it to non-rational functions.
Our definition of \textit{maximum-principle functions} (MPF)
now covers any such function $w$ that is known to far.
The function from \cite{fs:convexity}
\begin{align*}
w=\frac{(a-b)^2(a+b)^{2\sigma}}{(a\,b)^2}
\end{align*}
is an example of a MPF for a normal velocity $F=H^\sigma$ with $1<\sigma\leq 5$.
Our main question is when MPF cease to exist
for the large set of contracting and expanding
normal velocities $F^\sigma_\xi$.
We are particularly interested in normal velocities
$F^\sigma_0=\sgn(\sigma)\cdot K^{\sigma/2}$,
$F^\sigma_1=\sgn(\sigma)\cdot H^\sigma$, and
$F^\sigma_2=\sgn(\sigma)\cdot |A|^\sigma$,
for all powers $\sigma\in{\mathbb{R}}\setminus\{0\}$.
Here, we either are able to prove the non-existence of MPF, or give an example
of a MPF. We present MPF from literature, and
newly discovered MPF.
Small gaps only exist for $H^\sigma$ with $5.17<\sigma<5.98$, and $|A|^\sigma$ with $8.15<\sigma<9.89$.
We summarize our main results in the table for $F^\sigma_\xi$ (p. \pageref{thm:necessary}), and in the tables
for $F^\sigma_0$ (p. \pageref{cor:gauss curvature}), for $F^\sigma_1$ (p. \pageref{cor:mean curvature}),
for $F^\sigma_2$ (p. \pageref{cor:|A|}), and for $F^\sigma_\sigma$ (p. \pageref{cor:trA}). \\
\textit{The paper is structured as follows:}
In the chapter on notation, we
give a brief introduction to differential geometric quantities
like induced metric, second fundamental form, and principal curvatures.
The next chapter is on $F^\sigma_\xi$.
This is the set of normal velocities for which we investigate
the existence and non-existence of MPF.
We proceed with a chapter on the MPF,
motivating the definition and taking a closer
look at the linear operator $L$ and the $\alpha$-conditions.
Chapter 5 contains main Theorem \ref{thm:necessary}.
Using Euler's Theorem on homogeneous functions
we are also able to prove the non-existence of MPF
for any normal velocity with $0<\hom F\leq 1$.
In chapter 6 we present vanishing functions,
which are sometimes MPF,
depending on the given normal velocity $F^\sigma_\xi$.
Furthermore, we present a few technical lemmas on
vanishing functions, which play an important role
in the following chapters
on $F^\sigma_0$, $F^\sigma_1$, $F^\sigma_2$, and $F^\sigma_\sigma$.
We conclude our paper with an outlook suggesting
an improved MPF Ansatz.
\section{Notation}
For a brief introduction of the standard notation we adopt the
a corresponding chapter from \cite{os:surfacesA2}.
We use $X=X(x,\,t)$ to denote the embedding vector of a $2$-manifold $M_t$ into ${\mathbb{R}}^3$ and
$\frac{d}{dt} X=\dot{X}$ for its total time derivative.
It is convenient to identify $M_t$ and its embedding in ${\mathbb{R}}^3$.
The normal velocity $F$ is a homogeneous symmetric function of the principal curvatures.
We choose $\nu$ to be the outer unit normal vector to $M_t$.
The embedding induces a metric $g_{ij} := \langle X_{,i},\, X_{,j} \rangle$ and
the second fundamental form $h_{ij} := -\langle X_{,ij},\,\nu \rangle$ for all $i,\,j = 1,\,2$.
We write indices preceded by commas to indicate differentiation with respect to space components,
e.\,g.\ $X_{,k} = \frac{\partial X}{\partial x_k}$ for all $k=1,\,2$.
We use the Einstein summation notation.
When an index variable appears twice in a single term it implies summation of that term over all the values of the index.
Indices are raised and lowered with respect to the metric
or its inverse $\left(g^{ij}\right)$,
e.\,g.\ $h_{ij} h^{ij} = h_{ij} g^{ik} h_{kl} g^{lj} = h^k_j h^j_k$.
The principal curvatures $a,\,b$ are the eigenvalues of the second fundamental
form $\left(h_{ij}\right)$ with respect to the induced metric $\left(g_{ij}\right)$. A surface is called strictly convex,
if all principal curvatures are strictly positive.
We will assume this throughout the paper.
Therefore, we may define the inverse of the second fundamental
form denoted by $(\tilde h^{ij})$.
Symmetric functions of the principal
curvatures are well-defined, we will use the Gauss curvature
$K=\frac{\det h_{ij}}{\det g_{ij}} = a\cdot b$, the mean curvature
$H=g^{ij} h_{ij} = a+b$, the square of the norm of the second fundamental form
$|A|^2= h^{ij} h_{ij} = a^2+b^2$, and the trace of powers of the second fundamental form
$\tr A^{\sigma} = \tr \left(h^i_j\right)^{\sigma} = a^\sigma+b^\sigma$. We write indices preceded by semi-colons
to indicate covariant differentiation with respect to the induced metric,
e.\,g.\ $h_{ij;\,k} = h_{ij,k} - \Gamma^l_{ik} h_{lj} - \Gamma^l_{jk} h_{il}$,
where $\Gamma^k_{ij} = \frac{1}{2} g^{kl} \left(g_{il,j} + g_{jl,i} - g_{ij,l}\right)$.
It is often convenient to choose normal coordinates, i.\,e.\ coordinate systems such
that at a point the metric tensor equals the Kronecker delta, $g_{ij}=\delta_{ij}$,
in which $\left(h_{ij}\right)$ is diagonal, $(h_{ij})=\diag(a,\,b)$.
Whenever we use this notation, we will also assume that we have
fixed such a coordinate system. We will only use a Euclidean metric
for ${\mathbb{R}}^3$ so that the indices of $h_{ij;\,k}$ commute according to
the Codazzi-Mainardi equations.
A normal velocity $F$ can be considered as a function of $(a,\,b)$
or $(h_{ij},\,g_{ij})$. We set $F^{ij}=\fracp{F}{h_{ij}}$,
$F^{ij,\,kl}=\fracp{^2F}{h_{ij}\partial h_{kl}}$.
Note that in coordinate
systems with diagonal $h_{ij}$ and $g_{ij}=\delta_{ij}$ as mentioned
above, $F^{ij}$ is diagonal.
\section{Contracting and expanding normal velocities $F^\sigma_\xi$}
In this chapter, we specify what we mean by a contracting and an expanding normal velocity
and define the important quantity $\beta=\frac{F_a}{F_b}$.
In Remark \ref{rem:normal velocity xi},
we introduce the set of normal velocities $F^\sigma_\xi$.
This set includes
powers of the Gauss curvature, $F^\sigma_0$,
powers of the mean curvature, $F^\sigma_1$,
powers of the norm of the second fundamental form, $F^\sigma_2$,
and the trace of powers of the second fundamental form, $F^\sigma_\sigma$.
Throughout this paper, our goal is to determine,
when MPF exist and cease to exist for $F^\sigma_\xi$.
\begin{definition}[Normal velocity $F$]\label{def:normal velocity}
Let $a$ and $b$ be principal curvatures.
Let $F(a,b)\in C^2\left({\mathbb{R}}^2_{+}\right)$ be a symmetric homogeneous function of degree $\sigma\in{\mathbb{R}}\setminus\{0\}$.
In this paper, we call $F$ a \textit{normal velocity} if
\begin{align*}
F_a,\;F_b > 0
\end{align*}
for all $0<a,b$.
Furthermore, we call $F$ \textit{contracting} if
$F>0$ for all $0<a,b$,
and we call $F$ \textit{expanding} if
$F<0$ for all $0<a,b$.
\end{definition}
\begin{definition}[Quantity $\beta$]\label{def:beta}
Let $F$ be a normal velocity.
We define the quantity $\beta$ as
\begin{align}
\beta = \frac{F_a}{F_b}
\end{align}
for all $0<a,b$.
We later choose $(a,b)=(\rho,1)$ and write $\beta_F(\rho)$.
\end{definition}
\begin{remark}[Normal velocity $F^\sigma_\xi$]\label{rem:normal velocity xi}
In this paper, we investigate the normal velocity $F^\sigma_\xi$.
We define $F^\sigma_\xi$ as
\begin{eqnarray}\label{def:fxi}
F^\sigma_\xi(a,b) =
\begin{cases}
\sgn(\sigma)\cdot\left(a^\xi+b^\xi\right)^{\sigma/\xi}, & \text{if } \xi\neq 0, \\
\sgn(\sigma)\cdot(a\,b)^{\sigma/2}, & \text{if } \xi=0,
\end{cases}
\end{eqnarray}
for all $\sigma\in{\mathbb{R}}\setminus\{0\}$.
Calculating $\big(F^\sigma_\xi\big)_a$, $\big(F^\sigma_\xi\big)_b$, we obtain
\begin{eqnarray}\label{def:fxia}
\left(F^\sigma_\xi\right)_a =
\begin{cases}
|\sigma|\cdot\left(a^\xi+b^\xi\right)^{\sigma/\xi-1}a^{\xi-1}, & \text{if } \xi\neq 0, \\
|\sigma|\cdot(a\,b)^{\sigma/2-1}\frac{b}{2}, & \text{if } \xi=0,
\end{cases}
\end{eqnarray}
and
\begin{eqnarray}\label{def:fxib}
\left(F^\sigma_\xi\right)_b =
\begin{cases}
|\sigma|\cdot\left(a^\xi+b^\xi\right)^{\sigma/\xi-1}b^{\xi-1}, & \text{if } \xi\neq 0, \\
|\sigma|\cdot(a\,b)^{\sigma/2-1}\frac{a}{2}, & \text{if } \xi=0,
\end{cases}
\end{eqnarray}
for all $0<a,b$. Therefore,
\begin{align*}
\left(F^\sigma_\xi\right)_a,\,\left(F^\sigma_\xi\right)_b>0,
\end{align*}
for all $0<a,b$. Hence, $F^\sigma_\xi$ is a normal velocity for all $\xi\in{\mathbb{R}}$.
$F^\sigma_\xi$ is a \textit{contracting} normal velocity, if $\sigma>0$, and
$F^\sigma_\xi$ is an \textit{expanding} normal velocity, if $\sigma<0$.
Calculating the quantity $\beta(\rho)$ we obtain
\begin{align*}
\beta_{F^\sigma_\xi}(\rho) = \rho^{\xi-1},
\end{align*}
for all $\xi\in{\mathbb{R}}$. Note that $\beta_{F^\sigma_\xi}(\rho)$ is independent of $\sigma$.
\end{remark}
\begin{example}[Gauss curvature]
Let $\xi=0$. Then we have $\beta_{F^\sigma_0}(\rho)=\rho^{-1}$, and
\begin{align*}
F^\sigma_0=&\,\sgn(\sigma)\cdot(a\,b)^{\sigma/2} \\
=&\,\sgn(\sigma)\cdot K^{\sigma/2}.
\end{align*}
\end{example}
\begin{example}[Mean curvature]
Let $\xi=1$. Then we have $\beta_{F^\sigma_1}(\rho)=1$, and
\begin{align*}
F^\sigma_1=&\,\sgn(\sigma)\cdot(a+b)^\sigma \\
=&\,\sgn(\sigma)\cdot H^\sigma.
\end{align*}
\end{example}
\begin{example}[Norm of the second fundamental form]
Let $\xi=2$. Then we have $\beta_{F^\sigma_2}(\rho)=\rho$, and
\begin{align*}
F^\sigma_2=&\,\sgn(\sigma)\cdot(a^2+b^2)^{\sigma/2} \\
=&\,\sgn(\sigma)\cdot |A|^\sigma.
\end{align*}
\end{example}
\begin{example}[Trace of the second fundamental form]
Let $\xi=\sigma$. Then we have $\beta_{F^\sigma_\sigma}(\rho)=\rho^{\sigma-1}$, and
\begin{align*}
F^\sigma_\sigma=&\,\sgn(\sigma)\cdot(a^\sigma+b^\sigma)^{\sigma/\sigma} \\
=&\,\sgn(\sigma)\cdot \tr A^\sigma.
\end{align*}
\end{example}
\section{Definition and motivation of rational MPF, and MPF}
In the context of geometric evolution equations \eqref{eq:flow},
\begin{align*}
\frac{d}{dt} X = -F\nu,
\end{align*}
the question arises, if after appropriate rescaling,
closed strictly convex surfaces $M_t$ converge to spheres.
This is also referred to as convergence to a round point
or convergence to a sphere at $\infty$.
The answer is affirmative for many normal velocities $F$,
including the Gauss curvature flow, $F^2_0=K$, B. Andrews \cite{ba:gauss},
the mean curvature flow, $F^1_1=H$, G. Huisken \cite{gh:flow}, and
the inverse Gauss curvature flow, $F^{-1}_0=-\frac{1}{K}$, O. Schn\"urer \cite{os:surfaces}.
For many normal velocities $F$ proofs rely on the fact that a certain geometrically meaningful quantity $w$
is monotone, i.\,e.\ $\max_{M_t} w$ is non-increasing in time.
In \cite{os:surfacesA2} and \cite{os:surfaces}, O. Schn\"urer proposes criteria
for selecting such monotone quantities for contracting flows and for expanding flows, respectively.
To date, to the author's knowledge, all known quantities which fulfill these criteria
can be used to show convergence to a round point or to a sphere at $\infty$.
Here, we decided to work with these criteria as a definition.
Their monotonicity is proven using the maximum-principle.
This is why we name these quantities \textit{rational maximum-principle functions} (RMPF).
\begin{definition}[RMPF]\label{def:rmpf}
Let $a$ and $b$ be principal curvatures.
Let $w(a,b)\in C^2({\mathbb{R}}^2_+)$ be a symmetric homogeneous function of degree $\chi\in{\mathbb{R}}\setminus\{0\}$.
We call $w$ a \textit{rational maximum-principle function} for a normal velocity $F$ if
\begin{enumerate}
\item\label{rI}
\begin{enumerate}
\item $w>0$ for all $0<a,b$, $a\neq b$,
\item $w=0$ for all $0<a=b$.
\end{enumerate}
\item\label{rII}
\begin{enumerate}
\item $\chi>0$ if $F$ is contracting,
\item $\chi<0$ if $F$ is expanding.
\end{enumerate}
\item\label{rIII}
\begin{enumerate}
\item $w_a<0$ for all $0<a<b$,
\item $w_a>0$ for all $0<b<a$.
\end{enumerate}
\item\label{rIV}
Let $Lw:=\frac{d}{dt} w-F^{ij}w_{;\,ij}$ be the linear operator
corresponding to the geometric flow equation \eqref{eq:flow}.
We achieve $Lw\leq 0$ for all $0<a,b$ by assuming that
\begin{enumerate}
\item terms without derivatives of $\left(h_{ij}\right)$ are non-positive for all $0<a,b$, and
\item terms involving derivatives of $\left(h_{ij}\right)$, at a critical point of $w$,
\textit{i.e.} $w_{;i}=0$ for all $i=1,\,2$, are non-positive for all $0<a,b$.
\end{enumerate}
\item\label{rV}
$w(a,b)$ is a rational function.
\end{enumerate}
\end{definition}
The RMPF conditions \eqref{rI} through \eqref{rV} as in \cite{os:surfacesA2}, \cite{os:surfaces} are motivated as follows:
For all geometric flow equations \eqref{eq:flow} we assume that spheres stay spherical.
For contracting (expanding) normal velocities they contract to a point (expand to infinity).
So we can only aim to find monotone quantities, if $w(a,a)=0$ for all $a>0$, or if $\chi\leq 0$ (if $\chi\geq 0$).
If $\chi\leq 0$ (if $\chi\geq 0$), we obtain that $w$ is non-increasing on any self-similarly contracting (expanding) surface.
So this does not imply convergence to a round point.
RMPF condition \eqref{rIII} ensures that the quantity decreases if the principal curvatures approach each other.
By RMPF condition \eqref{rIV}, we check that we can apply the maximum principle to prove monotonicity.
The first found monotone quantities are all rational functions, e.\,g.\ $w=(a-b)^2$ for $F^2_0=K$ in \cite{ba:gauss}, or
$w=\frac{(a-b)^2(a+b)}{(a\,b)}$ for $F^2_2=|A|^2$ in \cite{os:surfacesA2}.
This motivates RMPF condition \eqref{rV}.
\begin{definition}[MPF]\label{def:mpf}
Let $a$ and $b$ be principal curvatures.
Let $w(a,b)\in C^2({\mathbb{R}}^2_+)$ be a symmetric homogeneous function of degree $\chi\in{\mathbb{R}}\setminus\{0\}$.
We call $w$ a \textit{maximum-principle function} for a normal velocity $F$ if
\begin{enumerate}
\item\label{I}
\begin{enumerate}
\item $w>0$ for all $0<a,b$, $a\neq b$,
\item $w=0$ for all $0<a=b$.
\end{enumerate}
\item\label{II}
\begin{enumerate}
\item $\chi>0$ if $F$ is contracting,
\item $\chi<0$ if $F$ is expanding.
\end{enumerate}
\item\label{III}
\begin{enumerate}
\item $w_a<0$ for all $0<a<b$,
\item $w_a>0$ for all $0<b<a$.
\end{enumerate}
\item\label{IV}
We assume that the \textit{constant terms} $C_w(a,b)$ and the \textit{gradient terms} $E_w(a,b)$ and $G_w(a,b)$ are non-positive for all $0<a,b$.
\begin{align*}
C_w(a,b):=&\, w_a\,a \left(\left(F_a\,a^2+F_b\,b^2\right) + \left(F-F_a\,a-F_b\,b\right)a\right) \\
&\quad + w_b\,b \left(\left(F_a\,a^2+F_b\,b^2\right) + \left(F-F_a\,a-F_b\,b\right)b\right), \\
E_w(a,b):=&\, w_a \left(F_{aa} + 2\,F_{ab}\,\alpha + F_{bb}\,\alpha^2\right) \\
&\quad -F_a \left(w_{aa} + 2\,w_{ab}\,\alpha + w_{bb}\,\alpha^2\right) \\
&\quad +2\,\frac{w_b\,F_a-w_a\,F_b}{a-b}\,\alpha^2, \\
G_w(a,b):=&\, \frac{1}{\alpha^2}\cdot\bigg(w_b \left(F_{aa} + 2\,F_{ab}\,\alpha + F_{bb}\,\alpha^2\right)\bigg. \\
&\qquad\qquad -F_b \left(w_{aa} + 2\,w_{ab}\,\alpha + w_{bb}\,\alpha^2\right) \\
&\qquad\qquad \bigg. +2\, \frac{w_b\,F_a - w_a\,F_b}{a-b}\bigg).
\end{align*}
\item\label{V}
We define
\begin{align*}
\alpha = -\frac{w_a}{w_b},\;
\alpha_a = \frac{\partial\alpha}{\partial a}.
\end{align*}
Let $c>0$, $d\in{\mathbb{R}}$ be some constants.
Set $(a,b)=(\rho,1)$, where $0<\rho<1$.
If $F$ is contracting, we assume for $\rho\rightarrow 0$
\begin{align*}
\alpha =&\, c, \qquad \qquad \quad \; \; \, \, \alpha_a = 0, \\
\text{or} \qquad \alpha =&\, \frac{c}{\rho}+o\left(\rho^{-1}\right), \quad
\alpha_a = -\frac{c}{\rho^2}+o\left(\rho^{-1}\right).
\end{align*}
If $F$ is expanding, we assume for $\rho\rightarrow 0$
\begin{align*}
\alpha =&\, \frac{c}{\rho}+o\left(\rho^{-1}\right), \qquad \quad \; \,
\alpha_a = -\frac{c}{\rho^2}+o\left(\rho^{-1}\right), \\
\text{or} \qquad \alpha =&\, \frac{c+d\,\rho}{\rho^2}+o\left(\rho^{-1}\right), \quad
\alpha_a = -\frac{2\,c+d\,\rho}{\rho^2}+o\left(\rho^{-1}\right).
\end{align*}
\end{enumerate}
\end{definition}
F. Schulze and O. Schn\"urer present in \cite{fs:convexity}
one of the first non-rational monotone quantities. They use
\begin{align*}
w=\frac{(a-b)^2(a+b)^{2\sigma}}{(a\,b)^2}
\end{align*}
to show convergence to a round point for $F^\sigma_1=H^\sigma$, for all $1<\sigma\leq 5$.
Here, $w$ is not rational, if $2\sigma\notin{\mathbb{N}}$.
Therefore, in MPF Definition \ref{def:mpf},
we extend RMPF Definition \ref{def:rmpf} to not necessarily rational functions.
We call them \textit{maximum-principle functions} (MPF).
In Lemma \ref{lem:rmpf}, we prove that MPF condition \eqref{V} holds for any RMPF
in the case of contracting flows $F^\sigma_\xi$, if $\xi>0$, or $\xi=0$ and $1<\sigma\leq 2$.
For other $F^\sigma_\xi$,
a similar Lemma is in progress.
A first step in the this direction is Lemma \ref{lem:alpha o estimates}.
To date, all known RMPF fulfill MPF condition \eqref{V}.
In this paper, the main question is, when MPF cease to exist.
As it turns out,
this is more an algebra than a differential geometry question.
Correspondingly, in MPF condition \eqref{IV}
we formulate three inequality conditions.
In Lemmas \ref{lem:1stDerivatives} through \ref{lem:LwCriticalPoint}, we show that these three inequality conditions
are equivalent to RMPF condition \eqref{rIV}.
In condensed form, MPF Definition \ref{def:mpf} is a more algebraic formulation of RMPF Definition \ref{def:rmpf},
which is extended to what seems to be the proper class of non-rational functions.
\subsection{Linear operator L}\label{sc:terms}
In Lemmas \ref{lem:1stDerivatives} through \ref{lem:LwCriticalPoint}, we show that the three inequality conditions
$C_w(a,b),\,E_w(a,b),\,G_w(a,b)\leq 0$
in condition MPF \eqref{IV} are equivalent
to RMPF condition \eqref{rIV} as follows:
In Lemma \ref{lem:1stDerivatives}, we show two helpful identities for the first derivatives
with respect to the induced metric $g_{ij}$, and for the derivatives with respect to the second fundamental form $h_{ij}$.
Next, we cite the evolution equations for $g_{ij}$ and $h_{ij}$ from O. Schn\"urer \cite{os:alpbach}.
We need both Lemmas in Lemma \ref{lem:Lw}.
Here, we compute the linear operator $Lw$, where $w$ is a function of the principal curvatures $a$ and $b$.
Sometimes, we also denote them by $\lambda_1$ and $\lambda_2$.
In Lemma \ref{lem:2ndDerivatives}, we cite another helpful identity for the second derivatives
with respect to the second fundamental form.
Finally, in Lemma \ref{lem:LwCriticalPoint} we compute the constant terms $C_w(a,b)$, and the two gradient terms $E_w(a,b)$ and $G_w(a,b)$.
This concludes the proof of our claim that condition \eqref{IV} of RMPF Definition \ref{def:rmpf} and of MPF Definition \ref{def:mpf}
are equivalent.
\begin{lemma}[First derivatives]\label{lem:1stDerivatives}
Let $f$ be a normal velocity $F$ or a function $w$ of the principal curvatures $\lambda_1$ and $\lambda_2$.
Then we have
\begin{align}
f^{kl} =&\, f^k_j g^{lj}, \label{eq:fhkl} \\
\frac{\partial f}{\partial g_{kl}} =&\, -f^{il}h^k_i. \label{eq:fgkl}
\end{align}
The matrices $\left(f^{kl}\right)$ and $\left(\frac{\partial f}{\partial g_{kl}}\right)$ are symmetric.
\end{lemma}
\begin{proof}
We consider $f = f\left(h^j_i\big((h_{kl}),(g_{kl})\big)\right)$.
The matrices $\left(h_{ij}\right)$ and $\left(g_{ij}\right)$ are symmetric.
So we have $h_{ij} = \frac{1}{2}\left(h_{ij} + h_{ji}\right)$ and $g^{ij} = \frac{1}{2}\left(g^{ij} + g^{ji}\right)$. \\
Now, we differentiate $f$ with respect to $h_{kl}$.
\begin{align*}
f^{kl} =&\, \frac{\partial f}{\partial h_{kl}} {\displaybreak[1]} \\
=&\,\frac{\partial f}{\partial h^j_i} \frac{\partial h^j_i}{h_{kl}} {\displaybreak[1]} \\
=&\, f^i_j\, \frac{\partial}{\partial h_{kl}} \frac{1}{2}\left(h_{im}+h_{mi}\right)\frac{1}{2}\left(g^{mj}+g^{jm}\right) {\displaybreak[1]} \\
=&\, f^j_i\, \frac{1}{4} \frac{\partial}{\partial h_{kl}} \left(h_{im} g^{mj} + h_{im} g^{jm}+h_{mi} g^{mj}+h_{mi} g^{jm}\right) {\displaybreak[1]} \\
=&\, f^j_i\, \frac{1}{4} \left(\delta^k_i\delta^l_m g^{mj} + \delta^k_i\delta^l_m g^{jm} + \delta^k_m\delta^l_i g^{mj} + \delta^k_m \delta^l_i g^{jm}\right) {\displaybreak[1]} \\
=&\, f^i_j\, \frac{1}{4} \left(2\,\delta^k_i g^{jl} + 2\,\delta^l_i g^{jk}\right) {\displaybreak[1]} \\
=&\, \frac{1}{2} \left(f^k_j g^{jl} + f^l_j g^{jk}\right).
\end{align*}
Since $f^{kl} - f^{lk} = 0$ for all $1 \leq k,\, l \leq n$, the matrix $\left(f^{kl}\right)$ is symmetric and formula \eqref{eq:fhkl} follows. \\
Next, we differentiate $f$ with respect to $g_{kl}$.
\begin{align*}
\frac{\partial f}{\partial g_{kl}}
=&\, \frac{\partial f}{\partial h^j_i} \frac{\partial}{\partial g_{kl}} \frac{1}{2}
\left(h_{im} + h_{mi}\right) \frac{1}{2} \left(g^{mj} + g^{jm}\right) {\displaybreak[1]}\\
=&\, f^i_j\,\frac{\partial}{\partial g_{kl}}\,\frac{1}{4} \left(h_{im} g^{mj} + h_{im} g^{jm} + h_{mi} g^{mj} + h_{mi} g^{jm}\right) {\displaybreak[1]}\\
=&\, -f^i_j\,\frac{1}{4}\,\left(h_{im} g^{mk} g^{jl} + h_{im} g^{jk} g^{ml} + h_{mi} g^{mk} g^{jl} + h_{mi} g^{jk} g^{ml}\right) \\
&\, \left(\textit{use } \frac{\partial g^{ij}}{\partial g_{kl}} = -g^{ik} g^{jl}\right) {\displaybreak[1]}\\
=&\, -f^i_j\,\frac{1}{2}\,\left(h^k_i g^{jl} + h^l_i g^{jk}\right)
\end{align*}
Since $\frac{\partial f}{\partial g_{kl}} - \frac{\partial f}{\partial g_{lk}} = 0$ for all $1 \leq k,\,l \leq n$ the matrix $\left(\frac{\partial f}{\partial g_{kl}}\right)$ is symmetric and formula \eqref{eq:fgkl} follows.
\end{proof}
\begin{lemma}[Evolution equations]\label{lem:evolution}
Let $X$ be a solution to a geometric flow equation \eqref{eq:flow}.
Then we have the following evolution equations:
\begin{align}
\frac{d}{dt}g^{ij} =&\, 2 F h^{ij}, \label{eq:gij} \\
Lh_{ij} =&\,F^{kl} h^m_k h_{lm} \cdot h_{ij} - \left(F - F^{kl} h_{kl}\right) h^m_i h_{jm} + F^{kl,rs} h_{kl;i} h_{rs;j}, \label{eq:hij}
\end{align}
where $L$ is the linear operator of RMPF Definition \ref{def:rmpf}.
\end{lemma}
\begin{proof}
We refer to O. Schn\"urer \cite{os:alpbach}.
\end{proof}
\begin{lemma}[Linear operator]\label{lem:Lw}
Let $w=w\big(h^j_i\big)$ be a function of the principal curvatures.
Then we have
\begin{equation}
\label{eq:whji}
\begin{split}
Lw=&\, w^{ij} \left(h_{ij} F^{kl} h^m_k h_{lm} + h^m_i h_{jm} \left(F-F^{kl}h_{kl}\right) \right) \\
&\quad + \left(w^{ij} F^{kl,rs} - F^{ij} w^{kl,rs} \right) h_{kl;i} h_{rs;j},
\end{split}
\end{equation}
where $L$ is the linear operator of RMPF Definition \ref{def:rmpf}.
\end{lemma}
\begin{proof}
We consider $w = w\left(h^j_i\big((h_{kl}),(g_{kl})\big)\right)$.
\begin{align*}
Lw =&\, \frac{d}{dt} w - F^{rs} w_{;rs} {\displaybreak[1]}\\
=&\, \frac{\partial w}{\partial h^j_i} \frac{d}{dt} \left(h_{im} g^{mj}\right) - F^{rs}\bigg(\frac{\partial w}{\partial h^j_i} \left(h_{im} g^{mj}\right)_{;r}\bigg)_{;s} {\displaybreak[1]}\\
=&\, w^i_j \left(\dot{h}_{im} g^{mj} + h_{im} \dot{g}^{mj}\right) - F^{rs} \left(w^i_j\left(h_{im;r} g^{mj} \right)\right)_{;s} {\displaybreak[1]}\\
=&\, 2\,F\,w^i_j h^{mj}h_{im} + w^i_j g^{mj}\dot{h}_{im} - F^{rs} \left(w^i_j g^{mj} h_{im;r}\right)_{;s} {\displaybreak[1]}\\
&\, \left(\textit{use evolution equation \eqref{eq:gij}}\right) {\displaybreak[1]}\\
=&\, 2\,F\,w^{ij} h^m_j h_{im} + w^{im} \dot{h}_{im} - F^{rs} \left(w^{im} h_{im;r}\right)_{;s} {\displaybreak[1]}\\
&\, \left(\textit{use formula \eqref{eq:fhkl}: } w^{kl} = w^k_j g^{lj} \right) {\displaybreak[1]}\\
=&\, 2\,F\,w^{ij} h^m_j h_{im} + w^{im} \dot{h}_{im} - F^{rs} \frac{\partial w^{im}}{\partial h^l_k} \left(h_{kn} g^{nl}\right)_{;s} h_{im;r}
-F^{rs} w^{im} h_{im;rs} {\displaybreak[1]}\\
=&\, 2\,F\,w^{ij} h^m_j h_{im} + w^{im} \dot{h}_{im} - F^{rs} \left(w^{im}\right)^k_l g^{nl} h_{kn;s} h_{im;r} - F^{rs} w^{im} h_{im;rs} {\displaybreak[1]}\\
&\, \left(\textit{use formula \eqref{eq:fhkl}: } w^{kl} = w^k_j g^{lj} \right) {\displaybreak[1]}\\
=&\, 2\,F\,w^{ij} h^m_j h_{im} + w^{im}\left(\dot{h}_{im} - F^{rs} h_{im;rs}\right) - F^{rs} w^{im,kn} h_{im;r} h_{kn;s} {\displaybreak[1]}\\
=&\, 2\,F\,w^{ij} h^m_i h_{jm} + w^{ij}\left(\dot{h}_{ij} - F^{kl} h_{ij,kl}\right) - F^{ij} w^{kl,rs} h_{kl;i} h_{rs;j} {\displaybreak[1]}\\
&\, \left(\textit{rename indices}\right) {\displaybreak[1]}\\
=&\, w^{ij}\left(F^{kl} h^m_k h_{lm}\cdot h_{ij} + \left(F-F^{kl}h_{kl}\right)h^m_i h_{jm} + F^{kl,rs} h_{kl;i} h_{rs;j}\right) \\
&\quad - F^{ij} w^{kl,rs} h_{kl;i} h_{rs;j} {\displaybreak[1]}\\
&\, \left(\textit{use evolution equation \eqref{eq:hij}}\right) {\displaybreak[1]}\\
=&\, w^{ij} \left(h_{ij} F^{kl} h^m_k h_{lm} + h^m_i h_{jm} \left(F - F^{kl} h_{kl}\right)\right) \\
&\quad +\left(w^{ij} F^{kl,rs} - F^{ij} w^{kl,rs}\right) h_{kl;i} h_{rs;j}.\qedhere
\end{align*}
\end{proof}
\begin{lemma}[Second derivatives]\label{lem:2ndDerivatives}
Let $f$ be a normal velocity $F$ or a function $w$ of the principal curvatures $\lambda_1$ and $\lambda_2$.
Then we have
\begin{align}\label{eq:fijkl}
f^{ij,kl} \eta_{ij} \eta_{kl}
= \sum_{i,\,j} \frac{\partial^2 f}{\partial \lambda_i \partial \lambda_j} \eta_{ii} \eta_{jj}
+ \sum_{i\neq j} \frac{\frac{\partial f}{\partial \lambda_i}-\frac{\partial f}{\partial \lambda_j}}{\lambda_i - \lambda_j} \eta_{ij}^2
\end{align}
for any symmetric matrix $\left(\eta_{ij}\right)$ and $\lambda_1 \neq \lambda_2$, or $\lambda_1 = \lambda_2$ and the last term is interpreted as a limit.
\end{lemma}
\begin{proof}
We refer to C. Gerhardt \cite{cg:curvature}.
\end{proof}
\begin{lemma}[Linear operator at a critical point]\label{lem:LwCriticalPoint}
Let $w=w(h^j_i)$ be a symmetric function of the principal curvatures $a$ and $b$.
At a critical point of $w$, i.\,e.\ $w_{;i}=0$ for all $i=1,2$,
we choose normal coordinates,
i.\,e.\ $g_{ij} = \delta_{ij}$ and $\left(h_{ij}\right) = \diag\left(a,\,b\right)$. \\
Then we have
\begin{align*}
Lw = C_w\left(a,\,b\right) + E_w\left(a,\,b\right) h_{11;1}^2 + G_w\left(a,\,b\right) h_{22;2}^2,
\end{align*}
where $L$ is the linear operator of RMPF Definition \ref{def:rmpf}, and
$\alpha$ is the quantity of MPF Definition \ref{def:mpf}. \\
The constant terms are
\begin{align*}
C_w\left(a,\,b\right) =&\, w_a\,a \left(\left(F_a\,a^2+F_b\,b^2\right) + \left(F-F_a\,a-F_b\,b\right)a\right) \\
&\quad + w_b\,b \left(\left(F_a\,a^2+F_b\,b^2\right) + \left(F-F_a\,a-F_b\,b\right)b\right) \\
=&\,F\left(w_a\,a^2+w_b\,b^2\right) \\
&\quad + F_a\,w_b\,a\,b\left(a-b\right) \\
&\quad -F_b\,w_a\,a\,b\left(a-b\right),
\end{align*}
and the two gradient terms are
\begin{align*}
E_w(a,b):=&\, w_a \left(F_{aa} + 2\,F_{ab}\,\alpha + F_{bb}\,\alpha^2\right) \\
&\quad -F_a \left(w_{aa} + 2\,w_{ab}\,\alpha + w_{bb}\,\alpha^2\right) \\
&\quad +2\,\frac{w_b\,F_a-w_a\,F_b}{a-b}\,\alpha^2, {\displaybreak[1]} \\
G_w(a,b):=&\, \frac{1}{\alpha^2}\cdot\bigg(w_b \left(F_{aa} + 2\,F_{ab}\,\alpha + F_{bb}\,\alpha^2\right)\bigg. \\
&\qquad\qquad -F_b \left(w_{aa} + 2\,w_{ab}\,\alpha + w_{bb}\,\alpha^2\right) \\
&\qquad\qquad \bigg. +2\, \frac{w_b\,F_a - w_a\,F_b}{a-b}\bigg).
\end{align*}
We obtain $G_w\left(a,b\right) = E_w\left(b,a\right)$ for all $0<a,b$.
\end{lemma}
\begin{proof}
We consider $w = w\big(h^j_i\big)$.
At a critical point of $w$, we have
\begin{align*}
w_{;k} = \frac{\partial w}{\partial h^j_i} \big(h^j_i\big)_{;k}
= w^i_j \left(h_{il} g^{lj}\right)_{;k}
= w^i_j g^{lj} h_{il;k}
= w^{ij} h_{ij;k}
= 0.
\end{align*}
Here, we also choose normal coordinates and get
$w_{;1} = w_a\,h_{11;1} + w_b\,h_{22;1} = 0$, implying
\begin{align}
h_{22;1} = -\frac{w_a}{w_b}\,h_{11;1} = \alpha\cdot h_{11;1}, \label{eq:h111}
\end{align}
and $w_{;2} = w_a\,h_{11;2} + w_b\,h_{22;2} = 0$, implying
\begin{align}
h_{11;2} = -\frac{w_b}{w_a}\,h_{22;2} = \frac{1}{\alpha}\cdot h_{22;2}. \label{eq:h222}
\end{align}
Now, we compute the linear operator $Lw$ of Lemma \ref{lem:Lw}
at a critical point of $w$, where we choose normal coordinates.
\begin{align*}
Lw =&\, w^{ij} \left(h_{ij}\,F^{kl} h^m_k h_{lm} + h^m_i h_{jm} \left(F - F^{kl} h_{kl}\right)\right) \\
&\quad +\left(w^{ij}\,F^{kl,rs} - F^{ij}\,w^{kl,rs}\right) h_{kl;i} h_{rs;j} \\
=&\, w_a\left(\left(F_a\,a^2+F_b\,b^2\right)a+\left(F-F_a\,a-F_b\,b\right)a^2\right) \\
&\quad +w_b\left(\left(F_a\,a^2+F_b\,b^2\right)b+\left(F-F_a\,a-F_b\,b\right)b^2\right) \\
&\quad +w_a\left(F_{aa}\,h^2_{11;1} + 2\,F_{ab}\,h_{11;1} h_{22;1} + F_{bb}\,h^2_{22;1} + 2\,\frac{F_a - F_b}{a-b}\,h^2_{12;1}\right) \\
&\quad -F_a\left(w_{aa}\,h^2_{11;1} + 2\,w_{ab}\,h_{11;1} h_{22;1} + w_{bb}\,h^2_{22;1} + 2\,\frac{w_a - w_b}{a-b}\,h^2_{12;1}\right) \\
&\quad +w_b\left(F_{aa}\,h^2_{11;2} + 2\,F_{ab}\,h_{11;2} h_{22;2} + F_{bb}\,h^2_{22;2} + 2\,\frac{F_a - F_b}{a-b}\,h^2_{12;2}\right) \\
&\quad -F_b\left(w_{aa}\,h^2_{11;2} + 2\,w_{ab}\,h_{11;2} h_{22;2} + w_{bb}\,h^2_{22;2} + 2\,\frac{w_a - w_b}{a-b}\,h^2_{12;2}\right) \\
&\qquad \left(\textit{use formula \eqref{eq:fijkl}}\right).
\end{align*}
Clearly, the constant terms $C_w\left(a,b\right)$ are the first two lines of terms after the last equation equal sign.
It remains to compute the gradient terms using identities \eqref{eq:h111} and \eqref{eq:h222}.
We obtain the first gradient terms
\begin{align*}
E_w\left(a,b\right)\cdot h^2_{11;1} =&\,w_a\left(F_{aa}\,h^2_{11;1} + 2\,F_{ab}\,\alpha\,h^2_{11;1} + F_{bb}\,\alpha^2\,h^2_{11;1}\right) \\
&\quad -F_a\left(w_{aa}\,h^2_{11;1} + 2\,w_{ab}\,\alpha\,h^2_{11;1} + w_{bb}\,\alpha^2\,h^2_{11;1}\right) \\
&\quad +2\,\frac{1}{a-b}\,\left(w_b\,F_a-w_b\,F_b-w_a\,F_b+w_b\,F_b\right)\alpha^2\,h^2_{11;1} \\
=&\,\bigg(w_a\left(F_{aa} + 2\,F_{ab}\,\alpha + F_{bb}\,\alpha^2\right) \bigg. \\
&\quad -F_a\left(w_{aa} + 2\,w_{ab}\,\alpha + w_{bb}\,\alpha^2\right) \\
&\quad \left.\,+2\,\frac{w_b\,F_a - w_a\,F_b}{a-b}\,\alpha^2\right)\,h^2_{11;1},
\end{align*}
and we obtain second gradient terms
\begin{align*}
G_w\left(a,b\right)\cdot h^2_{22;2}
=&\,w_b\left(F_{aa}\,\frac{1}{\alpha^2}\,h^2_{22;2} + 2\,F_{ab}\,\frac{1}{\alpha}\,h^2_{22;2} + F_{bb}\,h^2_{22;2} \right) \\
&\quad -F_b\left(w_{aa}\,\frac{1}{\alpha^2}\,h^2_{22;2} + 2\,w_{ab}\,\frac{1}{\alpha}\,h^2_{22;2} + w_{bb}\,h^2_{22;2} \right) \\
&\quad +2\,\frac{1}{a-b}\left(w_a\,F_a - w_a\,F_b - w_a\,F_a + w_b\,F_a\right)\frac{1}{\alpha^2}\,h^2_{22;2} \\
=&\, \frac{1}{\alpha^2}\cdot\bigg(w_b \left(F_{aa} + 2\,F_{ab}\,\alpha + F_{bb}\,\alpha^2\right)\bigg. \\
&\qquad\qquad -F_b \left(w_{aa} + 2\,w_{ab}\,\alpha + w_{bb}\,\alpha^2\right) \\
&\qquad\qquad \bigg. +2\, \frac{w_b\,F_a - w_a\,F_b}{a-b}\bigg)h_{22;2}^2. \qedhere
\end{align*}
\end{proof}
\subsection{$\alpha$-conditions}\label{sc:rmpf}
RMPF Definition \ref{def:rmpf} and MPF Definition \ref{def:mpf} are the same up to condition \eqref{III},
and in Lemma \ref{lem:LwCriticalPoint} we have shown that RMPF and MPF conditions \eqref{IV} are equivalent.
So it only remains to motivate MPF condition \eqref{V}.
RMPF condition \eqref{rV} assumes any such function to be rational.
As mentioned before, the example of
\begin{align*}
w = \frac{(a-b)^2(a+b)^{2\sigma}}{(a\,b)^2},
\end{align*}
given by O. Schn\"urer and F. Schulze \cite{fs:convexity} fulfills all RMPF conditions
except RMPF condition \eqref{rV} for any $2\sigma\notin {\mathbb{N}}$.
But it can be used to show convergence to a round point
using the maximum-principle.
As a first step, we analyzed all functions that were used
to show convergence to a round point and convergence to a sphere at $\infty$ in
\cite{ba:gauss}, \cite{ac:surfaces}, \cite{ql:surfaces},
\cite{os:surfacesA2}, \cite{os:surfaces}, and \cite{fs:convexity}.
All of them have in common that they fulfill MPF condition \eqref{V}.
As a second step, we seek to prove that all RMPF are MPF.
This is Lemma \ref{lem:rmpf}.
Here, we show that for contracting $F^\sigma_\xi$
with $\xi>0$, or $\xi=0$ and $1<\sigma\leq 2$,
RMPF fulfill MPF condition \eqref{V}.
This includes the contracting normal velocities
$F^\sigma_0=K^{\sigma/2}$, $F^\sigma_1=H^\sigma$, and $F^\sigma_2=|A|^\sigma$.
A similar Lemma for the remaining contracting and the expanding $F^\sigma_\xi$
is in progress.
In \cite{mf:on}, we show that there are no RMPF for any $F=K^{\sigma/2}$
with $\sigma>2$.
And due to B. Andrews we know that convex surfaces do not necessarily
converge to a round point for any power $\frac{1}{2} \leq \sigma \leq 1$.
For the power $\sigma=\frac{1}{2}$, they converge to ellipsoids \cite{ba:contraction}.
For any power $\frac{1}{2}\leq \sigma \leq 1$, surfaces
contract homothetically in the limit \cite{ba:motion}.
This is why we cannot expect any RMPF to exist for any $F=K^{\sigma/2}$
with $0<\sigma\leq 1$. \\
We begin this chapter by showing that the quantity $\alpha=-\frac{w_a}{w_b}$
is strictly positive for
all $0<a,b$, if $w$ is a RMPF or a MPF.
In Remark \ref{rem:examples condition}, we give an example for each of the $\alpha$-conditions in MPF condition \eqref{V}.
Then we proceed with Lemma \ref{lem:alpha o estimates},
which contains some calculations for Lemma \ref{lem:rmpf}.
By this Lemma, we motivate MPF condition \eqref{V}.
\begin{lemma}[Quantity $\alpha$ is strictly positive]\label{lem:alpha}
Let $w$ be a RMPF or a MPF.
Then we have
\begin{align*}
\alpha>0
\end{align*}
for all $0<a,\,b$.
Note that $\alpha=-\frac{w_a}{w_b}$ as in MPF Definition \ref{def:mpf}.
\end{lemma}
\begin{proof}
Let $0<a,\,b$.
MPF condition \eqref{III} and the symmetry of $w$ in $a$ and $b$ imply
\begin{align*}
w_a(a,b) <&\, 0 \text{ for all } 0 < a < b, \\
w_a(a,b) >&\, 0 \text{ for all } 0 < b < a, \\
w_b(a,b) >&\, 0 \text{ for all } 0 < a < b, \\
w_b(a,b) <&\, 0 \text{ for all } 0 < b < a.
\end{align*}
Recall, $\alpha=-\frac{w_a}{w_b}$.
Hence, the claim follows for all $0<a,b$ with $a\neq b$.
Now, let $0<a,b$ with $a=b$.
By MPF condition \eqref{IV}, in particular $C_w\leq 0$ for all $0<a,b$, we have
\begin{align*}
&\, w_a\,\underbrace{a\left(\left(F_a\,a^2+F_b\,b^2\right) + \left(F-F_a\,a-F_b\,b\right)a\right)}_{:={\varphi}} \\
&\quad+w_b\,\underbrace{b\left(\left(F_a\,a^2+F_b\,b^2\right) + \left(F-F_a\,a-F_b\,b\right)b\right)}_{:=\psi} \\
\leq&\, 0
\end{align*}
for all $0<a,b$. This implies
\begin{align*}
w_a\,{\varphi}\leq-w_b\,\psi
\end{align*}
for all $0<a,b$. \\
Letting $0<a<b$ yields $\alpha\cdot{\varphi} \geq \psi$, and
letting $0<b<a$ yields $\alpha\cdot{\varphi} \leq \psi$.
Hence, for $a\rightarrow b$ we get
\begin{align*}
\alpha\cdot {\varphi} \leq \psi \leq \alpha\cdot {\varphi}.
\end{align*}
This implies $\psi=\alpha\cdot{\varphi}$ for all $0<a,b$ with $a=b$.
We get
\begin{align*}
{\varphi} =&\, \alpha\cdot{\varphi}, \text{ since } \psi={\varphi} \text{ at } a=b, \\
\text{and } \alpha =&\, 1, \text{ if } {\varphi}\neq 0 \text{ at } a=b.
\end{align*}
Let $a=b=1$, then
\begin{align*}
{\varphi}(1,1)
=&\, 1\cdot ((F_a\cdot 1+F_b\cdot 1)+(F-F_a\cdot 1-F_b\cdot 1)\cdot 1) \\
=&\, F \neq 0 \text{ by Definition \ref{def:normal velocity} of a normal velocity } F.
\end{align*}
Since ${\varphi}$ is homogeneous, we have ${\varphi}\neq 0$ for all $0<a,b$ with $a=b$.
This concludes the proof.
\end{proof}
\begin{remark}[$\alpha$-conditions]\label{rem:examples condition}
We give an example for each of the $\alpha$-conditions in MPF condition \eqref{V}.
Let $F=F^\sigma_0=\sgn(\sigma)\cdot K^{\sigma/2}$. Then we have
\begin{center}
\begin{tabular}{| l | c || c | c | c |}
\hline
& MPF & & & $\alpha$-condition \\
\hline \hline
& & & & \\
$\Bigg.\Bigg.$ $\sigma=2$ & $(a-b)^2$ & B. Andrews & \cite{ba:gauss} & $\alpha = 1$ \\
\hline
& & & & \\
$\Bigg.\Bigg.$ $\sigma\in(1,2)$ & $\frac{(a-b)^2(a\,b)^\sigma}{(a\,b)^2}$ & B. Andrews, X. Chen & \cite{ac:surfaces} & $\alpha = \frac{-\sigma+2}{\sigma\rho} + o\left(\rho^{-1} \right)$ \\
\hline \hline
& & & & \\
$\Bigg.\Bigg.$ $\sigma\in(-2,0)$ & $\frac{(a-b)^2(a\,b)^{\sigma/2}}{(a\,b)}$ & Q. Li & \cite{ql:surfaces} & $\alpha = \frac{-\sigma+2}{(\sigma+2)\rho} + o\left(\rho^{-1} \right)$ \\
\hline
& & & & \\
$\Bigg.\Bigg.$ $\sigma=-2$ & $\frac{(a-b)^2}{(a\,b)^2}$ & O. Schn\"urer & \cite{os:surfaces} & $\alpha = \frac{1}{\rho^2}$\\
\hline
\end{tabular}
\end{center}
\end{remark}
\begin{lemma}[$o$-estimates for the quantity $\alpha$]\label{lem:alpha o estimates}
Let $F^\sigma_\xi$ be defined as in Remark \ref{rem:normal velocity xi}.
Assume $w$ to be a RMPF or a MPF for a $F^\sigma_\xi$.
Let $F^\sigma_\xi$ be a contracting normal velocity, i.\,e.\ $\sigma>0$.
Then for $\rho\rightarrow 0$ we get
\begin{align}
\Big. \Big. \alpha \geq&\, \frac{1}{\sigma\,\rho} + o\left(\rho^{-1}\right), \text{ if } \xi>0, {\displaybreak[1]} \\
\Big. \Big. \alpha \geq&\, \frac{-\sigma+2}{\rho} + o\left(\rho^{-1}\right), \text{ if } \xi=0 \text{ and } \sigma\neq 2, {\displaybreak[1]} \\
\Big. \Big. \alpha \geq&\, 1, \text{ if } \xi=0 \text{ and } \sigma=2, {\displaybreak[1]} \\
\Big. \Big. \alpha \geq&\, \frac{-\sigma+1}{\sigma\,\rho^{-\xi}} + \sigma\left(\rho^{-1}\right), \text{ if } -1<\xi<0, {\displaybreak[1]} \\
\Big. \Big. \alpha \geq&\, \frac{-\sigma+1}{(\sigma+1)\rho^2} + \frac{1}{\rho} + o\left(\rho^{-1}\right), \text{ if } \xi=-1, {\displaybreak[1]} \\
\Big. \Big. \alpha \geq&\, \frac{-\sigma+1}{\rho^2} + \frac{\sigma}{\rho} + o\left(\rho^{-1}\right), \text{ if } \xi<-1.
\end{align}
Let $F^\sigma_\xi$ be an expanding normal velocity, i.\,e.\ $\sigma<0$.
Then for $\rho\rightarrow 0$ we get
\begin{align}
\Big. \Big. \alpha \geq&\, \frac{1}{\sigma\,\rho} + o\left(\rho^{-1}\right), \text{ if } \xi>0, {\displaybreak[1]} \\
\Big. \Big. \alpha \geq&\, \frac{-\sigma+2}{\rho} + o\left(\rho^{-1}\right), \text{ if } \xi=0, {\displaybreak[1]} \\
\Big. \Big. \alpha \geq&\, \frac{-\sigma+1}{\sigma\,\rho^{-\xi}} + o\left(\rho^{-1}\right), \text{ if } -1<\xi<0, {\displaybreak[1]} \\
\Big. \Big. \alpha \geq&\, \frac{-\sigma+1}{(\sigma+1)\rho^2} + \frac{1}{\rho} + o\left(\rho^{-1}\right), \text{ if } \xi=-1 \text{ and } \sigma<-1, {\displaybreak[1]} \\
\Big. \Big. \alpha \leq&\, \frac{1}{\rho^3}, \text{ if } \xi=-1 \text{ and } \sigma=-1, {\displaybreak[1]} \\
\Big. \Big. \alpha \leq&\, \frac{-\sigma+1}{(\sigma+1)\rho^2} + \frac{1}{\rho} + o\left(\rho^{-1}\right), \text{ if } \xi=-1 \text{ and } -1<\sigma<0, {\displaybreak[1]} \\
\Big. \Big. \alpha \leq&\, \frac{-\sigma+1}{\rho^2} + \frac{\sigma}{\rho} + o\left(\rho^{-1}\right), \text{ if } \xi<-1.
\end{align}
\end{lemma}
\begin{proof}
Assume $w$ to be a RMPF or a MPF for a $F^\sigma_\xi$.
By Lemma \ref{lem:necessary}, we have
\begin{align*}
0 \geq&\, C^\alpha_\beta(\rho) \\
=&\, \sgn(\sigma)\cdot\left( -\alpha\,\rho\left(\sigma+\rho(-\sigma+1)+\beta\,\rho^2 \right)
+1 +\beta\,\rho(1-\sigma+\rho\,\rho) \right),
\end{align*}
which implies
\begin{align}\label{id:alpha sign}
\begin{split}
&\, \sgn(\sigma)\cdot\Big( \alpha\,\rho \big(\overbrace{ \sigma+\rho(-\sigma+1) + \beta\,\rho^2 }^{=:A}\big)\Big) \\
\geq&\, \sgn(\sigma)\cdot \Big( \underbrace{1+\beta\,\rho\big(1-\sigma+\rho\,\sigma\big)}_{=:B}\Big).
\end{split}
\end{align}
We seek estimates for $\alpha$ in some zero neighborhood of $\rho$.
So we need to know the sign of $A$ in such a neighborhood.
For our convenience, we calculate the quantity $\alpha$ for
\begin{align*}
\beta = \rho^\xi.
\end{align*}
Towards the end of the proof we just need to add one to $\xi$
to obtain the proper results for $F^\sigma_\xi$.
This yields
\begin{align*}
A = \sigma + \rho(-\sigma+1) + \rho^{\xi+2}.
\end{align*}
We get
\begin{align*}
{\lim_{\rho\rightarrow 0}\,} A =&\, \sigma, \text{ if } \xi>-2, \\
{\lim_{\rho\rightarrow 0}\,} A =&\, \sigma+1, \text{ if } \xi=-2 \text{ and } \sigma\neq -1, \\
{\lim_{\rho\rightarrow 0}\,} \rho^{-1}\cdot A =&\, 2, \text{ if } \xi=-2 \text{ and } \sigma=-1, \\
{\lim_{\rho\rightarrow 0}\,} \rho^{-\xi-2}\cdot A =&\, 1, \text{ if } \xi<-2.
\end{align*}
Therefore, $A$ is strictly positive in some zero neighborhood of $\rho$, if $F^\sigma_\xi$
is a contracting normal velocity, i.\,e.\ $\sigma>0$. \\
If $F^\sigma_\xi$ is an expanding normal velocity, i.\,e.\ $\sigma<0$, we obtain
a zero neighborhood of $\rho$, where
\begin{itemize}
\item $A$ is strictly negative, if $\xi>-2$,
\item $A$ is strictly negative, if $\xi=-2$ and $\sigma<-1$,
\item $A$ is strictly positive, if $\xi=-2$ and $\sigma=-1$,
\item $A$ is strictly positive, if $\xi=-2$ and $-1<\sigma<0$,
\item $A$ is strictly positive, if $\xi<-2$.
\end{itemize}
Note that by \eqref{id:alpha sign}, we have
\begin{align*}
\alpha\,\rho\,A \geq B, \text{ if } \sigma>0, \\
\alpha\,\rho\,A \leq B, \text{ if } \sigma<0,
\end{align*}
for all $0<\rho<1$. Now, we use the results on $A$.
Let $F^\sigma_\xi$ be a contracting normal velocity.
Then we have
\begin{align}\label{id:alpha estimate}
\alpha \geq \frac{B}{\rho\,A}
\end{align}
in some zero neighborhood of $\rho$.
Let $F^\sigma_\xi$ be an expanding normal velocity.
Then we have
\begin{align}\label{id:alpha estimate I}
\alpha \geq \frac{B}{\rho\,A}
\end{align}
in some zero neighborhood of $\rho$,
if $\sigma>-2$, or $\xi=-2$ and $\sigma<-1$, \\
and we have
\begin{align}\label{id:alpha estimate II}
\alpha \leq \frac{B}{\rho\,A}
\end{align}
in some zero neighborhood of $\rho$,
if $\xi=-2$ and $-1\leq \sigma<0$, or $\xi<-2$. \\
In the following, we can calculate the desired $o$-estimates for the quantity $\alpha$. \\
First, let $F^\sigma_\xi$ be a contracting normal velocity. \\
By \eqref{id:alpha estimate}, we have
\begin{align*}
\alpha \geq \frac{1+\rho^{\xi+1}(1-\sigma+\rho\,\sigma)}{\rho\left( \sigma + \rho(-\sigma+1) + \rho^{\xi+2} \right)}.
\end{align*}
Let $\xi>-1$. For $\rho\rightarrow 0$ we get
\begin{align*}
\alpha \geq \frac{1}{\sigma\,\rho} + o\left(\rho^{-1}\right).
\end{align*}
Let $\xi=-1$. For $\rho\rightarrow 0$ we get
\begin{align*}
\alpha \geq&\, \frac{-\sigma+2}{\rho} + o\left(\rho^{-1}\right), \text{ if } \sigma\neq 2, \\
\alpha \geq&\, 1, \text{ if } \sigma=2.
\end{align*}
Let $-2<\xi<-1$. For $\rho \rightarrow 0$ we get
\begin{align*}
\alpha \geq \frac{-\sigma+1}{\sigma\,\rho^{-\xi}} + o\left(\rho^{-1}\right).
\end{align*}
Let $\xi=-2$. For $\rho\rightarrow 0$ we get
\begin{align*}
\alpha \geq \frac{-\sigma+1}{(\sigma+1)\rho^2} + \frac{1}{\rho} + o\left(\rho^{-1}\right).
\end{align*}
Let $\xi<-2$. For $\rho\rightarrow 0$ we get
\begin{align*}
\alpha \geq \frac{-\sigma+1}{\rho^2} + \frac{\sigma}{\rho} + o\left(\rho^{-1}\right).
\end{align*}
Next, let $F^\sigma_\xi$ be an expanding normal velocity.
Here, we use \eqref{id:alpha estimate I}, and \eqref{id:alpha estimate II}.
Let $\xi>-1$. For $\rho\rightarrow 0$ we get
\begin{align*}
\alpha \geq \frac{1}{\sigma\,\rho} + o\left(\rho^{-1}\right).
\end{align*}
Let $\xi=-1$. For $\rho\rightarrow 0$ we get
\begin{align*}
\alpha\geq \frac{-\sigma+2}{\rho} + o\left(\rho^{-1}\right).
\end{align*}
Let $-2<\xi<-1$. For $\rho\rightarrow 0$ we get
\begin{align*}
\alpha \geq \frac{-\sigma+1}{\sigma\,\rho^{-\xi}} + o\left(\rho^{-1}\right).
\end{align*}
Let $\xi=-2$ and $\sigma<-1$. For $\rho\rightarrow 0$ we get
\begin{align*}
\alpha \geq \frac{-\sigma+1}{(\sigma+1)\rho^2} + \frac{1}{\rho} + o\left(\rho^{-1}\right).
\end{align*}
Let $\xi=-2$ and $\sigma=-1$. For $\rho\rightarrow 0$ we get
\begin{align*}
\alpha \leq \frac{1}{\rho^3}.
\end{align*}
Let $\xi=-2$ and $-1<\sigma<0$. For $\rho\rightarrow 0$ we get
\begin{align*}
\alpha \leq \frac{-\sigma+1}{(\sigma+1)\rho^2} + \frac{1}{\rho} + o\left(\rho^{-1}\right).
\end{align*}
Let $\xi<-2$. For $\rho\rightarrow 0$ we get
\begin{align*}
\alpha \leq \frac{-\sigma+1}{\rho^2} + \frac{\sigma}{\rho} + o\left(\rho^{-1}\right).
\end{align*}
This concludes the proof.
\end{proof}
\begin{lemma}[RMPF fulfill $\alpha$-condition]\label{lem:rmpf}
Let $F^\sigma_\xi$ be defined as in Remark \ref{rem:normal velocity xi}.
Let $w$ be a RMPF for a contracting normal velocity $F^\sigma_\xi$, i.\,e.\ $\sigma>0$,
with $\xi>0$, or $\xi=0$, and $1<\sigma<2$.
Then for $\rho\rightarrow 0$ we get
\begin{align*}
\alpha = \frac{c}{\rho} + o\left(\rho^{-1}\right)
\end{align*}
for some $c>0$, which is the second part of MPF condition \eqref{V}.
The function $w=(a-b)^2$ from \cite{ba:gauss} is a RMPF
for $F^\sigma_\xi$ with $\xi=0$ and $\sigma=2$.
For $\rho\rightarrow 0$ we get
\begin{align*}
\alpha=1,
\end{align*}
which is the first part of MPF condition \eqref{V}.
\end{lemma}
\begin{proof}
Let $w \not\equiv 0$ be a rational symmetric homogeneous function of degree $\chi\in{\mathbb{R}}\setminus\{0\}$.
Then $w$ has the form
\begin{align*}
w = \frac{p(a,b)}{q(a,b)\cdot(a\,b)^l},
\end{align*}
for some $l\in{\mathbb{Z}}$, and some symmetric homogeneous polynomials $p(a,b)$, and $q(a,b)$, of degree $g$, and $h$, respectively.
\begin{align*}
p\left(a,b\right) =&\,\sum_{i=0}^{\lfloor g/2\rfloor} c_{i+1}\left(a^{g-i}b^i + a^ib^{g-i}\right), \\
q\left(a,b\right) =&\,\sum_{j=0}^{\lfloor h/2\rfloor} d_{j+1}\left(a^{h-j}b^j + a^jb^{h-j}\right),
\end{align*}
where $c_1>0$, and $d_1>0$, respectively.
Next, we calculate
\begin{align*}
\alpha = -\frac{w_a}{w_b}
= -\frac{b\left(p\,q\,l+a\left(p\,q_a-p_a\,q\right)\right)}{a\left(p\,q\,l+b\left(p\,q_b-p_b\,q\right)\right)}.
\end{align*}
Setting $(a,b)=(\rho,1)$, where $0<\rho<1$, yields
\begin{align}\label{id:alpha p q}
\alpha =&\, -\frac{p\,q\,l+\rho\left(p\,q_a-p_a\,q\right)}{\rho\left(p\,q\,l+p\,q_b-p_b\,q\right)}.
\end{align}
Firstly, we calculate
\begin{align*}
p(a,b)=&\,\sum_{i=0}^{\lfloor g/2\rfloor} c_{i+1}\left(a^{g-i}b^i + a^ib^{g-i}\right), \\
p(\rho,1)=&\,\sum_{i=0}^{\lfloor g/2\rfloor} c_{i+1}\left(\rho^{g-i} + \rho^i\right), \\
p(\rho,0)=&\,c_1. \Bigg .\Bigg.
\end{align*}
Secondly, we calculate
\begin{align*}
p_a(a,b) =&\,\sum_{i=0}^{\lfloor g/2\rfloor} c_{i+1}\left(\left(g-i\right)a^{g-i-1}b^i + i\,a^{i-1}b^{g-i}\right), \\
p_a(\rho,1) =&\,\sum_{i=0}^{\lfloor g/2\rfloor} c_{i+1}\left(\left(g-i\right)\rho^{g-i-1} + i\,\rho^{i-1}\right), \\
p_a(0,1) =&\, c_2. \Bigg .\Bigg.
\end{align*}
Thirdly, we calculate
\begin{align*}
p_b(a,b) =&\,\sum_{i=0}^{\lfloor g/2\rfloor} c_{i+1}\left(i\,a^{g-i}b^{i-1} + \left(g-i\right)a^i b^{g-i-1}\right), \\
p_b(\rho,1) =&\,\sum_{i=0}^{\lfloor g/2\rfloor} c_{i+1}\left(i\,\rho^{g-i} + \left(g-i\right)\rho^i\right), \\
p_b(0,1) =&\,g\,c_1. \Bigg .\Bigg.
\end{align*}
Analogously, we obtain
\begin{align*}
q(0,1) =&\, d_1, \\
q_a(0,1) =&\, d_2, \\
q_b(0,1) =&\, h\,d_1.
\end{align*}
Now, we can combine identity \eqref{id:alpha p q}, and the calculations for $p$, $p_a$, $p_b$,
and $q$, $q_a$, $q_b$.
Hence, for $\rho\rightarrow 0$ we get
\begin{align}\label{id:alpha l}
\alpha=\frac{l}{\rho(g-h-l)} + o\left(\rho^{-1}\right).
\end{align}
Due to RMPF condition \eqref{rII}, we have $\chi>0$, if $F$ is a contracting normal velocity.
Also we have $l\geq 0$, which is a direct consequence of
RMPF conditions \eqref{rI} and \eqref{rIII}.
This implies,
\begin{align*}
g-h-l\geq g-h-2l=\chi>0.
\end{align*}
By Lemma \ref{lem:alpha o estimates}, for $\rho \rightarrow 0$ we get
\begin{align*}
\Big. \Big. \alpha \geq&\, \frac{1}{\sigma\,\rho} + o\left(\rho^{-1}\right), \text{ if } \xi>0, \\
\Big. \Big. \alpha \geq&\, \frac{-\sigma+2}{\rho} + o\left(\rho^{-1}\right), \text{ if } \xi=0 \text{ and } 0<\sigma<2,
\end{align*}
Combining these $o$-estimates with identity \eqref{id:alpha l} yields $l>0$, if $\xi>0$, or $\xi=0$ and $0<\sigma<2$.
This translates into the second part of MPF condition \eqref{V}.
Here, the condition on $\alpha$ implies the condition on $\alpha_a$ since $w$ is assumed to be a rational function.
The part on $w=(a-b)^2$ follows by direct calculations.
This concludes the proof.
\end{proof}
\section{Necessary conditions for the existence of MPF}
In this chapter, we start with Euler's Theorem on homogeneous functions.
This is Theorem \ref{thm:euler} and implies Corollary \ref{cor:euler}.
We use Euler's Theorem \ref{thm:euler} in Lemma \ref{lem:one}, and in Lemma \ref{lem:zeroone}.
Here, we show that there are no MPF for contracting normal velocities
which are homogeneous of degree $\sigma\in(0,1]$.
In Lemma \ref{lem:necessary}, we use Euler's Theorem \ref{thm:euler} and
its Corollary \ref{cor:euler}.
Here, we reformulate MPF condition \eqref{IV}, i.\,e.\
$C_w(a,b)\leq 0$,
$E_w(a,b)\leq 0$, $G_w(a,b)\leq 0$ for all $0<a,b$.
We replace the function $w$ and the normal velocity $F$
by the quantities $\alpha=-\frac{w_a}{w_b}$, and
$\beta=\frac{F_a}{F_b}$. We set $(a,b)=(\rho,1)$, where $0<\rho<1$.
We obtain the three necessary conditions
$C^\alpha_\beta(\rho)\leq 0$, $E^\alpha_\beta(\rho)\leq 0$, and $G^\alpha_\beta(\rho)\leq 0$
for all $0<\rho<1$.
In particular, these new constant and gradient terms now match the form of
MPF condition \eqref{V} ($\alpha$-conditions).
Combining the new constant and gradient terms from Lemma \ref{lem:necessary} and
MPF condition \eqref{V} yields Theorem \ref{thm:necessary}.
We are interested in MPF $w$ for normal velocities $F^\sigma_\xi$ as defined in
Remark \ref{rem:normal velocity xi}.
Theorem \ref{thm:necessary} gives us necessary conditions for the existence of MPF
in terms of $\xi$ and $\sigma$.
In particular, there exist no MPF for contracting $F^\sigma_\xi$ with $\xi<0$
by Theorem \ref{thm:necessary}.
We discuss further implications of Theorem \ref{thm:necessary}
in the following chapters
on $F^\sigma_0=\sgn(\sigma)\cdot K^{\sigma/2}$,
$F^\sigma_1=\sgn(\sigma)\cdot H^\sigma$,
$F^\sigma_2=\sgn(\sigma)\cdot |A|^\sigma$, and
$F^\sigma_\sigma=\sgn(\sigma)\cdot \tr A^\sigma$.
\subsection{Euler's Theorem on homogeneous functions}
\begin{definition}[Cone]\label{def:set}
Let $\Omega \subset {\mathbb{R}}^n\setminus\{0\}$ be an open set.
We call $\Omega$ a \textit{cone}
if $\forall x\in\Omega : \{r\,x : 0 < r < \infty \} \subset \Omega$.
\end{definition}
\begin{definition}[Homogeneous function]\label{def:homogeneous}
Let $\Omega$ be a cone.
We call $f:\Omega\to{\mathbb{R}}$ a \textit{homogeneous function} of degree $\chi\in{\mathbb{R}}$ if
\begin{align*}
f(s\,x) = s^\chi\,f(x),
\end{align*}
for all $s>0$, and for all $x\in\Omega$.
\end{definition}
\begin{theorem}[Euler's Theorem on homogeneous functions]\label{thm:euler}
Let $\Omega\subset{\mathbb{R}}^2$ be a cone.
Let $f(a,b)\in C^1\left(\Omega\right)$ be a function.
Then $f$ is homogeneous of degree $\chi$, if and only if
\begin{align}\label{eq:euler}
a\,f_a+b\,f_b=\chi\,f.
\end{align}
\end{theorem}
\begin{proof}
We refer to S. Hildebrandt \cite{sh:analysis2}.
\end{proof}
\begin{corollary}[Euler's Corollary]\label{cor:euler}
Let $f(a,b)\in C^2\left({\mathbb{R}}^2\setminus\{0\}\right)$ be
a homogeneous function of degree $\chi$.
Then we have
\begin{align}
a\,f_{aa}+b\,f_{ab}=(\chi-1)\,f_a, \label{eq:euleraa}\\
a\,f_{ab}+b\,f_{bb}=(\chi-1)\,f_b. \label{eq:eulerbb}
\end{align}
\end{corollary}
\begin{proof}
We use Theorem \ref{thm:euler}. First, by differentiating equation \eqref{eq:euler} with respect to $a$, we obtain
equation \eqref{eq:euleraa}. Next, by differentiating equation \eqref{eq:euler} with respect to $b$, we obtain
equation \eqref{eq:eulerbb}.
\end{proof}
\subsection{No MPF for normal velocities with $0 < \hom\,F \leq 1$}
\begin{lemma}[Contracting normal velocities homogeneous of degree one]\label{lem:one}
There exists no MPF for any contracting normal velocity $F$
homogeneous of degree one.
\end{lemma}
\begin{proof}
Suppose there exists a MPF $w$ for some contracting normal velocity
homogeneous of degree one. Then we have by MPF condition \eqref{IV}
\begin{align*}
C_w(a,b)\leq 0
\end{align*}
for all $0<a,b$. Furthermore,
Euler's Theorem \ref{thm:euler} implies
\begin{align*}
\chi\,w =&\, a\,w_a+b\,w_b, \\
\text{and}\qquad F=&\, a\,F_a+b\,F_b.
\end{align*}
Combining these identities with MPF condition \eqref{IV} yields
\begin{align*}
0\geq&\, C_w(a,b) \\
=&\, a\,w_a\left(\left(a^2\,F_a+b^2\,F_b\right)+a\left(F-a\,F_a-b\,F_b\right)\right) \\
&\quad +b\,w_b\left(\left(a^2\,F+b^2\,F_b\right)+b\left(F-a\,F_a-b\,F_b\right)\right) \\
=&\, a\,w_a\left(a^2\,F_a+b^2\,F_b\right)+b\,w_b\left(a^2\,F_a+b\,F_b\right) \\
=&\, \chi\,w\left(a^2\,F_a+b^2\,F_b\right) \\
>&\, 0
\end{align*}
for all $0<a,b$ with $a\neq b$.
Note that $\chi>0$ for contracting normal velocities by MPF condition \eqref{II},
$w>0$ for all $0<a,b$ with $a\neq b$ by MPF condition \eqref{I},
and $F_a,\,F_b>0$ for all $0<a,b$, since $F$ is a normal velocity.
Hence, the claim follows.
\end{proof}
\begin{lemma}[Contracting normal velocities homogeneous of degree between zero and one]\label{lem:zeroone}
There exists no MPF for any contracting normal velocity $F$ homogeneous of degree $\sigma\in(0,1)$.
\end{lemma}
\begin{proof}
Suppose there exists a MPF $w$ for some contracting normal velocity homogeneous of degree $\sigma\in(0,1)$.
Then we have by MPF condition \eqref{IV}
\begin{align*}
C_w(a,b) \leq 0
\end{align*}
for all $0<a,b$. Furthermore, Euler's Theorem \ref{thm:euler} implies
\begin{align*}
\chi\,w =&\, a\,w_a+b\,w_b, \\
\text{and}\qquad \sigma\,F =&\, a\,F_a+b\,F_b.
\end{align*}
Combining these identities with MPF condition \eqref{IV} yields
\begin{align*}
0 \geq&\, C_w(a,b) \\
=&\, a\,w_a\left(\left(a^2\,F_a+b^2\,F_b\right)+a\left(F-a\,F_a-b\,F_b\right)\right) \\
&\quad +b\,w_b\left(\left(a^2\,F_a+b^2\,F_b\right)+b\left(F-a\,F_a-b\,F_b\right)\right) \\
=&\,(a\,w_a+b\,w_b)\left(a^2\,F_a+b^2\,F_b\right)+\left(a^2\,w_a+b^2\,w_b\right)(F-a\,F_a-b\,F_b) \\
=&\, \chi\,w\left(a^2\,F_a+b^2\,F_b\right)+(1-\sigma)F\left(a^2\,w_a+b^2\,w_b\right).
\end{align*}
Note that $\chi>0$ for contracting normal velocities by MPF condition \eqref{II},
$w>0$ for all $0<a,b$ with $a\neq b$ by MPF condition \eqref{I},
and $F,\,F_a,\,F_b>0$ for all $0<a,b$, since $F$ is a contracting normal velocity.
So we get
\begin{align*}
0\geq a^2\,w_a+b^2\,w_b
\end{align*}
for all $0<a,b$.
By MPF condition \eqref{III}, we have
\begin{align*}
w_a <&\, 0, \\
w_b >&\, 0
\end{align*}
for all $0<a<b$, see proof of Lemma \ref{lem:alpha} for details.
Using the last three inequalities we obtain
\begin{align*}
\alpha = -\frac{w_a}{w_b} \geq \frac{b^2}{a^2}
\end{align*}
for all $0<a<b$.
Setting $(a,b)=(\rho,1)$, where $0<\rho<1$, yields
\begin{align}\label{ineq:rho2}
\alpha \geq \frac{1}{\rho^2}
\end{align}
for all $0<\rho<1$.
If $F$ is contracting, we have for $\rho\rightarrow 0$
\begin{align*}
\alpha =&\, c, \\
\text{or}\quad \alpha =&\, \frac{c}{\rho} + o\left(\rho^{-1}\right)
\end{align*}
by MPF condition \eqref{V}.
This results in a contradiction to inequality \eqref{ineq:rho2}.
Hence, the claim follows.
\end{proof}
\begin{lemma}[Necessary conditions for the existence of MPF]\label{lem:necessary}
Let $w$ be a MPF for a normal velocity $F$. Set $(a,b)=(\rho,1)$, where $0<\rho<1$.
Then the constant terms $C^\alpha_\beta(\rho)$, and the gradient terms
$E^\alpha_\beta(\rho)$ and $G^\alpha_\beta$ are non-positive for all $0<\rho<1$.
The constant terms are
\begin{align}
C^\alpha_\beta(\rho) = \sgn(\sigma)\cdot\left(\left(-\alpha\,\rho^2+1\right)\left(\beta\,\rho+1\right)+\left(\alpha+\beta\right)\left(-1+\rho\right)\rho\,\sigma\right).
\end{align}
The gradient terms are
\begin{align}
\begin{split}
E^\alpha_\beta(\rho) =&\,-\alpha_a\,\beta(\alpha\,\rho-1)(\beta\,\rho+1)(1-\rho) \\
&\,\quad-\alpha\,\beta_a(\alpha\,\rho-1)^2(1-\rho) \\
&\,\quad-\alpha(\alpha+\beta)(\alpha(1+\rho+(1-\rho)\sigma)
+\beta(1-\rho)(\sigma-1)+2\,\alpha\,\beta\,\rho),
\end{split}
\end{align}
and
\begin{align}
\begin{split}
G^\alpha_\beta(\rho) =&\,-\alpha_a(\alpha\,\rho-1)(\beta\,\rho+1)(1-\rho) \\
&\,\quad+\beta_a(\alpha\,\rho-1)^2(1-\rho) \\
&\,\quad-(\alpha+\beta)(\alpha(1-\rho)(1-\sigma)
+\beta(1+\rho-(1-\rho)\sigma)+2).
\end{split}
\end{align}
\end{lemma}
\begin{proof}
In this proof, we make the general assumption $0<a<b$.
We begin with a few preliminary calculations.
Recall the quantities $\alpha=-\frac{w_a}{w_b}$, and $\beta=\frac{F_a}{F_b}$
from MPF Definition \ref{def:mpf}, and Definition \ref{def:beta}, respectively.
They imply
\begin{align}
w_a =&\, -\alpha\,w_b, \label{id:wa} \\
F_a =&\, \beta\,F_b. \label{id:fa}
\end{align}
Furthermore, we deduce from Euler's Corollary \ref{cor:euler} the identities
\begin{align}
w_{aa} =&\, \frac{(\chi-1)w_a-b\,w_{ab}}{a}, \label{id:waaHelp} \\
w_{bb} =&\, \frac{(\chi-1)w_b-a\,w_{ab}}{b}, \label{id:wbbHelp} \\
F_{aa} =&\, \frac{(\sigma-1)F_a-b\,F_{ab}}{a}, \label{id:faaHelp} \\
F_{bb} =&\, \frac{(\sigma-1)F_b-a\,F_{ab}}{b}. \label{id:fbbHelp}
\end{align}
Calculations for $w_{ab}$ in $\alpha,\alpha_a$:
We have
\begin{align*}
\alpha_a = \frac{\partial}{\partial a}\alpha = \frac{-w_b\,w_{aa}+w_a\,w_{ab}}{w_b^2},
\end{align*}
which implies
\begin{align*}
w_{ab} = \frac{w_b(w_{aa}+w_b\,\alpha_a)}{w_a}.
\end{align*}
Using identities \eqref{id:waaHelp} and \eqref{id:wa} we get
\begin{align}
w_{ab} = \frac{w_b((\chi-1)\alpha-\alpha_a\,a)}{\alpha\,a-b}. \label{id:wab}
\end{align}
Calculations for $w_{aa}$ in $\alpha$, $\alpha$:
Using identities \eqref{id:waaHelp}, \eqref{id:wab}, and \eqref{id:wa} we get
\begin{align}
w_{aa} = \frac{w_b\left(-(\chi-1)\alpha^2+\alpha_a\,b\right)}{\alpha\,a-b}. \label{id:waa}
\end{align}
Calculations for $w_{bb}$ in $\alpha$, $\alpha_a$:
Using identities \eqref{id:wbbHelp}, \eqref{id:wab}, and \eqref{id:wa} we get
\begin{align}
w_{bb} = \frac{w_b\left(\alpha_a\,a^2-(\chi-1)\right)}{(\alpha\,a-b)b}. \label{id:wbb}
\end{align}
Calculations for $F_{ab}$ in $\beta$, $\beta_a$:
We have
\begin{align*}
\beta_a = \frac{\partial}{\partial a} \beta = \frac{F_b\,F_{aa}-F_a\,F_{ab}}{F_b^2},
\end{align*}
which implies
\begin{align*}
F_{ab} = \frac{F_b(F_{aa}-F_b\,\beta_a)}{F_a}.
\end{align*}
Using identities \eqref{id:faaHelp} and \eqref{id:fa} we get
\begin{align}
F_{ab} = \frac{F_b(\beta(\sigma-1)+\beta_a\,a)}{\beta\,a+b}. \label{id:fab}
\end{align}
Calculations for $F_{aa}$ in $\beta$, $\beta_a$:
Using identities \eqref{id:faaHelp}, \eqref{id:fab}, and \eqref{id:fa} we get
\begin{align}
F_{aa} = \frac{F_b\left(\beta^2(\sigma-1)+\beta_a\,b\right)}{(\beta\,a+b)b}. \label{id:faa}
\end{align}
Calculations for $F_{bb}$ in $\beta$, $\beta_a$:
Using identities \eqref{id:fbbHelp}, \eqref{id:fab}, and \eqref{id:fa} we get
\begin{align}
F_{bb} = \frac{F_b\left(\beta_a\,a^2+(\sigma-1)b\right)}{(\beta\,a+b)b}. \label{id:fbb}
\end{align}
Firstly, we combine identities \eqref{id:waa}, \eqref{id:wab}, and \eqref{id:wbb}
to obtain the useful identity
\begin{align}
w_{aa} + 2\,w_{ab}\,\alpha + w_{bb}\,\alpha^2
= \frac{(\alpha\,a-b)\alpha_a}{b}\,w_b. \label{id:I}
\end{align}
Secondly, we combine identities \eqref{id:faa}, \eqref{id:fab}, and \eqref{id:fbb}
to obtain the useful identity
\begin{align}
F_{aa} + 2\,F_{ab}\,\alpha + F_{bb}\,\alpha^2
= \frac{(\alpha+\beta)^2(\sigma-1)b+\beta_a(\alpha\,a-b)^2}{(\beta\,a+b)b}\,F_b. \label{id:II}
\end{align}
Thirdly, we use identities \eqref{id:wa} and \eqref{id:fa}
to obtain the useful identity
\begin{align}
2\,\frac{w_b\,F_a-w_a\,F_b}{a-b} = 2\,\frac{\alpha+\beta}{a-b}\,w_b\,F_b. \label{id:III}
\end{align}
Now, we can further investigate the constant terms $C_w(a,b)$, and
the gradient terms $E_w(a,b)$ and $G_w(a,b)$ from MPF Definition \ref{def:mpf}.
We compute necessary conditions for the existence of MPF:
$C^\alpha_\beta(\rho)$, $E^\alpha_\beta(\rho)$, and $G^\alpha_\beta(\rho)$ have to be
non-positive for all $0<\rho<1$.
\textbf{Constant terms \textit{C}.} By MPF Definition \ref{def:mpf} we have
\begin{align*}
C_w(a,b) = F\left(w_a\,a^2+w_b\,b^2\right)+F_a\,w_b\,a\,b(a-b)-F_b\,w_a\,a\,b(a-b) \leq 0
\end{align*}
for all $0<a,b$.
Using Euler's Theorem \ref{thm:euler} for normal velocity $F$,
i.\,e.\ $F=\frac{1}{\sigma}\left(a\,F_a+b\,F_b\right)$, and identities \eqref{id:wa} and \eqref{id:fa} we obtain
\begin{align*}
C_w(a,b) = \left(\left(-\alpha\,a^2+b^2\right)(\beta\,a+b)-(\alpha+\beta)a\,b(b-a)\sigma\right)\frac{w_b\,F_b}{\sigma} \leq 0
\end{align*}
for all $0<a,b$.
Setting $(a,b)=(\rho,1)$, where $0<\rho<1$, yields
\begin{align*}
C^\alpha_\beta(\rho) = &\, \sgn(\sigma)\cdot\left(\left(-\alpha\,\rho^2+1\right)(\beta\,\rho+1)+(\alpha+\beta)(-1+\rho)\rho\,\sigma\right) \\
= &\, C_w(\rho,1) \frac{|\sigma|}{w_b\,F_b} \\
\leq &\, 0
\end{align*}
for all $0 < \rho < 1$. Note that $w_b,\,F_b>0$ for all $0<\rho<1$. \\
\textbf{Gradient terms \textit{E}.} By MPF Definition \ref{def:mpf}, we have
\begin{align*}
E_w(a,b) = &\, w_a \left(F_{aa} + 2\,F_{ab}\,\alpha + F_{bb}\,\alpha^2\right) \\
&\quad -F_a \left(w_{aa} + 2\,w_{ab}\,\alpha + w_{bb}\,\alpha^2\right) \\
&\quad +2\,\frac{w_b\,F_a-w_a\,F_b}{a-b}\,\alpha^2, \\
\leq &\, 0
\end{align*}
for all $0<a,b$. \\
Using identities \eqref{id:I}, \eqref{id:II}, and \eqref{id:III},
we obtain
\begin{align*}
E_w(a,b) = &\, w_a \left(\frac{(\alpha+\beta)^2(\sigma-1)b+\beta_a(\alpha\,a-b)^2}{(\beta\,a+b)b}\,F_b\right) \\
&\quad -F_a \left(\frac{(\alpha\,a-b)\alpha_a}{b}\,w_b\right) \\
&\quad +\left(2\,\frac{\alpha+\beta}{a-b}\,w_b\,F_b\right)\alpha^2, \\
\leq &\, 0
\end{align*}
for all $0<a,b$.
We use identities \eqref{id:wa} and \eqref{id:fa} and set $(a,b)=(\rho,1)$, where $0<\rho<1$.
This yields
\begin{align*}
E^\alpha_\beta(\rho) = &\, -\alpha_a\,\beta(\alpha\,\rho-1)(\beta\,\rho+1)(1-\rho) \\
&\, \quad-\alpha\,\beta_a(\alpha\,\rho-1)^2(1-\rho) \\
&\, \quad-\alpha(\alpha+\beta)(\alpha(1+\rho+(1-\rho)\sigma)
+ \beta(1-\rho)(\sigma-1)+2\,\alpha\,\beta\,\rho), \\
= &\, E_w(\rho,1)\,\frac{(\beta\,\rho+1)(1-\rho)}{w_b\,F_b} \\
\leq &\, 0
\end{align*}
for all $0<\rho<1$. Note that $w_b,\,F_b>0$ for all $0<\rho<1$. \\
\textbf{Gradient terms \textit{G}.} By MPF Definition \eqref{def:mpf}, we have
\begin{align*}
G_w(a,b) = &\, \frac{1}{\alpha^2}\cdot\bigg(w_b \left(F_{aa} + 2\,F_{ab}\,\alpha + F_{bb}\,\alpha^2\right)\bigg. \\
& \qquad\qquad -F_b \left(w_{aa} + 2\,w_{ab}\,\alpha + w_{bb}\,\alpha^2\right) \\
& \qquad\qquad \bigg. +2\, \frac{w_b\,F_a - w_a\,F_b}{a-b}\bigg) \\
\leq &\, 0
\end{align*}
for all $0<a,b$.
We use identities \eqref{id:wa} and \eqref{id:fa} and set $(a,b)=(\rho,1)$, where $0<\rho<1$.
This yields
\begin{align*}
G_w(a,b) = &\, \frac{1}{\alpha^2}\cdot\bigg(w_b \left(\frac{(\alpha+\beta)^2(\sigma-1)b+\beta_a(\alpha\,a-b)^2}{(\beta\,a+b)b}\,F_b\right)\bigg. \\
& \qquad\qquad -F_b \left(\frac{(\alpha\,a-b)\alpha_a}{b}\,w_b\right) \\
& \qquad\qquad \bigg. +2\, \frac{\alpha+\beta}{a-b}\,w_b\,F_b\bigg) \\
\leq &\, 0
\end{align*}
for all $0<a,b$.
We use identities \eqref{id:wa} and \eqref{id:fa} and set $(a,b)=(\rho,1)$, where $0<\rho<1$.
This yields
\begin{align*}
G^\alpha_\beta(\rho) = &\, -\alpha_a(\alpha\,\rho-1)(\beta\,\rho+1)(1-\rho) \\
&\, \quad+\beta_a(\alpha\,\rho-1)^2(1-\rho) \\
&\, \quad-(\alpha+\beta)(\alpha(1-\rho)(1-\sigma)
+ \beta(1+\rho-(1-\rho)\sigma)+2) \\
= &\, G_w(\rho,1)\,\frac{\alpha^2(\beta\,\rho+1)(1-\rho)}{w_b\,F_b} \\
\leq &\, 0
\end{align*}
for all $0<\rho<1$. Note that $w_b,\,F_b>0$ for all $0<\rho<1$.
This concludes the proof.
\end{proof}
\subsection{When MPF cease to exist for normal velocities $F^\sigma_\xi$}
\begin{theorem}[Necessary conditions for the existence of MPF]\label{thm:necessary}
Let $w$ be a MPF.
Let $F^\sigma_\xi$ be defined as in Remark \ref{rem:normal velocity xi}.
Then the necessary conditions from Lemma \ref{lem:necessary}
imply the following necessary conditions for the existence of MPF
for contracting normal velocities $F^\sigma_\xi$, and
for the existence of MPF
for expanding normal velocities $F^\sigma_\xi$.
\begin{center}
\begin{tabular}{| l | c || c |}
\hline
& $\Big. \Big.$ contracting $F^\sigma_\xi$, $\sigma>1$ & expanding $F^\sigma_\xi$, $\sigma<0$ \\
\hline \hline
$\Big. \Big.$ $\xi>0$ & $c=1/\sigma$ & $\sigma\in [-1,0)$ \\
\hline
$\Big. \Big.$ $\xi=0$ & $\sigma\in(1,2]$ & $\sigma\in[-2,0)$ \\
\hline
$\Big. \Big.$ $\xi<0$ & MPF non-existent & open problem \\
\hline
\end{tabular}
\end{center}
Note that $c>0$ is from MPF Definition \ref{def:mpf}.
\end{theorem}
\begin{proof}
We recall from Remark \ref{rem:normal velocity xi} that
\begin{align}
\beta_{F^\sigma_\xi} = \rho^{\xi-1}.
\end{align}
For our convenience, we calculate the constant terms $C^\alpha_\beta(\rho)$, and
the gradient terms $E^\alpha_\beta(\rho)$ and $G^\alpha_\beta(\rho)$
from Lemma \ref{lem:necessary} for
\begin{align*}
\beta = \rho^\xi,\;\beta_a=\xi\,\rho^{\xi-1}.
\end{align*}
Towards the end of the proof we just need to add one to $\xi$
to obtain the proper results for $F^\sigma_\xi$.
Furthermore, we recall from MPF Definition \ref{def:mpf}
the $\alpha$-conditions:
\begin{align*}
&\,\alpha_1\text{-condition (contracting)}: \\
&\,\qquad \alpha=c,
\;\alpha_a=0, \\
&\,\alpha_2\text{-condition (contracting/expanding)}: \\
&\,\qquad\alpha=\frac{c}{\rho}+o\left(\rho^{-1}\right),
\;\alpha_a=-\frac{c}{\rho^2}+o\left(\rho^{-1}\right), \\
&\,\alpha_3\text{-condition (expanding)}: \\
&\,\qquad\alpha=\frac{c+d\,\rho}{\rho^2}+o\left(\rho^{-1}\right),
\;\alpha_a=-\frac{2\,c+d\,\rho}{\rho^3}+o\left(\rho^{-1}\right).
\end{align*}
The proof idea is the following.
For all $\alpha$-conditions and all $\beta=\rho^{\xi}$, where $\xi\in{\mathbb{R}}$, we calculate
\begin{align*}
{\lim_{\rho\rightarrow 0}\,} \rho^{\xi_C}\cdot C^\alpha_\beta(\rho), \\
{\lim_{\rho\rightarrow 0}\,} \rho^{\xi_E}\cdot E^\alpha_\beta(\rho), \\
{\lim_{\rho\rightarrow 0}\,} \rho^{\xi_G}\cdot G^\alpha_\beta(\rho),
\end{align*}
for some $\xi_C,\xi_E,\xi_G\in{\mathbb{R}}$, which depend on $\xi$.
These calculations are somewhat tedious.
Since $w$ is MPF, these limits have to be non-negative for continuity reasons.
Calculating these limits is the first part of the proof.
In the second part, we draw conclusions from the results
of the first part. Here, we get the necessary conditions
for $F^\sigma_\xi$ in terms of $\xi$, $\sigma$, and $c>0$.
\textbf{Limits of \textit{C}, \textit{E}, and \textit{G}.} \\
We begin by inserting $\beta=\rho^\xi$ and $\beta_a=\xi\,\rho^{\xi-1}$
into $C^\alpha_\beta(\rho)$, $E^\alpha_\beta(\rho)$, and $G^\alpha_\beta(\rho)$.
This yields
\begin{align*}
C^\alpha_\xi(\rho):=&\,
\sgn(\sigma)\left(\left(-\alpha\,\rho^2+1\right)\left(\rho^{\xi+1}+1\right)+\left(\alpha+\rho^\xi\right)\left(-1+\rho\right)\rho\,\sigma\right), \\
E^\alpha_\xi(\rho):=&\,-\alpha_a\,\rho^\xi(\alpha\,\rho-1)(\rho^{\xi+1}+1)(1-\rho) \\
&\,\quad-\alpha\,\xi\,\rho^{\xi-1}(\alpha\,\rho-1)^2(1-\rho) \\
&\,\quad-\alpha(\alpha+\rho^\xi)(\alpha(1+\rho+(1-\rho)\sigma)
+\rho^\xi(1-\rho)(\sigma-1)+2\,\alpha\,\rho^{\xi+1}), \\
G^\alpha_\xi(\rho):=&\,-\alpha_a(\alpha\,\rho-1)(\rho^{\xi+1}+1)(1-\rho) \\
&\,\quad+\xi\,\rho^{\xi-1}(\alpha\,\rho-1)^2(1-\rho) \\
&\,\quad-(\alpha+\rho^\xi)(\alpha(1-\rho)(1-\sigma)
+\rho^\xi(1+\rho-(1-\rho)\sigma)+2).
\end{align*}
Now we calculate the desired limits for each $\alpha$-condition. \\
\textit{Calculate the constant terms $C^\alpha_\xi(\rho)$ for the $\alpha_1$-condition.} \\
Let $\xi>-1$:
\begin{align}\label{eq:x1}
{\lim_{\rho\rightarrow 0}\,} C^{\alpha_1}_{\xi>-1}(\rho) = \sgn(\sigma).
\end{align}
Let $\xi=-1$:
\begin{align}\label{eq:x2}
{\lim_{\rho\rightarrow 0}\,} C^{\alpha_1}_{\xi=-1}(\rho) = \sgn(\sigma)\cdot(-\sigma+2).
\end{align}
Let $\xi<-1$:
\begin{align}
{\lim_{\rho\rightarrow 0}\,} \rho^{-(\xi+1)}\cdot C^{\alpha_1}_{\xi<-1}(\rho) = \sgn(\sigma)\cdot(-\sigma+1).
\end{align}
\textit{Calculate the gradient terms $E^\alpha_\xi(\rho)$ and $G^\alpha_\xi(\rho)$ for the $\alpha_1$-condition.} \\
Let $\xi>1$:
\begin{align}
{\lim_{\rho\rightarrow 0}\,} E^{\alpha_1}_{\xi>1}(\rho) =&\, -c^3(\sigma+1), \\
{\lim_{\rho\rightarrow 0}\,} G^{\alpha_1}_{\xi>1}(\rho) =&\, c(c(\sigma-1)-2).
\end{align}
Let $\xi=1$:
\begin{align}
{\lim_{\rho\rightarrow 0}\,} E^{\alpha_1}_{\xi=1}(\rho) =&\, -c(c^2(\sigma+1)+1), \\
{\lim_{\rho\rightarrow 0}\,} G^{\alpha_1}_{\xi=1}(\rho) =&\, c(c(\sigma-1)-2)+1.
\end{align}
Let $0<\xi<1$:
\begin{align}
{\lim_{\rho\rightarrow 0}\,} \rho^{-(\xi-1)}\cdot E^{\alpha_1}_{0<\xi<1}(\rho) =&\, -c\,\xi, \\
{\lim_{\rho\rightarrow 0}\,} \rho^{-(\xi-1)}\cdot G^{\alpha_1}_{0<\xi<1}(\rho) =&\, \xi.
\end{align}
Let $\xi=0$:
\begin{align}
{\lim_{\rho\rightarrow 0}\,} E^{\alpha_1}_{\xi=0}(\rho) =&\, -c(c+1)(c(\sigma+1)+\sigma-1), \\
{\lim_{\rho\rightarrow 0}\,} G^{\alpha_1}_{\xi=0}(\rho) =&\, (c+1)(c(\sigma-1)+\sigma-3).
\end{align}
Let $-1<\xi<0$:
\begin{align}
{\lim_{\rho\rightarrow 0}\,} \rho^{-(\xi-1)}\cdot E^{\alpha_1}_{-1<\xi<0}(\rho) =&\, -c\,\xi, \\
{\lim_{\rho\rightarrow 0}\,} \rho^{-(\xi-1)}\cdot G^{\alpha_1}_{-1<\xi<0}(\rho) =&\, \xi.
\end{align}
Let $\xi=-1$:
\begin{align}
{\lim_{\rho\rightarrow 0}\,} \rho^2\cdot E^{\alpha_1}_{\xi=-1}(\rho) =&\, -c(\sigma-2), \label{eq:x3} \\
{\lim_{\rho\rightarrow 0}\,} \rho^2\cdot G^{\alpha_1}_{\xi=-1}(\rho) =&\, \sigma-2. \label{eq:x4}
\end{align}
Let $\xi<-1$:
\begin{align}
{\lim_{\rho\rightarrow 0}\,} \rho^{-2\,\xi}\cdot E^{\alpha_1}_{\xi<-1}(\rho) =&\, -c(\sigma-1), \\
{\lim_{\rho\rightarrow 0}\,} \rho^{-2\,\xi}\cdot G^{\alpha_1}_{\xi<-1}(\rho) =&\, \sigma-1. \label{eq:x5}
\end{align}
\textit{Calculate the constant terms $C^\alpha_\xi(\rho)$ for the $\alpha_2$-condition.} \\
Let $\xi>-1$:
\begin{align}\label{eq:x6}
{\lim_{\rho\rightarrow 0}\,} C^{\alpha_2}_{\xi>-1}(\rho) =&\, \sgn(\sigma)(-(c\,\sigma-1)).
\end{align}
Let $\xi=-1$:
\begin{align}\label{eq:x9}
{\lim_{\rho\rightarrow 0}\,} C^{\alpha_2}_{\xi=-1}(\rho) =&\, \sgn(\sigma)(-(c\,\sigma+\sigma-2)).
\end{align}
Let $\xi<-1$
\begin{align}\label{eq:x12}
{\lim_{\rho\rightarrow 0}\,} C^{\alpha_2}_{\xi<-1}(\rho) =&\, \sgn(\sigma)(-\sigma+1).
\end{align}
\textit{Calculate the gradient terms $G^\alpha_\xi(\rho)$ and $E^\alpha_\xi(\rho)$ for the $\alpha_2$-condition.} \\
Let $\xi>-1$:
\begin{align}
{\lim_{\rho\rightarrow 0}\,} \rho^3\cdot E^{\alpha_2}_{\xi>-1}(\rho) =&\, -c^3(\sigma+1), \label{eq:x7} \\
{\lim_{\rho\rightarrow 0}\,} \rho^2\cdot G^{\alpha_2}_{\xi>-1}(\rho) =&\, -(-c\,\sigma+1). \label{eq:x8}
\end{align}
Let $\xi=-1$:
\begin{align}
{\lim_{\rho\rightarrow 0}\,} \rho^3\cdot E^{\alpha_2}_{\xi=-1}(\rho) =&\, -c(c+1)(c\,\sigma+2\,c+\sigma), \label{eq:x10} \\
{\lim_{\rho\rightarrow 0}\,} \rho^2\cdot G^{\alpha_2}_{\xi=-1}(\rho) =&\, (c+1)(c\,\sigma+\sigma-2). \label{eq:x11}
\end{align}
Let $\xi<-1$:
\begin{align}
{\lim_{\rho\rightarrow 0}\,} \rho^{-(2\,\xi-1)}\cdot E^{\alpha_2}_{\xi<-1}(\rho) =&\, -c(c+\sigma), \label{eq:x13} \\
{\lim_{\rho\rightarrow 0}\,} \rho^{-2\,\xi}\cdot G^{\alpha_2}_{\xi<-1}(\rho) =&\, \sigma-1. \label{eq:x14}
\end{align}
\textit{Calculate the constant terms $C^\alpha_\xi(\rho)$ for the $\alpha_3$-condition.} \\
Let $\xi>-2$:
\begin{align}\label{eq:x15}
{\lim_{\rho\rightarrow 0}\,} \rho\cdot C^{\alpha_3}_{\xi>-2}(\rho) =&\, c\,\sigma.
\end{align}
Let $\xi=-2$:
\begin{align}\label{eq:x22}
{\lim_{\rho\rightarrow 0}\,} \rho\cdot C^{\alpha_3}_{\xi=-2}(\rho) =&\, c\,\sigma+c+\sigma-1.
\end{align}
Let $\xi<-2$:
\begin{align}\label{eq:x25}
{\lim_{\rho\rightarrow 0}\,} \rho^{-(\xi+1)}\cdot C^{\alpha_3}_{\xi<-2}(\rho) =&\, c+\sigma-1.
\end{align}
\textit{Calculate the gradient terms $E^\alpha_\xi(\rho)$ and $G^\alpha_\xi(\rho)$ for the $\alpha_3$-conditions.} \\
Let $\xi>-1$:
\begin{align}
{\lim_{\rho\rightarrow 0}\,} \rho^6\cdot E^{\alpha_3}_{\xi>-1}(\rho) =&\, -c^3(\sigma+1), \label{eq:x16} \\
{\lim_{\rho\rightarrow 0}\,} \rho^4\cdot G^{\alpha_3}_{\xi>-1}(\rho) =&\, c^2(\sigma+1). \label{eq:x17}
\end{align}
Let $\xi=-1$:
\begin{align}
{\lim_{\rho\rightarrow 0}\,} \rho^6\cdot E^{\alpha_3}_{\xi=-1}(\rho) =&\, -c^3(\sigma+2), \label{eq:x18} \\
{\lim_{\rho\rightarrow 0}\,} \rho^4\cdot G^{\alpha_3}_{\xi=-1}(\rho) =&\, c^2(\sigma+2). \label{eq:x19}
\end{align}
Let $-2<\xi<-1$:
\begin{align}
{\lim_{\rho\rightarrow 0}\,} \rho^{-(\xi-5)}\cdot E^{\alpha_3}_{-2<\xi<-1}(\rho) =&\, -c^3(\xi+2), \label{eq:x20} \\
{\lim_{\rho\rightarrow 0}\,} \rho^{-(\xi-2)}\cdot G^{\alpha_3}_{-2<\xi<-1}(\rho) =&\, c^2(\xi+2). \label{eq:x21}
\end{align}
Let $\xi=-2$:
\begin{align}
{\lim_{\rho\rightarrow 0}\,} \rho^6\cdot E^{\alpha_3}_{\xi=-2}(\rho) =&\, -c(c^2(\sigma+3)+2\,c(\sigma+2)+\sigma+1+d), \label{eq:x23} \\
{\lim_{\rho\rightarrow 0}\,} \rho^4\cdot G^{\alpha_3}_{\xi=-2}(\rho) =&\, c^2(\sigma+1)+2\,c\,\sigma+\sigma-1-c\,d. \label{eq:x24}
\end{align}
Let $-3<\xi<-2$:
\begin{align}
{\lim_{\rho\rightarrow 0}\,} \rho^{-(\xi-5)}\cdot E^{\alpha_3}_{-3<\xi<-2}(\rho) =&\, -c^3(\xi+2), \label{eq:x26} \\
{\lim_{\rho\rightarrow 0}\,} \rho^{-(\xi-3)}\cdot G^{\alpha_3}_{-3<\xi<-2}(\rho) =&\, -c^3(\xi+2). \label{eq:x27}
\end{align}
Let $\xi=-3$:
\begin{align}
{\lim_{\rho\rightarrow 0}\,} \rho^8\cdot E^{\alpha_3}_{\xi=-3}(\rho) =&\, c(c^2-2\,c-\sigma-1-d), \label{eq:x28} \\
{\lim_{\rho\rightarrow 0}\,} \rho^6\cdot G^{\alpha_3}_{\xi=-3}(\rho) =&\, -c^2+\sigma-1. \label{eq:x29}
\end{align}
Let $\xi<-3$:
\begin{align}
{\lim_{\rho\rightarrow 0}\,} \rho^{-(2\,\xi-2)}\cdot E^{\alpha_3}_{\xi<-3}(\rho) =&\, -c(2\,c+\sigma+1+d), \label{eq:x30} \\
{\lim_{\rho\rightarrow 0}\,} \rho^{-2\,\xi}\cdot G^{\alpha_3}_{\xi<-3}(\rho) =&\, \sigma-1. \label{eq:x31}
\end{align}
\textbf{Conclusions for the limits of \textit{C}, \textit{E}, and \textit{G}.} \\
\textit{Conclusions for the $\alpha_1$-condition in the contracting case, $\sigma>1$.} \\
Let $\xi>-1$: \\
By \eqref{eq:x1} we get the necessary condition
\begin{align*}
1 \leq 0,
\end{align*}
thus, here a MPF $w$ cannot exist. \\
Let $\xi=-1$: \\
By \eqref{eq:x2}, \eqref{eq:x3}, \eqref{eq:x4}, we get the necessary conditions
\begin{align*}
-\sigma+2\leq&\, 0, \\
-\sigma+2\leq&\, 0, \\
\sigma-2\leq&\, 0.
\end{align*}
Thus, here a MPF $w$ cannot exist if $\sigma\neq 2$. \\
Let $\xi<-1$: \\
By \eqref{eq:x5}, we get the necessary condition
\begin{align*}
\sigma-1 \leq&\, 0,
\end{align*}
thus, here a MPF $w$ cannot exist. \\
Therefore, a MPF $w$ can only exist if
\begin{align}\label{eq:xI}
\xi=-1, \text{ and } \sigma=2.
\end{align}
\newline
\textit{Conclusions for the $\alpha_2$-condition in the contracting case, $\sigma>1$.} \\
Let $\xi>-1$:
By \eqref{eq:x6}, \eqref{eq:x7}, \eqref{eq:x8}, we get the necessary conditions
\begin{align*}
-(c\,\sigma-1) \leq&\, 0, \\
-(\sigma+1) \leq&\, 0, \\
c\,\sigma-1 \leq&\, 0.
\end{align*}
Thus, here a MPF $w$ cannot exist if $\sigma\neq \frac{1}{\sigma}$. \\
Let $\xi=-1$: \\
By \eqref{eq:x9}, \eqref{eq:x10}, \eqref{eq:x11}, we get the necessary conditions,
\begin{align*}
-(c\,\sigma+\sigma-2) \leq 0, \\
-(c\,\sigma+2\,c+\sigma) \leq 0, \\
c\,\sigma+\sigma-2 \leq 0.
\end{align*}
This implies $\frac{2-\sigma}{\sigma}=c>0$.
Thus, here a MPF $w$ cannot exist if $\sigma\not\in(0,2)$. \\
Let $\xi<-1$: \\
By \eqref{eq:x14} we get the necessary condition,
\begin{align*}
\sigma-1 \leq 0,
\end{align*}
thus, here a MPF $w$ cannot exist. \\
Therefore, a MPF $w$ can only exist if
\begin{align}\label{eq:xII}
\begin{split}
\xi>&\, -1, \text{ and } c=\frac{1}{\sigma}, \\
\xi=&\, -1, \text{ and } \sigma \in (0,2), \\
\xi<&\, -1, \text{ MPF non-existent}.
\end{split}
\end{align}
\textit{Conclusions for the $\alpha_2$-condition in the expanding case, $\sigma<0$.} \\
Let $\xi>-1$: \\
By \eqref{eq:x6}, \eqref{eq:x7}, \eqref{eq:x8}, we get the necessary conditions,
\begin{align*}
c\,\sigma-1 \leq&\, 0, \\
-(\sigma-1) \leq&\, 0, \\
c\,\sigma-1 \leq&\, 0,
\end{align*}
thus, here MPF $w$ cannot exist if $\sigma\not\in [-1,0)$. \\
Let $\xi=-1$: \\
By \eqref{eq:x9}, \eqref{eq:x10}, \eqref{eq:x11}, we get the necessary conditions
\begin{align*}
c\,\sigma+\sigma-2 \leq&\, 0, \\
-c\,\sigma-2\,c-\sigma \leq&\, 0 \\
c\,\sigma+\sigma-2 \leq&\, 0,
\end{align*}
thus, here MPF cannot exist if $\sigma\not\in(-2,0)$. \\
Let $\xi<-1$: \\
By \eqref{eq:x12}, \eqref{eq:x13}, \eqref{eq:x14}, we get the necessary conditions
\begin{align*}
\sigma-1 \leq&\, 0, \\
-(c+\sigma) \leq&\, 0, \\
\sigma-1 \leq&\, 0,
\end{align*}
thus, here we cannot determine any further necessary conditions. \\
Therefore, a MPF $w$ can only exist if
\begin{align}
\begin{split}\label{eq:xIII}
\xi>&\, -1, \text{ and } \sigma\in[-1,0), \\
\xi=&\, -1, \text{ and } \sigma\in(-2,0), \\
\xi<&\, -1, \text{ no further necessary conditions}.
\end{split}
\end{align}
\textit{Conclusions for the $\alpha_3$-condition in the contracting case, $\sigma<0$.} \\
Let $\xi>-1$: \\
By \eqref{eq:x15}, \eqref{eq:x16}, \eqref{eq:x17}, we get the necessary conclusions
\begin{align*}
c\,\sigma \leq&\, 0, \\
-(\sigma+1) \leq&\, 0, \\
(\sigma+1) \leq&\, 0,
\end{align*}
thus, here a MPF $w$ cannot exist if $\sigma\neq -1$. \\
Let $\xi=-1$: \\
By \eqref{eq:x15}, \eqref{eq:x18}, \eqref{eq:x19}, we get the necessary conditions
\begin{align*}
c\,\sigma \leq&\, 0, \\
-(\sigma+2) \leq&\, 0, \\
\sigma+2\leq&\, 0,
\end{align*}
thus, here a MPF $w$ cannot exist if $\sigma\neq -2$. \\
Let $-2<\xi<-1$: \\
By \eqref{eq:x15}, \eqref{eq:x20}, \eqref{eq:x21}, we get the necessary conditions
\begin{align*}
c\,\sigma \leq&\, 0, \\
-(\xi+2) \leq&\, 0, \\
\xi+2 \leq&\, 0,
\end{align*}
thus, here a MPF $w$ cannot exist. \\
Let $\xi=-2$: \\
By \eqref{eq:x22}, \eqref{eq:x23}, \eqref{eq:x24}, we get the necessary conditions
\begin{align*}
c\,\sigma+c+\sigma-1 \leq&\, 0, \\
-(c^2(\sigma+3)+2\,c(\sigma+2)+\sigma+1+d) \leq&\, 0, \\
c^2(\sigma+1)+2\,c\,\sigma+\sigma-1-c\,d \leq&\, 0,
\end{align*}
thus, here we cannot determine any further necessary conditions. \\
Let $-3<\xi<-2$: \\
By \eqref{eq:x25}, \eqref{eq:x26}, \eqref{eq:x27}, we get the necessary conditions
\begin{align*}
c+\sigma-1\leq&\, 0, \\
-(\xi+2) \leq&\, 0, \\
\xi+2 \leq&\, 0,
\end{align*}
thus, here a MPF $w$ cannot exist. \\
Let $\xi=-3$: \\
By \eqref{eq:x25}, \eqref{eq:x28}, \eqref{eq:x29}, we get the necessary conditions
\begin{align*}
c+\sigma-1\leq&\, 0, \\
c^3-2\,c-\sigma-1-d\leq&\, 0, \\
-c^2+\sigma-1\leq&\, 0,
\end{align*}
thus, here we cannot determine any further necessary conditions. \\
Let $\xi<-3$: \\
By \eqref{eq:x25}, \eqref{eq:x30}, \eqref{eq:x31}, we get the necessary conditions
\begin{align*}
c+\sigma-1 \leq&\, 0, \\
-2\,c-\sigma+1+d \leq&\, 0, \\
\sigma-1 \leq&\, 0,
\end{align*}
thus, here we cannot determine any further necessary conditions. \\
Therefore, a MPF $w$ can only exist if
\begin{align}\label{eq:xIV}
\begin{split}
\xi>&\, -1, \text{ and } \sigma=-1, \\
\xi=&\, -1, \text{ and } \sigma=-2, \\
-2<&\,\xi<-1, \text{ MPF non-existent}, \\
\xi=&\,-2, \text{ no further necessary conditions}, \\
-3<&\,\xi<-2: \text{ MPF non-existent}, \\
\xi=&\,-3, \text{ no further necessary conditions}, \\
\xi<&\,-3, \text{ no further ncessary conditions}.
\end{split}
\end{align}
As mentioned in the beginning of the proof, we need to add one to $\xi$
to obtain the proper results for $F^\sigma_\xi$ since $\beta_{F^\sigma_\xi} = \rho^{\xi-1}$. \\
\textbf{Conclusions in the contracting case, and in the expanding case.} \\
Consider $F^\sigma_\xi$, where $\sigma>1$.
MPF $w$ can only exist in the contracting case if the $\alpha_1$-condition, or if the corresponding $\alpha_2$-condition is met.
Combining \eqref{eq:xI} and \eqref{eq:xII},
we obtain the necessary conditions
\begin{align*}
\xi>&\, 0:\;c=\frac{1}{\sigma}, \\
\xi=&\, 0:\;\sigma\in(0,2], \\
\xi<&\, 0:\;\text{MPF non-existent}.
\end{align*}
Consider $F^\sigma_\xi$, where $\sigma<0$.
MPF $w$ can only exist in the expanding case if the corresponding $\alpha_2$-condition, or if the $\alpha_3$-condition is met.
Combining \eqref{eq:xIII} and \eqref{eq:xIV},
we obtain the necessary conditions
\begin{align*}
\xi>&\, 0:\;\sigma\in[-1,0), \\
\xi=&\, 0:\;\sigma\in[-2,0), \\
\xi<&\, 0:\;\text{no further necessary conditions}.
\end{align*}
This concludes the proof.
\end{proof}
\section{Vanishing functions}
G. Huisken \cite{gh:flow} uses a function similar to
\begin{align*}
v=\frac{(a-b)^2}{(a+b)^2}
\end{align*}
to show convergence to a round point.
Here, $v$ fulfills all MPF conditions except condition \eqref{II},
since $v$ is homogeneous of degree zero.
But it is a vanishing function,
i.\,e.\ a function whose constant terms $C_v(a,b)$ of
MPF condition \eqref{IV} vanish for all $0<a,b$.
B. Andrews \cite{ba:gauss} uses the MPF
\begin{align*}
v=(a-b)^2
\end{align*}
to show convergence to a round point for $F=K$,
and O. Schn\"urer and F. Schulze \cite{fs:convexity}
use the MPF
\begin{align}\label{id:MPF Schnulze}
v=\frac{(a-b)^2(a+b)^{2\sigma}}{(a\,b)^2}
\end{align}
to show a corresponding result for $F=H^\sigma$ with $1<\sigma\leq 5$.
All of these functions are vanishing functions.
In fact, vanishing functions seem to be the natural set of functions,
when it comes to MPF for contracting normal velocities $F^\sigma_\xi$
with $\xi\geq 0$.
We conjecture that, if a vanishing function is not MPF
for a contracting normal velocity $F^\sigma_\xi$ with $\xi\geq 0$, then there just exist
no MPF for that $F^\sigma_\xi$.
For other contracting and expanding $F^\sigma_\xi$,
this is different. Here,
vanishing functions do not play such an important role.
By Lemma \ref{lem:alpha o estimates}, we know that
vanishing functions cannot be MPF in most cases.
We see this by interchanging the $\leq$ and $\geq$
by equal signs in the $o$-estimates of this Lemma.
That is basically what we do in Lemma \ref{lem:alpha vanishing V}.
However, for contracting normal velocities $F^\sigma_\xi$ with $\xi\geq 0$,
vanishing functions remain interesting.
In later chapters, the MPF tables for normal velocities
$F^\sigma_0=\sgn(\sigma)\cdot K^{\sigma/2}$,
$F^\sigma_1=\sgn(\sigma)\cdot H^\sigma$,
$F^\sigma_2=\sgn(\sigma)\cdot |A|^\sigma$,
and
$F^\sigma_\sigma=\sgn(\sigma)\cdot \tr A^\sigma$,
all contain vanishing functions.
In Lemma \ref{lem:alpha vanishing VI} we see that in some sense
they are the optimal fit.
In Remark \ref{rem:not vanishing}, we present
some MPF for $F=H^\sigma$ with $\sigma=3,4$
by O. Schn\"urer \cite{os:surfacesA2}, and with $\sigma=2,5$,
which are not vanishing functions.
We want to stress that vanishing functions are useful
but not the only existing MPF.
The rest of the somewhat technical Lemmas
on vanishing functions
is needed in later chapters.
\begin{definition}[Vanishing functions]\label{def:vanishing}
Let $a$ and $b$ be principal curvatures.
Let $v(a,b)\in C^2\left({\mathbb{R}}^2_{+}\right)$ with $v\not\equiv 0$.
Let $C_w(a,b)$ be defined as in MPF condition \eqref{IV}.
We call $v$ a \textit{vanishing function}
for a normal velocity $F$ if
$C_v(a,b)=0$
for all $0<a,b$.
\end{definition}
\begin{example}[Vanishing functions]\label{exm:vanishing}
By O. Schn\"urer and F. Schulze \cite{fs:convexity},
and B. Andrews and X. Chen \cite{ac:surfaces},
we have the following example of a
vanishing function for a normal velocity $F$
\begin{align*}
v=\frac{(a-b)^2\,F^2}{(a\,b)^2}.
\end{align*}
Note that $F$ can be an arbitrary normal velocity.
\end{example}
\begin{remark}[MPF which are not vanishing functions]\label{rem:not vanishing}
We give examples of MPF which are not vanishing functions.
Let $F=F^\sigma_1=\sgn(\sigma)\cdot H^\sigma$. Then we have
\begin{center}
\begin{tabular}{| l | c || c | c |}
\hline
& MPF \, $w_{H^\sigma}$ & & \\
\hline \hline
& & & \\
$\Bigg.\Bigg.$ $\sigma=2$ & $\frac{(a-b)^2\left(a^2+4a\,b+b^2\right)}{(a+b)(a\,b)}$ & & \\
\hline
& & & \\
$\Bigg.\Bigg.$ $\sigma=3$ & $\frac{(a-b)^2(a+b)^2\left(a^2+2a\,b+b^2\right)}{\left(a^2-a\,b+b^2\right)(a\,b)}$ & O. Schn\"urer & \cite{os:surfacesA2} \\
\hline \hline
& & & \\
$\Bigg.\Bigg.$ $\sigma=4$ & $\frac{(a-b)^2(a+b)^6\left(a^2+a\,b+b^2\right)}{(a\,b)^2}$ & O. Schn\"urer & \cite{os:surfacesA2} \\
\hline
& & & \\
$\Bigg.\Bigg.$ $\sigma=5$ & $\frac{(a-b)^2(a+b)^2\left(16(a+b)^8-(a\,b)^4\right)}{(a\,b)^2}$ & & \\
\hline
\end{tabular}
\end{center}
\end{remark}
\begin{lemma}[Quantity $\alpha$ of a vanishing function I]\label{lem:alpha vanishing I}
Let $v$ be a vanishing function for a normal velocity $F$.
Then we have
\begin{align}\label{id:alphavF}
\alpha_{v,F}(a,b) = \frac{b\left(b\,F-a(-a+b)F_a\right)}{a\left(a\,F+b(-a+b)F_b\right)}.
\end{align}
Set $(a,b)=(\rho,1)$, where $0<\rho<1$.
Furthermore, we have
\begin{align}\label{id:alphavFrho}
\alpha_{v,F}(\rho) = \frac{F-\rho(1-\rho)F_a}{\rho\left(\rho\,F+(1-\rho)F_b\right)}.
\end{align}
Note that $\alpha_{v,F}(a,b)$ is unique for all vanishing functions $v$
for a normal velocity $F$.
\end{lemma}
\begin{proof}
Let $v$ be a vanishing function for a normal velocity $F$.
Then we have $C_v(a,b)=0$ for all $0<a,b$.
Using $\alpha_{v,F}:=-\frac{v_a}{v_b}$ yields identity \eqref{id:alphavF}.
Now we set $(a,b)=(\rho,1)$, where $0<\rho<1$, and obtain identity \eqref{id:alphavFrho}.
This concludes the proof.
\end{proof}
\begin{lemma}[Quantity $\alpha$ of a vanishing function II]\label{lem:alpha vanishing II}
Let $v$ be a vanishing function for a normal velocity $F$.
Set $(a,b)=(\rho,1)$, where $0<\rho<1$.
Then we have
\begin{align}\label{id:alphavFbeta}
\alpha_{v,\beta,\sigma}(\rho) = \frac{1+\beta\,\rho(1-(1-\rho)\sigma)}{\rho((1-\rho)\sigma+\rho(\beta\,\rho+1))}
\end{align}
\end{lemma}
\begin{proof}
Let $v$ be a vanishing function for a normal velocity $F$.
Set $(a,b)=(\rho,1)$, where $0<\rho<1$.
Let $C^\alpha_\beta(\rho)$ be defined as in Lemma \ref{lem:necessary}.
For vanishing function $v$, we get $C^\alpha_\beta(\rho)=0$ for all $0<\rho<1$.
This implies identity \eqref{id:alphavFbeta}.
\end{proof}
\begin{lemma}[Quantity $\alpha$ of a vanishing function III]\label{lem:alpha vanishing III}
Let $F^\sigma_\xi$ be defined as in Remark \ref{rem:normal velocity xi}.
Let $v$ be a vanishing function for a contracting normal velocity $F^\sigma_\xi$, i.\,e.\ $\sigma>0$.
Set $(a,b)=(\rho,1)$, where $0<\rho<1$.
Let $\alpha_{v,\beta,\sigma}(\rho)$ be defined as in Lemma \ref{lem:alpha vanishing II}.
Then $\alpha_{v,\beta,\sigma}(\rho)$ is strictly decreasing in $\sigma$
for all $0<\rho<1$.
\end{lemma}
\begin{proof}
Let $\sigma,\,\psi>0$.
Let $\beta=\rho^{\xi-1}$.
We calculate $\alpha_{v,\beta,\sigma}(\rho)$ and $\alpha_{v,\beta,\sigma+\psi}(\rho)$,
and we subtract the second term from the first term.
We obtain
\begin{align*}
0<\frac{(1-\rho)(1+\rho^{\xi})(1+\rho^{1+\xi})\psi}{\rho\left(\rho+\rho^{1+\xi}+(1-\rho)\sigma\right)
\left(\rho+\rho^{1+\xi}+(1-\rho)(\sigma+\psi)\right)}
\end{align*}
for all $0<\rho<1$. This concludes the proof.
\end{proof}
\begin{lemma}[Quantity $\alpha$ of a vanishing function IV]\label{lem:alpha vanishing IV}
Let $F^\sigma_\xi$ be defined as in Remark \ref{rem:normal velocity xi}.
Let $v$ be a vanishing function for a contracting normal velocity $F^\sigma_\xi$ with $\sigma>1$.
Set $(a,b)=(\rho,1)$, where $0<\rho<1$.
Let $\alpha_{v,\beta,\sigma}(\rho)$ be defined as in Lemma \ref{lem:alpha vanishing II}.
Then we have
\begin{align*}
\alpha_{v,\beta,\sigma}(\rho)<\frac{1}{\rho}
\end{align*}
for all $0<\rho<1$.
\end{lemma}
\begin{proof}
By Lemma \ref{lem:alpha vanishing III}, $\alpha_{v,\beta,\sigma}(\rho)$ is strictly decreasing in $\sigma$
for all $0<\rho<1$ and all $\sigma>0$.
We get
\begin{align*}
\alpha_{v,\beta,1}(\rho)=\frac{1}{\rho}
\end{align*}
for all $0<\rho<1$. Therefore, the claim follows.
\end{proof}
\begin{figure}[t]
\vspace*{2mm}
\center\includegraphics[width=8.3cm]{figure1}
\caption{
Let $v_{H^\sigma}$ be vanishing functions as in \eqref{id:MPF Schnulze}, and let $w_{H^\sigma}$ be MPF as defined as in Remark \ref{rem:not vanishing}, which are not vanishing functions.
Set $(a,b)=(\rho,1)$, where $0\leq\rho\leq 1$.
We plot the quantity $\rho\cdot \alpha$ for $v_{H^\sigma}$ with $\sigma=1,\ldots,7$, (dashed lines), and for $w_{H^\sigma}$ with $\sigma=2,\ldots,5$, on the interval $[0,1]$.
Note that we have ${\lim_{\rho\rightarrow 0}\,} \rho\cdot \alpha_{w,H^\sigma} = {\lim_{\rho\rightarrow 0}\,} \rho\cdot \alpha_{v,H^\sigma} = \frac{1}{\sigma}$.
}
\end{figure}
\begin{lemma}[Quantity $\alpha$ of a vanishing function V]\label{lem:alpha vanishing V}
Let $F^\sigma_\xi$ be defined as in Remark \ref{rem:normal velocity xi}.
Let $v$ be a vanishing function for a contracting normal velocity $F^\sigma_\xi$.
Set $(a,b)=(\rho,1)$, where $0<\rho<1$.
Let $\alpha_{v,\beta,\sigma}(\rho)$ be defined as in Lemma \ref{lem:alpha vanishing II}.
Then we have for $\rho\rightarrow 0$
\begin{align*}
\alpha_{v,\beta,\sigma}(\rho) = \frac{1}{\sigma\,\rho} + o\left(\rho^{-1}\right),&\, \text{ if } \xi > 0, \\
\alpha_{v,\beta,\sigma}(\rho) = \frac{2-\sigma}{\sigma\,\rho} + o\left(\rho^{-1}\right),&\, \text{ if } \xi = 0,\,\sigma\neq2, \\
\alpha_{v,\beta,\sigma}(\rho) = 1 + o\left(\rho^{-1}\right),&\, \text{ if } \xi = 0,\,\sigma=2.
\end{align*}
Furthermore, we have
\begin{align*}
\alpha_{v,\beta,\sigma}(1) = 1
\end{align*}
for all $\sigma>0$.
\end{lemma}
\begin{proof}
Let $\alpha_{v,\beta,\sigma}(\rho)$ be defined as in Lemma \ref{lem:alpha vanishing II}.
The claim follows by direct calculations.
\end{proof}
\begin{lemma}[Quantity $\alpha$ of a vanishing function VI]\label{lem:alpha vanishing VI}
Let $v,\,w$ be a vanishing and a maximum-principle function, respectively,
for a contracting normal velocity $F$.
Set $(a,b)=(\rho,1)$, where $0<\rho<1$.
Then we have
\begin{align*}
\alpha_{v,F}(\rho) \leq \alpha_{w,F}(\rho)
\end{align*}
for all $0<\rho<1$.
\end{lemma}
\begin{proof}
Let $w$ be a MPF for a normal velocity $F$. Then we have $C_w(a,b)\leq 0$ for all $0<a,b$.
Set $(a,b)=(\rho,1)$, where $0<\rho<1$. Using $\alpha_{w,F} := -\frac{w_a}{w_b}$ yields
\begin{align*}
\frac{F-\rho(1-\rho)F_a}{\rho\left(\rho\,F+(1-\rho)F_b\right)} \leq \alpha_{w,F}.
\end{align*}
By Lemma \ref{lem:alpha vanishing I}, the left term equals $\alpha_{v,F}$.
Hence, the claim follows.
Note that the denominator of $\alpha_{v,F}$ is strictly positive for all $0<\rho<1$.
\end{proof}
\begin{lemma}[Quantity $\alpha$ of a vanishing function VII]\label{lem:alpha vanishing VII}
Let $v$ be a vanishing function for some $F=H^\sigma$ with $\sigma>1$,
i.\,e.\ $\beta(a,b)=1$ for all $0<a,b$.
Set $(a,b)=(\rho,1)$, where $0<\rho<1$.
Let $\alpha_{v,\beta,\sigma}(\rho)$ be defined as in Lemma \ref{lem:alpha vanishing II}.
Then we have
\begin{align*}
\alpha_{v,1,\sigma}(\rho)=\frac{1+\rho-(1-\rho)\rho\,\sigma}{\rho\left(\rho+\rho^2+(1-\rho)\sigma\right)}
\end{align*}
Furthermore, $\alpha_{v,1,\sigma}(\rho)$ has these roots
\begin{align*}
\rho_{\mp} = \frac{-1+\sigma\mp \sqrt{1-6\,\sigma+\sigma^2}}{2\,\sigma}.
\end{align*}
In particular, we have
\begin{align*}
\rho_{-} \in \left(0,-1+\sqrt{2}\right],
\quad \rho_{+} \in&\, \left[-1+\sqrt{2},1\right),
\quad \text{for all } \sigma\geq 3+\sqrt{2}, \\
\rho_{-} \in \left(0,1/3\right],
\quad \rho_{+} \in&\, \left[1/2,1\right),
\quad \text{for all } \sigma\geq 6,
\end{align*}
where $-1+\sqrt{2}\approx 0.414$, and $3+\sqrt{2} \approx 5.828$.
Note that $\alpha_{v,1,3+\sqrt{3}}(\rho)$
has one root $\rho_{\mp}=-1+\sqrt{2}$, and
$\alpha_{v,1,6}(\rho)$ has two roots $\rho_{-}=1/3$ and $\rho_{+}=1/2$.
\end{lemma}
\begin{proof}
We set $\beta=1$ in $\alpha_{v,\beta,\sigma}(\rho)$ from Lemma \ref{lem:alpha vanishing II}
to obtain $\alpha_{v,1,\sigma}(\rho)$.
Note that the denominator of $\alpha_{v,1,\sigma}(\rho)$ is strictly positive
for all $0<\rho<1$.
Since the numerator is a quadratic equation in $\rho$, we obtain the roots
\begin{align*}
\rho_{\mp} = \frac{-1+\sigma\mp \sqrt{1-6\,\sigma+\sigma^2}}{2\,\sigma}.
\end{align*}
The discriminant is also a quadratic equation, this time in $\sigma$.
Thus, $\alpha_{v,1,\sigma}(\rho)$ has
\begin{itemize}
\item no roots, if $3-2\,\sqrt{2} < \sigma < 3+\sqrt{2}$,
\item one root, if $\sigma=3-2\,\sqrt{2}$ or if $\sigma = 3+\sqrt{2}$,
\item two roots, if $\sigma<3-2\,\sqrt{2}$ or if $3+\sqrt{2}<\sigma$,
\end{itemize}
where $3-2\,\sqrt{2} \approx 0.172$, and $3+\sqrt{2} \approx 5.828$.
We obtain the roots for $\sigma=3+\sqrt{2}$ and for $\sigma=6$ by direct calculations.
By Lemma \ref{lem:alpha vanishing V}, $\alpha_{v,1,\sigma}(\rho)$ is
strictly positive in some zero neighborhood of $\rho$, and
$\alpha_{v,1,\sigma}(1)=1$.
And by Lemma \ref{lem:alpha vanishing III}, $\alpha_{v,1,\sigma}(\rho)$ is
strictly decreasing in $\sigma$ for all $0<\rho<1$.
Therefore, we get $\rho_{-} \in \left(0,-1+\sqrt{2}\right]$ and $\rho_{+}\in[-1+\sqrt{2},1)$
for all $\sigma\geq 3+\sqrt{2}$, and
$\rho_{-} \in \left(0,1/3\right]$ and $\rho_{+} \in \left[1/2,1\right)$
for all $\sigma\geq 6$. This concludes the proof.
\end{proof}
\begin{lemma}[Quantity $\alpha$ of a vanishing function VIII]\label{lem:alpha vanishing VIII}
Let $v$ be a vanishing function for some $F=|A|^\sigma$ with $\sigma>1$,
i.\,e.\ $\beta(a,b)=\frac{a}{b}$ for all $0<a,b$.
Set $(a,b)=(\rho,1)$, where $0<\rho<1$.
Let $\alpha_{v,\beta,\sigma}(\rho)$ be defined as in Lemma \ref{lem:alpha vanishing II}.
Then we have
\begin{align*}
\alpha_{v,\rho,\sigma}(\rho)=\frac{1+\rho^2-(1-\rho)\rho^2\,\sigma}{\rho\left(\rho+\rho^3+(1-\rho)\sigma\right)}
\end{align*}
Furthermore, $\alpha_{v,\rho,\sigma}(\rho)$ has these roots
\begin{align*}
\rho_0 =&\, \frac{1}{3\sigma} \left(
\sigma-1
-\frac{1+\sqrt{3}\,\imath}{2^{2/3}}(\sigma-1)^2\frac{1}{\varrho}
-\frac{1-\sqrt{3}\,\imath}{2\cdot 2^{1/3}}\varrho
\right), {\displaybreak[1]} \\
\rho_{-} =&\, \frac{1}{3\sigma} \left(
\sigma-1
-\frac{1-\sqrt{3}\,\imath}{2^{2/3}}(\sigma-1)^2\frac{1}{\varrho}
-\frac{1+\sqrt{3}\,\imath}{2\cdot 2^{1/3}}\varrho
\right), {\displaybreak[1]} \\
\rho_{+} =&\, \frac{1}{3\sigma} \left(
\sigma-1
+\frac{2}{2^{2/3}}(\sigma-1)^2\frac{1}{\varrho}
+\frac{1}{2^{1/3}}\varrho
\right), {\displaybreak[1]} \\
&\Big. \Big. \qquad \text{where} \\
\varrho =&\, \left( -2+6\sigma-33\sigma^2+2\sigma^3+3\sqrt{3}\sigma\sqrt{4-12\sigma+39\sigma^2-4\sigma^3} \right)^{1/3}.
\end{align*}
In particular, we have
\begin{align*}
\rho_{-} \in \left(0,\rho_\star\right],
\quad \rho_{+} \in&\, \left[\rho_\star,1\right),
\quad \text{for all } \sigma\geq \sigma_\star, {\displaybreak[1]} \\
\rho_{-} \in \left(0,1/2\right],
\quad \rho_{+} \in&\, \left[1/5\left(1+\sqrt{6}\right),1\right),
\quad \text{for all } \sigma\geq 10,
\end{align*}
where $\rho_\star\approx 0.596$,
$\sigma_{\star} := \frac{1}{12} \left( 39+ \left( 51759-5832\sqrt{2}\right)^{1/3} + 8\left( 71+8\sqrt{2} \right)^{1/3} \right) \approx 9.444$,
$1/5\left(1+\sqrt{6}\right) \approx 0.69$.
Note that $\alpha_{v,\rho,\sigma_\star}(\rho)$ has one root $\rho_{\mp}=\rho_\star$,
and $\alpha_{v,\rho,10}(\rho)$ has two roots $\rho_{-}=0.5$, and $\rho_{+}=1/5\left(1+\sqrt{6}\right)$.
Some results were obtained using a compute algebra program.
\end{lemma}
\begin{proof}
We set $\beta=\rho$ in $\alpha_{v,\beta,\sigma}(\rho)$ from Lemma \ref{lem:alpha vanishing II}
to obtain $\alpha_{v,\rho,\sigma}(\rho)$.
Note that the denominator of $\alpha_{v,\rho,\sigma}(\rho)$ is strictly positive
for all $0<\rho<1$.
Since the numerator is a cubic equation in $\rho$, we obtain the three roots
$\rho_0$, $\rho_{-}$, and $\rho_{+}$.
The radicand in $\varrho$ is also a cubic equation, this time in $\sigma$.
Thus, $\alpha_{v,\rho,\sigma}(\rho)$ has
\begin{itemize}
\item no roots, if $1<\sigma<\sigma_{\star}$,
\item one root, if $\sigma=\sigma_{\star}$,
\item two roots, if $\sigma_{\star} < \sigma$.
\end{itemize}
where $\sigma_{\star} := \frac{1}{12} \left( 39+ \left( 51759-5832\sqrt{2}\right)^{1/3} + 8\left( 71+8\sqrt{2} \right)^{1/3} \right) \approx 9.444$.
$\alpha_{v,\rho,\sigma_\star}(\rho)$ has one root $\rho_\star\approx 0.596$ for $0<\rho<1$.
All of the above computations are obtained using a computer algebra program.
We rewrite by hand
\begin{align*}
\alpha_{v,\rho,10}(\rho) = \left(2\rho-1\right)\left(5\rho-\left(1-\sqrt{6}\right)\right)\left(\rho-1/5\left(1+\sqrt{6}\right)\right).
\end{align*}
obtaining the roots
\begin{align*}
\rho_0 =&\, \frac{1}{5}\left( 1-\sqrt{6}\right) \approx -0.29, \\
\rho_{-} =&\, \frac{1}{2} = 0.5, \\
\rho_{+} =&\, \frac{1}{5}\left( 1+\sqrt{6} \right) \approx 0.69.
\end{align*}
By Lemma \ref{lem:alpha vanishing V}, $\alpha_{v,\rho,\sigma}(\rho)$ is
strictly positive in some zero neighborhood of $\rho$, and
$\alpha_{v,\rho,\sigma}(1)=1$.
And by Lemma \ref{lem:alpha vanishing III}, $\alpha_{v,\rho,\sigma}(\rho)$ is
strictly decreasing in $\sigma$ for all $0<\rho<1$.
Therefore, we get $\rho_{-} \in \left(0,\rho_\star\right]$ and $\rho_{+}\in\left[\rho_\star,1\right)$
for all $\sigma\geq \sigma_\star$, and
$\rho_{-} \in \left(0,1/2\right]$ and $\rho_{+} \in \left[1/5\left(1+\sqrt{6}\right),1\right)$
for all $\sigma\geq 10$. This concludes the proof.
\end{proof}
\section{Gauss curvature, $F^\sigma_0$}
In this chapter, we discuss $F^\sigma_0=K^{\sigma/2}$ for all $\sigma\in R\setminus\{0\}$.
For any $\sigma<-2$ and any $\sigma>2$, we get the non-existence of MPF.
This is a direct consequence of the necessary conditions in Theorem \ref{lem:necessary}.
For any $0<\sigma\leq 1$, we also get the non-existence of MPF.
This result is implied by Lemma \ref{lem:one}, and by Lemma \ref{lem:zeroone}.
For all other powers $\sigma$, we state MPF by
B. Andrews \cite{ba:gauss}, B. Andrews and X. Chen \cite{ac:surfaces}, Q. Li \cite{ql:surfaces},
and O. Schn\"urer \cite{os:surfaces}.
We make a summary of all results on $F^\sigma_0=\sgn(\sigma)\cdot K^{\sigma/2}$
in Corollary \ref{cor:gauss curvature}.
We should point out that C. Gerhardt \cite{cg:non} proves convergence to a sphere at $\infty$
for all $\sigma<0$ despite the fact that there exist no MPF for all $\sigma<-2$.
We discuss this further in chapter 'Outlook'.
\begin{theorem}[Gauss curvature I]\label{thm:gauss curvature I}
Let $F=F^\sigma_0=\sgn(\sigma)\cdot K^{\sigma/2}$.
Then there are no $MPF$, if $\sigma<-2$.
\end{theorem}
\begin{proof}
Let $F^\sigma_\xi$ be defined as in Remark $\ref{rem:normal velocity xi}$.
For powers of the Gauss curvature, we get
$F^\sigma_0=\sgn(\sigma)\cdot K^{\sigma/2}$.
By Theorem \ref{thm:necessary}, we have the necessary condition $\sigma\in[-2,0)$
for the existence of MPF in case of expanding normal velocities, i.\,e.\ $\sigma<0$.
Hence, the claim follows.
\end{proof}
\begin{theorem}[Gauss curvature II]\label{thm:gauss curvature II}
Let $F=F^\sigma_0=\sgn(\sigma)\cdot K^{\sigma/2}$.
Then there are no $MPF$, if $0<\sigma\leq 1$.
\end{theorem}
\begin{proof}
This is a direct consequence of Lemma \ref{lem:one} and Lemma \ref{lem:zeroone}.
\end{proof}
\begin{theorem}[Gauss curvature III]\label{thm:gauss curvature III}
Let $F=F^\sigma_0=\sgn(\sigma)\cdot K^{\sigma/2}$.
Then there are no $MPF$, if $\sigma>2$.
\end{theorem}
\begin{proof}
Let $F^\sigma_\xi$ be defined as in Remark $\ref{rem:normal velocity xi}$.
For powers of the Gauss curvature, we get
$F^\sigma_0=\sgn(\sigma)\cdot K^{\sigma/2}$.
By Theorem \ref{thm:necessary}, we have the necessary condition $\sigma\in(1,0]$
for the existence of MPF in case of contracting normal velocities, i.\,e.\ $\sigma>0$.
Hence, the claim follows.
\end{proof}
\newpage
\begin{corollary}[Gauss curvature]\label{cor:gauss curvature}
Let $F=F^\sigma_0=\sgn(\sigma)\cdot K^{\sigma/2}$.
Then we have
\begin{center}
\begin{tabular}{| l | c || c | c |}
\hline
& MPF & & \\
\hline \hline
& & & \\
$\sigma>2$ & non-existent & & Theorem \ref{thm:gauss curvature III} \\
\hline
& & & \\
$\sigma=2$ & $(a-b)^2$ & B. Andrews & \cite{ba:gauss} \\
\hline
& & & \\
$\sigma\in(1,2)$ & $\frac{(a-b)^2(a\,b)^\sigma}{(a\,b)^2}$ & B. Andrews, X. Chen & \cite{ac:surfaces} \\
\hline
& & & \\
$\sigma\in(0,1]$ & non-existent & & Theorem \ref{thm:gauss curvature II} \\
\hline \hline
& & & \\
$\sigma\in(-2,0)$ & $\frac{(a-b)^2(a\,b)^{\sigma/2}}{(a\,b)}$ & Q. Li & \cite{ql:surfaces} \\
\hline
& & & \\
$\sigma=-2$ & $\frac{(a-b)^2}{(a\,b)^2}$ & O. Schn\"urer & \cite{os:surfaces} \\
\hline
& & & \\
$\sigma<-2$ & non-existent & & Theorem \ref{thm:gauss curvature I} \\
\hline
\end{tabular}
\end{center}
\end{corollary}
\section{Mean curvature, $F^\sigma_1$}
In this chapter, we discuss $F^\sigma_1=\sgn(\sigma)\cdot H^\sigma$
for all $\sigma\in{\mathbb{R}}\setminus\{0\}$.
For $\sigma\geq 5.98$, we show the non-existence of MPF.
This is Theorem \ref{thm:mean curvature III}.
We show this by comparing an assumed MPF to a vanishing function.
Since vanishing functions develop roots for any
$\sigma\geq 3+\sqrt{2} \approx 5.828$,
we obtain a contradiction.
Note that the quantity $\alpha$ has to be strictly positive
for a MPF, see Lemma \ref{lem:alpha}.
Our proof technique does not work for $5.17<\sigma\leq 5.98$
since the quantity $\alpha$ of vanishing functions
does not develop any roots there.
For $5.89\leq\sigma<5.98$, $\alpha$ has roots but
our technique still does not work, since it involves
some estimates which apparently are not sharp.
For any $1<\sigma\leq 5.17$, we have a MPF by
O. Schn\"urer and F. Schulze \cite{fs:convexity},
which is also a vanishing function.
By Theorem \ref{thm:mean curvature II}, we get
the non-existence of MPF for any $0<\sigma\leq 1$.
In Lemma \ref{lem:H expanding}, we give a MPF,
which we present fully in a separate paper.
This is $-1<\sigma<0$.
And for $\sigma=-1$, we have another MPF by O. Schn\"urer.
For any $\sigma<-1$, the non-existence of MPF is shown
in Theorem \ref{thm:mean curvature I}.
We prove this using the necessary conditions for $F^\sigma_\xi$,
which we obtained in Theorem \ref{thm:necessary}.
We summarize our results for $F^\sigma_1=\sgn(\sigma)\cdot H^\sigma$
in the table of Corollary \ref{cor:mean curvature}.
\begin{theorem}[Mean curvature I]\label{thm:mean curvature I}
Let $F=F^\sigma_1=\sgn(\sigma)\cdot H^\sigma$.
Then there are no $MPF$, if $\sigma<-1$.
\end{theorem}
\begin{proof}
Let $F^\sigma_\xi$ be defined as in Remark $\ref{rem:normal velocity xi}$.
For powers of the mean curvature, we get
$F^\sigma_1=\sgn(\sigma)\cdot H^\sigma$.
By Theorem \ref{thm:necessary}, we have the necessary condition $\sigma\in[-1,0)$
for the existence of MPF in case of expanding normal velocities, i.\,e.\ $\sigma<0$.
Hence, the claim follows.
\end{proof}
\begin{theorem}[Mean curvature II]\label{thm:mean curvature II}
Let $F=F^\sigma_1=\sgn(\sigma)\cdot H^\sigma$.
Then there are no $MPF$, if if $0<\sigma\leq 1$.
\end{theorem}
\begin{proof}
This is a direct consequence of Lemma \ref{lem:one} and Lemma \ref{lem:zeroone}.
\end{proof}
\begin{theorem}[Mean curvature III]\label{thm:mean curvature III}
Let $F=F^\sigma_1=\sgn(\sigma)\cdot H^\sigma$.
Then there are no $MPF$, if $\sigma\geq \sigma_\delta$ with $\sigma_\delta=7$.
Using a computer algebra system, we even have $\sigma_\delta\approx 5.98$.
\end{theorem}
\begin{proof}
Let $\sigma_\delta := \sigma_0 + \delta$ for some $\sigma_0>1$ and some $\delta>0$.
Suppose there exists a MPF $w$ for any normal velocity $F=H^{\sigma_\delta}$.
Set $(a,b)=(\rho,1)$, where $0<\rho<1$.
We recall the gradient terms $G^\alpha_\beta(\rho)$ from Lemma \ref{lem:necessary}.
For $F=F^{\sigma_\delta}_1=\sgn(\sigma_\delta)\cdot H^{\sigma_\delta}$, we have $\beta=1$, and $\beta_a=0$.
By Lemma \ref{lem:necessary}, $G^\alpha_1(\rho)$ is non-positive for all $0<\rho<1$.
\begin{align*}
G^\alpha_1(\rho) =&\,-\alpha_a(\alpha\,\rho-1)(\beta\,\rho+1)(1-\rho) \\
&\,\quad+\beta_a(\alpha\,\rho-1)^2(1-\rho) \\
&\,\quad-(\alpha+\beta)(\alpha(1-\rho)(1-\sigma_0-\delta)
+\beta(1+\rho-(1-\rho)(\sigma_0+\delta))+2) {\displaybreak[1]} \\
=&\,-\alpha_a(\alpha\,\rho-1)(\rho+1)(1-\rho) \\
&\,\quad-(\alpha+\rho)(\alpha(1-\rho)(1-\sigma_0-\delta)
+(1+\rho-(1-\rho)(\sigma_0+\delta))+2)
\end{align*}
Since $w$ is MPF, using Theorem \ref{thm:necessary}, we get for $\rho\rightarrow 0$
\begin{align*}
\alpha_{w,1,\sigma_\delta}(\rho) =&\, \frac{c}{\rho} + o\left(\rho^{-1}\right) \\
=&\, \frac{1}{\sigma_\delta\,\rho} + o\left(\rho^{-1}\right).
\end{align*}
Now, let $v$ be a vanishing function for $F=H^{\sigma_0}$.
By Lemma \ref{lem:alpha vanishing V}, we get for $\rho\rightarrow 0$
\begin{align*}
\alpha_{v,1,\sigma_0}(\rho) = \frac{1}{\sigma_0\,\rho} + o\left(\rho^{-1}\right).
\end{align*}
By Lemma \ref{lem:alpha vanishing VII}, $\alpha_{v,1,\sigma_0}(\rho)$ has one root $\rho_0\in(0,-1+\sqrt{2}]$,
if $\sigma_0\geq 3+2\sqrt{2}$.
Combining the results on $\alpha_{w,1,\sigma_\delta}(\rho)$ and $\alpha_{v,1,\sigma_0}(\rho)$
for $\rho\rightarrow 0$ yields the existence of some zero neighborhood of $\rho$,
where
\begin{align}\label{id:alpha v w}
\alpha_{v,1,\sigma_0}(\rho) \geq \alpha_{w,1,\sigma_\delta}(\rho)>0.
\end{align}
We get the last inequality by Lemma \ref{lem:alpha}.
Thus, there exists a first boundary/intersection point $\rho_1\in(0,\rho_0)$
of $\alpha_{w,1,\sigma_\delta}(\rho)$ and $\alpha_{v,1,\sigma_0}(\rho)$,
such that
\begin{align}
\begin{split}\label{id:replace alpha}
\alpha_{w,1,\sigma_\delta}(\rho_1) =&\, \alpha_{v,1,\sigma_0}(\rho_1), \\
\alpha_{a\,(w,1,\sigma_\delta)}(\rho_1) =&\, \alpha_{a\,(v,1,\sigma_0)}(\rho_1) + {\varepsilon},
\end{split}
\end{align}
for some ${\varepsilon}\geq 0$.
Otherwise, we have $\alpha_{w,1,\sigma_\delta}(\rho_2)=0$ for some $\rho_2\in(0,\rho_0)$,
which results in a contradiction to inequality \eqref{id:alpha v w}.
In the following, we calculate the gradient terms $G^\alpha_1(\rho)$
at a first boundary/ intersection point $\rho_1\in(0,\rho_0)$.
As a preliminary step, we replace $\alpha_a$ by $\alpha_a+{\varepsilon}$.
This yields
\begin{align*}
G^\alpha_1(\rho_1) =&\,-(\alpha_a+{\varepsilon})(\alpha\,\rho_1-1)(\rho_1+1)(1-\rho_1) \\
&\,\quad-(\alpha+\rho_1)(\alpha(1-\rho_1)(1-\sigma_0-\delta)
+(1+\rho_1-(1-\rho_1)(\sigma_0+\delta))+2).
\end{align*}
By Lemma \ref{lem:alpha vanishing IV}, we have
\begin{align*}
\alpha_{v,1,\sigma_0}(\rho) \leq \frac{1}{\rho}
\end{align*}
for all $0<\rho<1$. Applying this to $G^\alpha_1(\rho_1)\leq 0$ yields
\begin{align*}
0\leq&\, {\varepsilon} \\
\leq&\,\frac{-\alpha_a(1-\alpha\,\rho_1)\left(1-\rho_1^2\right)-(\alpha+1)( \alpha(1-\rho_1)(\sigma_\delta-1)+(1-\rho_1)(\sigma_\delta+1)-4 )}{(1-\alpha\,\rho_1)\left(1-\rho_1^2\right)} \\
=:&\, \Phi(\rho_1,\sigma_0,\delta).
\end{align*}
Therefore, the term $\Phi(\rho_1,\sigma_0,\delta)$ is non-negative at a first boundary/intersection point $\rho_1\in (0,\rho_0)$.
So if $\Phi(\rho_1,\sigma_0,\delta)$ is strictly negative for all $\rho_1\in(0,\rho_0)$,
no boundary/ intersection point $\rho_1$ can occur.
By Lemma \ref{lem:alpha vanishing VII}, we have
\begin{align*}
\alpha_{v,1,\sigma_0}(\rho)=\frac{1+\rho-(1-\rho)\rho\,\sigma_0}{\rho\left(\rho+\rho^2+(1-\rho)\sigma_0\right)}.
\end{align*}
Differentiating $\alpha_{v,1,\sigma_0}(\rho)$ with respect to $\rho$ yields
\begin{align*}
\alpha_{a\,(v,1,\sigma_0)}(\rho) = \frac{-\sigma_0+2\rho(\sigma_0-1)+2\rho^2(\sigma_0-2)+2\rho^3(\sigma_0-1)-\rho^4\sigma_0}{\rho^2\left(\rho+\rho^2+(1-\rho)\sigma_0\right)^2}.
\end{align*}
Here, we replace $\alpha$ by $\alpha_{v,1,\sigma_0}(\rho_1)$ and $\alpha_a$ by $\alpha_{a\,(v,1,\sigma_0)}(\rho_1)$ in $\Phi(\rho_1,\sigma_0,\delta)$.
We can do so, since we assume the existence of a first boundary/intersection point $\rho_1\in(0,\rho_0)$, see \eqref{id:replace alpha}.
We get
\begin{align*}
\Phi(\rho_1,\sigma_0,\delta) = \frac{\delta\cdot\Phi_1(\rho_1,\sigma_0) + \Phi_2(\rho_1,\sigma_0)}{\Phi_3(\rho_1,\sigma_0)},
\end{align*}
where
\begin{align*}
\Phi_1(\rho_1,\sigma_0) =&\, -(1-\rho_1)\left(1+\rho_1^2\right)^2\left(\rho_1+\rho_1^2+(1-\rho_1)\sigma_0\right), \\
\Phi_2(\rho_1,\sigma_0) =&\, \rho_1(3\sigma_0-1)
-4\rho_1^2\sigma_0(\sigma_0-1)
+\rho_1^3(\sigma_0(8\sigma_0-9)+7) {\displaybreak[1]} \\
&\quad -\rho_1^4(\sigma_0^2-1)
+\rho_1^5(5\sigma_0+1)
-4\rho_1^6(\sigma_0-1)
+\rho_1^7(\sigma_0+1) \\
\Phi_3(\rho_1,\sigma_0) =&\, \rho_1^2(1-\rho_1)^2(\sigma_0-1)\left(\rho_1+\rho_1^2+(1-\rho_1)\sigma_0\right)^2,
\end{align*}
Clearly, the term $\Phi_1(\rho_1,\sigma_0)$ is strictly negative and the term $\Phi_3(\rho_1,\sigma_0)$ is strictly positive for all $0<\rho_1<1$,
and for all $\sigma_0>1$.
Recall that by Lemma \ref{lem:alpha vanishing VII}, $\alpha_{v,1,\sigma_0}(\rho)$ has one root $\rho_0\in(0,-1+\sqrt{2}]$,
if $\sigma_0\geq 3+2\sqrt{2}$.
As defined in the beginning of this proof we have, $\sigma_\delta := \sigma_0 + \delta$.
In the last two paragraphs of this proof, we show that
$\delta\cdot\Phi_1(\rho_1,\sigma_0) + \Phi_2(\rho_1,\sigma_0)$ is strictly negative for some fixed $\sigma_0>1$,
and some fixed $\delta>0$ for all $\rho_1\in(0,\rho_0)$.
As noted before, $\Phi_1(\rho_1,\sigma_0)$ is strictly negative for all $0<\rho_1<1$. Therefore, we have
$\delta\cdot\Phi_1(\rho_1,\sigma_0) + \Phi_2(\rho_1,\sigma_0)\geq \delta_1\cdot\Phi_1(\rho_1,\sigma_0) + \Phi_2(\rho_1,\sigma_0)$,
if $\delta\leq\delta_1$, for all $\rho_1\in(0,\rho_0)$.
So no boundary/intersection point can occur for any $\sigma_0+\delta_1\geq\sigma_0+\delta$.
Hence, no MPF $w$ exists for any $\sigma\geq\sigma_\delta$.
Using a computer algebra system, we compute $\delta\cdot\Phi_1(\rho_1,\sigma_0) + \Phi_2(\rho_1,\sigma_0)$
for $\sigma_0 = 3+2\sqrt{2}$, and $\delta=0.151$, i.\,e.\ $\sigma_\delta=\sigma_0+\delta\approx 5.98$.
We get that $\delta\cdot\Phi_1(\rho_1,\sigma_0) + \Phi_2(\rho_1,\sigma_0)$ is strictly negative for all
$0<\rho_1\leq-1+\sqrt{2}$.
By Lemma \ref{lem:alpha vanishing VII}, $\alpha_{v,1,\sigma_0}(\rho)$ has one root $\rho_0\in(0,-1+\sqrt{2}]$,
if $\sigma_0\geq 3+2\sqrt{2}$.
Hence, there are no MPF $w$ for any $F=H^\sigma$ with $\sigma\geq 5.98$.
By Lemma \ref{lem:tedious H}, we calculate $\delta\cdot\Phi_1(\rho_1,\sigma_0) + \Phi_2(\rho_1,\sigma_0)$
for $\sigma_0=6$, and $\delta=1$.
We get that $\delta\cdot\Phi_1(\rho_1,\sigma_0) + \Phi_2(\rho_1,\sigma_0)$ is strictly negative for all
$0<\rho_1\leq 1/3$.
By, Lemma \ref{lem:alpha vanishing VII}, $\alpha_{v,1,\sigma}(\rho)$ has one root $\rho_0\in(0,1/3]$ for all $\sigma\geq 6$.
Hence, there are no MPF for any $F=H^\sigma$ with $\sigma\geq 7$.
This concludes the proof.
\end{proof}
\begin{lemma}[Calculations for Theorem \ref{thm:mean curvature III}]\label{lem:tedious H}
Define
\begin{align*}
\tilde{\Phi}(\rho,\sigma,\delta) := \delta\cdot\Phi_1(\rho,\sigma) + \Phi_2(\rho,\sigma),
\end{align*}
where
\begin{align*}
\Phi_1(\rho,\sigma) =&\, -(1-\rho)\left(1+\rho^2\right)^2\left(\rho+\rho^2+(1-\rho)\sigma\right), \\
\Phi_2(\rho,\sigma) =&\, \rho(3\sigma-1)
-4\rho^2\sigma(\sigma-1)
+\rho^3(\sigma(8\sigma-9)+7) \\
&\quad -\rho^4(\sigma^2-1)
+\rho^5(5\sigma+1)
-4\rho^6(\sigma-1)
+\rho^7(\sigma+1),
\end{align*}
as in Theorem \ref{thm:mean curvature III}.
Let $\sigma=6$, and $\delta=1$.
Then we have that $\tilde{\Phi}(\rho,6,1)$ is strictly negative for all $0<\rho\leq 1/3$.
\end{lemma}
\begin{proof}
First, we calculate $\tilde{\Phi}(\rho,6,1)$. We get
\begin{align*}
\tilde{\Phi}(\rho,6,1) =&\, -6+28\rho-138\rho^2+264\rho^3-158\rho^4+44\rho^5-26\rho^6+8\rho^7.
\end{align*}
Next, we rewrite $\tilde{\Phi}(\rho,6,1)$ as a sum of four polynomials
\begin{align*}
\tilde{\Phi}(\rho,6,1) = \tilde{\Phi}_1(\rho) + \tilde{\Phi}_2(\rho) + \tilde{\Phi}_3(\rho) + \tilde{\Phi}_4(\rho),
\end{align*}
where
\begin{align*}
\tilde{\Phi}_1(\rho) =&\, -6+28\rho-38\rho^2,\\
\tilde{\Phi}_2(\rho) =&\, -100\rho^2+264\rho^3,\\
\tilde{\Phi}_3(\rho) =&\, -158\rho^4+44\rho^5,\\
\tilde{\Phi}_4(\rho) =&\, -26\rho^6+8\rho^7.
\end{align*}
Now, we show that $\tilde{\Phi}_1(\rho)$ is strictly negative, and
$\tilde{\Phi}_i(\rho),\,i=2,3,4$, is non-positive for all $0<\rho\leq 1/3$.
\begin{itemize}
\item $\tilde{\Phi}_1(\rho)$ has no roots, and $\tilde{\Phi}_1(0)=-6$,
\item $\tilde{\Phi}_2(\rho)/\rho^2$ has one root at $\rho=22/66\approx 0.379$, and $\tilde{\Phi}_2(1/6)=-14/9$,
\item $\tilde{\Phi}_3(\rho)/\rho^4$ has one root at $\rho=79/22\approx 3.591$, and $\tilde{\Phi}_3(1/6)=-113/972$,
\item $\tilde{\Phi}_4(\rho)/\rho^6$ has one root at $\rho=13/4=3.25$, and $\tilde{\Phi}_4(1/6)=-37/69984$.
\end{itemize}
This concludes the proof.
\end{proof}
\begin{lemma}[Mean curvature]\label{lem:H expanding}
Let $F=F^\sigma_1=\sgn(\sigma)\cdot H^\sigma$. Then
\begin{align*}
w=\frac{(a-b)^2(a^2+b^2)(a\,b)^{\sigma-2}}{(a+b)}
\end{align*}
is a maximum-principle function, if $\sigma\in(-1,0)$.
\end{lemma}
\begin{proof}
It is rather easy to see that $w$ is a MPF except for MPF condition \eqref{IV}.
Here, we need a computer algebra program.
We compute $C(a,b)$, $E(a,b)$, $G(a,b)$ for the given $w$ and $F$.
Then we use a Monte-Carlo method to check for non-positivity of these terms.
Since $C(a,b)$, $E(a,b)$, $G(a,b)$ are homogeneous, $C(a,b)=C(b,a)$, and $E(a,b)=G(b,a)$,
it suffices to check for non-positivity testing only on $(a,b)=(\rho,1)$, where $0<\rho\leq 1$.
In a separate paper, using this MPF we will show convergence to a sphere at $\infty$.
\end{proof}
\newpage
\begin{corollary}[Mean curvature]\label{cor:mean curvature}
Let $F=F^\sigma_1=\sgn(\sigma)\cdot H^\sigma$.
Then we have
\begin{center}
\begin{tabular}{| l | c || c | c |}
\hline
& MPF & & \\
\hline \hline
$\Bigg.\Bigg. \sigma\geq 5.
\begin{small}
\text{98}
\end{small}$ & non-existent & & Theorem \ref{thm:mean curvature III} \\
\hline
$\Bigg.\Bigg. \sigma\in(5.
\begin{small}
\text{17}
\end{small}
,5.
\begin{small}
\text{98}
\end{small}
)$ & open problem & & \\
\hline
$\Bigg.\Bigg. \sigma\in(1,5.
\begin{small}
\text{17}
\end{small}
]$ & $\frac{\left(a-b\right)^2\left(a+b\right)^{2\sigma}}{\left(a\,b\right)^2}$ & F. Schulze, O. Schn\"urer & \cite{fs:convexity} \\
\hline
$\Bigg.\Bigg. \sigma\in(0,1]$ & non-existent & & Theorem \ref{thm:mean curvature II} \\
\hline \hline
$\Bigg.\Bigg. \sigma\in(-1,0)$ & $\frac{\left(a-b\right)^2\left(a^2+b^2\right)\left(a\,b\right)^\sigma}{\left(a+b\right)\left(a\,b\right)^2}$ & & Lemma \ref{lem:H expanding} \\
\hline
$\Bigg.\Bigg. \sigma=-1$ & $\frac{\left(a-b\right)^2}{\left(a+b\right)\left(a\,b\right)}$ & O. Schn\"urer & \cite{os:surfaces} \\
\hline
$\Bigg.\Bigg. \sigma<-1$ & non-existent & & Theorem \ref{thm:mean curvature I} \\
\hline
\end{tabular}
\end{center}
\end{corollary}
\section{Norm of the second fundamental form, $F^\sigma_2$}
In this chapter, we discuss $F^\sigma_2=\sgn(\sigma)\cdot |A|^\sigma$
for all $\sigma\in{\mathbb{R}}\setminus\{0\}$.
For $\sigma\geq 9.89$, we show the non-existence of MPF.
This is Theorem \ref{thm:|A| III}.
We show this by comparing an assumed MPF to a vanishing function.
Since vanishing functions develop roots for any
$\sigma\geq 9.444$,
we obtain a contradiction.
Note that the quantity $\alpha$ has to be strictly positive
for a MPF, see Lemma \ref{lem:alpha}.
Our proof technique does not work for $8.15<\sigma<9.444$
since the quantity $\alpha$ of vanishing functions
does not develop any roots there.
For $9.444\leq\sigma<9.89$, $\alpha$ has roots but
our technique still does not work, since it involves
some estimates, which apparently are not sharp.
For any $1<\sigma\leq 8.15$, we have a MPF by
B. Andrews and X. Chen \cite{ba:contraction},
which is also a vanishing function.
By Theorem \ref{thm:|A| II}, we get
the non-existence of MPF for any $0<\sigma\leq 1$.
In Lemma \ref{lem:|A| expanding}, we give a MPF,
which we present fully in a separate paper.
This is $-1\leq\sigma<0$.
And for any $\sigma<-1$, the non-existence of MPF is shown
in Theorem \ref{thm:|A| I}.
We prove this using the necessary conditions for $F^\sigma_\xi$,
which we obtained in Theorem \ref{thm:necessary}.
We summarize our results for $F^\sigma_2=\sgn(\sigma)\cdot |A|^\sigma$
in the table of Corollary \ref{cor:|A|}.
\begin{theorem}[Norm of the second fundamental form I]\label{thm:|A| I}
Let $F=F^\sigma_2=\sgn(\sigma)\cdot |A|^\sigma$.
Then there are no $MPF$, if $\sigma<-1$.
\end{theorem}
\begin{proof}
Let $F^\sigma_\xi$ be defined as in Remark \ref{rem:normal velocity xi}.
For the norm of the second fundamental form, we get
$F^\sigma_2=\sgn(\sigma)\cdot |A|^\sigma$.
By Theorem \ref{thm:necessary}, we have the necessary condition $\sigma\in[-1,0)$
for the existence of MPF in case of expanding normal velocities, i.\,e.\ $\sigma<0$.
Hence, the claim follows.
\end{proof}
\begin{theorem}[Norm of the second fundamental form II]\label{thm:|A| II}
Let $F=F^\sigma_2=\sgn(\sigma)\cdot |A|^\sigma$.
Then there are no $MPF$, if $0<\sigma\leq 1$.
\end{theorem}
\begin{proof}
This is a direct consequence of Lemma \ref{lem:one} and Lemma \ref{lem:zeroone}.
\end{proof}
\begin{theorem}[Norm of the second fundamental form III]\label{thm:|A| III}
Let $F=F^\sigma_2=\sgn(\sigma)\cdot |A|^\sigma$.
Then there are no $MPF$, if $\sigma\geq \sigma_\delta$ with $\sigma_\delta=11$.
Using a computer algebra system, we even have $\sigma_\delta\approx 9.899$.
\end{theorem}
\begin{proof}
Let $\sigma_\delta := \sigma_0 + \delta$ for some $\sigma_0>1$ and some $\delta>0$.
Suppose there exists a MPF $w$ for any normal velocity $F=|A|^{\sigma_\delta}$.
Set $(a,b)=(\rho,1)$, where $0<\rho<1$.
We recall the gradient terms $G^\alpha_\beta(\rho)$ from Lemma \ref{lem:necessary}.
For $F=F^{\sigma_\delta}_2=\sgn(\sigma_\delta)\cdot |A|^{\sigma_\delta}$, we have $\beta=\rho$, and $\beta_a=1$.
By Lemma \ref{lem:necessary}, $G^\alpha_\rho(\rho)$ is non-positive for all $0<\rho<1$.
\begin{align*}
G^\alpha_\rho(\rho) =&\,-\alpha_a(\alpha\,\rho-1)(\beta\,\rho+1)(1-\rho) \\
&\,\quad+\beta_a(\alpha\,\rho-1)^2(1-\rho) \\
&\,\quad-(\alpha+\beta)(\alpha(1-\rho)(1-\sigma_0-\delta)
+\beta(1+\rho-(1-\rho)(\sigma_0+\delta))+2) {\displaybreak[1]} \\
=&\,-\alpha_a(\alpha\,\rho-1)\left(\rho^2+1\right)(1-\rho) \\
&\,\quad+1\cdot(\alpha\,\rho-1)^2(1-\rho) \\
&\,\quad-(\alpha+\rho)(\alpha(1-\rho)(1-\sigma_0-\delta)
+\rho(1+\rho-(1-\rho)(\sigma_0+\delta))+2).
\end{align*}
Since $w$ is MPF, using Theorem \ref{thm:necessary}, we get for $\rho\rightarrow 0$
\begin{align*}
\alpha_{w,\rho,\sigma_\delta}(\rho) =&\, \frac{c}{\rho} + o\left(\rho^{-1}\right) \\
=&\, \frac{1}{\sigma_\delta\,\rho} + o\left(\rho^{-1}\right).
\end{align*}
Now, let $v$ be a vanishing function for $F=|A|^{\sigma_0}$.
By Lemma \ref{lem:alpha vanishing V}, we get for $\rho\rightarrow 0$
\begin{align*}
\alpha_{v,\rho,\sigma_0}(\rho) = \frac{1}{\sigma_0\,\rho} + o\left(\rho^{-1}\right).
\end{align*}
By Lemma \ref{lem:alpha vanishing VIII}, $\alpha_{v,\rho,\sigma_0}(\rho)$ has one root $\rho_0\in(0,\rho_\star]$,
if $\sigma_0\geq \sigma_\star$, where $\rho_\star \approx 0.596$, $\sigma_\star\approx 9.444$.
Combining the results on $\alpha_{w,\rho,\sigma_\delta}(\rho)$ and $\alpha_{v,\rho,\sigma_0}(\rho)$
for $\rho\rightarrow 0$ yields the existence of some zero neighborhood of $\rho$,
where
\begin{align}\label{id:alpha v w A}
\alpha_{v,\rho,\sigma_0}(\rho) \geq \alpha_{w,\rho,\sigma_\delta}(\rho)>0.
\end{align}
We get the last inequality by Lemma \ref{lem:alpha}.
Thus, there exists a first boundary/intersection point $\rho_1\in(0,\rho_0)$
of $\alpha_{w,\rho,\sigma_\delta}(\rho)$ and $\alpha_{v,\rho,\sigma_0}(\rho)$,
such that
\begin{align}
\begin{split}\label{id:replace alpha A}
\alpha_{w,\rho,\sigma_\delta}(\rho_1) =&\, \alpha_{v,\rho,\sigma_0}(\rho_1), \\
\alpha_{a\,(w,\rho,\sigma_\delta)}(\rho_1) =&\, \alpha_{a\,(v,\rho,\sigma_0)}(\rho_1) + {\varepsilon},
\end{split}
\end{align}
for some ${\varepsilon}\geq 0$.
Otherwise, we have $\alpha_{w,\rho,\sigma_\delta}(\rho_2)=0$ for some $\rho_2\in(0,\rho_0)$,
which results in a contradiction to inequality \eqref{id:alpha v w A}.
In the following, we calculate the gradient terms $G^\alpha_1(\rho)$
at a first boundary/ intersection point $\rho_1\in(0,\rho_0)$.
As a preliminary step, we replace $\alpha_a$ by $\alpha_a+{\varepsilon}$.
This yields
\begin{align*}
G^\alpha_\rho(\rho_1) =&\,-(\alpha_a+{\varepsilon})(\alpha\,\rho-1)\left(\rho^2+1\right)(1-\rho) \\
&\,\quad+(\alpha\,\rho-1)^2(1-\rho) \\
&\,\quad-(\alpha+\rho)(\alpha(1-\rho)(1-\sigma_0-\delta)
+\rho(1+\rho-(1-\rho)(\sigma_0+\delta))+2).
\end{align*}
By Lemma \ref{lem:alpha vanishing IV}, we have
\begin{align*}
\alpha_{v,\rho,\sigma_0}(\rho) \leq \frac{1}{\rho}
\end{align*}
for all $0<\rho<1$. Applying this to $G^\alpha_\rho(\rho_1)\leq 0$ yields
\begin{align*}
0\leq&\, {\varepsilon} \\
\leq&\,\frac{-\alpha_a(1-\alpha\,\rho_1)(1-\rho_1)\left(1+\rho_1^2\right)}{(1-\alpha\,\rho_1)(1-\rho_1)\left(1+\rho_1^2\right)} \\
&\quad+ \frac{-(\alpha+\rho_1)(\alpha(1-\rho_1)(\sigma_\delta-1)+\rho_1(1-\rho_1)(\sigma_\delta+1)-2-2\rho_1}{(1-\alpha\,\rho_1)(1-\rho_1)\left(1+\rho_1^2\right)}\\
&\quad+ \frac{-(1-\rho_1)\left(\alpha\,\rho_1-1\right)^2}{(1-\alpha\,\rho_1)(1-\rho_1)\left(1+\rho_1^2\right)}\\
=:&\, \Phi(\rho_1,\sigma_0,\delta).
\end{align*}
Therefore, the term $\Phi(\rho_1,\sigma_0,\delta)$ is non-negative at a first boundary/intersection point $\rho_1\in (0,\rho_0)$.
So if $\Phi(\rho_1,\sigma_0,\delta)$ is strictly negative for all $\rho_1\in(0,\rho_0)$,
no boundary/ intersection point $\rho_1$ can occur.
By Lemma \ref{lem:alpha vanishing VIII}, we have
\begin{align*}
\alpha_{v,\rho,\sigma_0}(\rho)=\frac{1+\rho^2-(1-\rho)\rho^2\,\sigma_0}{\rho\left(\rho+\rho^3+(1-\rho)\sigma_0\right)}.
\end{align*}
Differentiating $\alpha_{v,\rho,\sigma_0}(\rho)$ with respect to $\rho$ yields
\begin{align*}
\alpha_{a\,(v,\rho,\sigma_0)}(\rho) =&\,
\frac{-\sigma_0+2\rho(\sigma_0-1)-\rho^2\sigma_0(\sigma_0-1)+2\rho^3\left(\sigma_0^2-2\right)}
{\rho^2\left(\rho+\rho^3+(1-\rho)\sigma_0\right)^2} \\
&\quad +\frac{-\rho^4\sigma_0(\sigma_0-1)+2\rho^5(\sigma_0-1)-\rho^6\sigma_0}
{\rho^2\left(\rho+\rho^3+(1-\rho)\sigma_0\right)^2}.
\end{align*}
Here, we replace $\alpha$ by $\alpha_{v,\rho,\sigma_0}(\rho_1)$ and $\alpha_a$ by $\alpha_{a\,(v,\rho,\sigma_0)}(\rho_1)$ in $\Phi(\rho_1,\sigma_0,\delta)$.
We can do so, since we assume the existence of a first boundary/intersection point $\rho_1\in(0,\rho_0)$, see \eqref{id:replace alpha A}.
We get
\begin{align*}
\Phi(\rho_1,\sigma_0,\delta) = \frac{\delta\cdot\Phi_1(\rho_1,\sigma_0) + \Phi_2(\rho_1,\sigma_0)}{\Phi_3(\rho_1,\sigma_0)},
\end{align*}
where
\begin{align*}
\Phi_1(\rho_1,\sigma_0) =&\, -(1-\rho_1)\left(1+\rho_1^3\right)^2\left(\rho_1+\rho_1^3+(1-\rho_1)\sigma_0\right), \\
\Phi_2(\rho_1,\sigma_0) =&\, \rho_1(3\sigma_0-1)
-3\rho_1^2(\sigma_0-1)
-\rho_1^3(\sigma_0(5\sigma_0-9)+4) \\
&\quad +\rho_1^4(\sigma_0(13\sigma_0-19)\sigma_0+12)
-2\rho_1^5(\sigma_0-1)(7\sigma_0-4) \\
&\quad +2\rho_1^6(\sigma_0-1)(5\sigma_0-6)
-\rho_1^7(5\sigma_0(\sigma_0-3)-4)
+\rho_1^8(\sigma_0-4)(\sigma_0-1) \\
&\quad -\rho_1^9(\sigma_0-1)
+\rho_1^{10}(\sigma_0+1), \\
\Phi_3(\rho_1,\sigma_0) =&\, \rho_1^2(1-\rho_1)^2(\sigma_0-1)\left(\rho_1+\rho_1^3+(1-\rho_1)\sigma_0\right)^2.
\end{align*}
Clearly, the term $\Phi_1(\rho_1,\sigma_0)$ is strictly negative and the term $\Phi_3(\rho_1,\sigma_0)$ is strictly positive for all $0<\rho_1<1$,
and for all $\sigma_0>1$.
Recall that by Lemma \ref{lem:alpha vanishing VIII}, $\alpha_{v,\rho,\sigma_0}(\rho)$ has one root $\rho_0\in(0,\rho_\star]$,
if $\sigma_0\geq \sigma_\star$, where $\rho_\star \approx 0.596$, $\sigma_\star\approx 9.444$.
As defined in the beginning of this proof we have, $\sigma_\delta := \sigma_0 + \delta$.
In the last two paragraphs of this proof, we show that
$\delta\cdot\Phi_1(\rho_1,\sigma_0) + \Phi_2(\rho_1,\sigma_0)$ is strictly negative for some fixed $\sigma_0>1$,
and some fixed $\delta>0$ for all $\rho_1\in(0,\rho_0)$.
As noted before, $\Phi_1(\rho_1,\sigma_0)$ is strictly negative for all $0<\rho_1<1$. Therefore, we have
$\delta\cdot\Phi_1(\rho_1,\sigma_0) + \Phi_2(\rho_1,\sigma_0)\geq \delta_1\cdot\Phi_1(\rho_1,\sigma_0) + \Phi_2(\rho_1,\sigma_0)$,
if $\delta\leq\delta_1$, for all $\rho_1\in(0,\rho_0)$.
So no boundary/intersection point can occur for any $\sigma_0 + \delta_1 \geq \sigma_0 + \delta$.
Hence, no MPF $w$ exists for any $\sigma\geq\sigma_\delta$.
Using a computer algebra system, we can compute $\delta\cdot\Phi_1(\rho_1,\sigma_0) + \Phi_2(\rho_1,\sigma_0)$
for $\sigma_0 = \sigma_\star$, and $\delta=0.4558$, i.\,e.\ $\sigma_\delta=\sigma_0+\delta\approx 9.899$.
We get that $\delta\cdot\Phi_1(\rho_1,\sigma_0) + \Phi_2(\rho_1,\sigma_0)$ is strictly negative for all
$0<\rho_1\leq \rho_\star$.
By Lemma \ref{lem:alpha vanishing VIII}, $\alpha_{v,\rho,\sigma_0}(\rho)$ has one root $\rho_0\in(0,\rho_\star]$,
if $\sigma_0\geq \sigma_\star$, where $\rho_\star \approx 0.596$, $\sigma_\star\approx 9.444$.
Hence, there are no MPF $w$ for any $F=|A|^\sigma$ with $\sigma\geq 9.899$.
By Lemma \ref{lem:tedious A}, we calculate $\delta\cdot\Phi_1(\rho_1,\sigma_0) + \Phi_2(\rho_1,\sigma_0)$
for $\sigma_0=10$, and $\delta=1$.
We get that $\delta\cdot\Phi_1(\rho_1,\sigma_0) + \Phi_2(\rho_1,\sigma_0)$ is strictly negative for all
$0<\rho_1\leq 1/2$.
By Lemma \ref{lem:alpha vanishing VIII}, $\alpha_{v,\rho,\sigma}(\rho)$ has one root $\rho_0\in(0,1/2]$ for all $\sigma\geq 10$.
Hence, there are no MPF for any $F=|A|^\sigma$ with $\sigma\geq 11$.
This concludes the proof.
\end{proof}
\begin{lemma}[Calculations for Theorem \ref{thm:|A| III}]\label{lem:tedious A}
Define
\begin{align*}
\tilde{\Phi}(\rho,\sigma,\delta) := \delta\cdot\Phi_1(\rho,\sigma) + \Phi_2(\rho,\sigma),
\end{align*}
where
\begin{align*}
\Phi_1(\rho,\sigma) =&\, -(1-\rho)\left(1+\rho^3\right)^2\left(\rho+\rho^3+(1-\rho)\sigma\right), \\
\Phi_2(\rho,\sigma) =&\, \rho(3\sigma-1)
-3\rho^2(\sigma-1)
-\rho^3(\sigma(5\sigma-9)+4) \\
&\quad +\rho^4(\sigma(13\sigma-19)\sigma+12)
-2\rho^5(\sigma-1)(7\sigma-4) \\
&\quad +2\rho^6(\sigma-1)(5\sigma-6)
-\rho^7(5\sigma(\sigma-3)-4)
+\rho^8(\sigma-4)(\sigma-1) \\
&\quad -\rho^9(\sigma-1)
+\rho^{10}(\sigma+1),
\end{align*}
as in Theorem \ref{thm:|A| III}.
Let $\sigma=10$, and $\delta=1$.
Then $\tilde{\Phi}(\rho,10,1)$ is strictly negative for all $0<\rho\leq 1/2$.
Furthermore, $\alpha_{v,\rho,\sigma}(\rho)$ from Lemma \ref{lem:alpha vanishing VIII} has
one root $\rho_0\in(0,1/2]$ for all $\sigma\geq 10$.
\end{lemma}
\begin{proof}
We begin this proof by calculating $\tilde{\Phi}(\rho,10,1)$. We obtain
\begin{align*}
\tilde{\Phi}(\rho,10,1) =&\, -10+48\rho-36\rho^2-435\rho^3+1161\rho^4-1206\rho^5 \\
&\quad +780\rho^6-333\rho^7+45\rho^8-10\rho^9+12\rho^{10}.
\end{align*}
Firstly, let $0<\rho\leq 0.3$. We rewrite $\tilde{\Phi}(\rho,10,1)$ as a sum of five polynomials
\begin{align*}
\tilde{\Phi}(\rho,10,1) = \tilde{\Phi}_1(\rho) + \tilde{\Phi}_2(\rho) + \tilde{\Phi}_3(\rho) + \tilde{\Phi}_4(\rho) + \tilde{\Phi}_5(\rho),
\end{align*}
where
\begin{align*}
\tilde{\Phi}_1(\rho) =&\, -10+48\rho-36\rho^2-85\rho^3,\\
\tilde{\Phi}_2(\rho) =&\, -350\rho^3+1161\rho^4,\\
\tilde{\Phi}_3(\rho) =&\, -1206\rho^5+780\rho^6,\\
\tilde{\Phi}_4(\rho) =&\, -333\rho^7+45\rho^8,\\
\tilde{\Phi}_5(\rho) =&\, -10\rho^9+12\rho{10}.
\end{align*}
Now, we show that $\tilde{\Phi}_1(\rho)$ is strictly negative, and
$\tilde{\Phi}_i(\rho),\,i=2,3,4,5$, is non-positive for all $0<\rho\leq 0.3$.
\begin{itemize}
\item $\tilde{\Phi}_1(\rho)$ has no critical points, and $\tilde{\Phi}_1(0)=-10$, and $\tilde{\Phi}_1(0.3)=-1.135$,
\item $\tilde{\Phi}_2(\rho)/\rho^3$ has no roots, and $\tilde{\Phi}_2(\rho)/\rho^3_{|\rho=1/4}=-239/4$,
\item $\tilde{\Phi}_3(\rho)/\rho^5$ has no roots, and $\tilde{\Phi}_3(\rho)/\rho^5_{|\rho=1/4}=-1011$,
\item $\tilde{\Phi}_4(\rho)/\rho^7$ has no roots, and $\tilde{\Phi}_4(\rho)/\rho^7_{|\rho=1/4}=-1287/4$.
\item $\tilde{\Phi}_5(\rho)/\rho^9$ has no roots, and $\tilde{\Phi}_5(\rho)/\rho^9_{|\rho=1/4}=-7$.
\end{itemize}
Secondly, let $0.3\leq\rho\leq 0.5$. We rewrite $\tilde{\Phi}(\rho,10,1)$ as a sum of five different polynomials
\begin{align*}
\tilde{\Phi}(\rho,10,1) = \tilde{\Phi}_1(\rho) + \tilde{\Phi}_2(\rho) + \tilde{\Phi}_3(\rho) + \tilde{\Phi}_4(\rho) + \tilde{\Phi}_5(\rho),
\end{align*}
where
\begin{align*}
\tilde{\Phi}_1(\rho) =&\, -10+48\rho-36\rho^2-57\rho^3,\\
\tilde{\Phi}_2(\rho) =&\, -378\rho^3+1161\rho^4-855\rho^5,\\
\tilde{\Phi}_3(\rho) =&\, -351\rho^5+780\rho^6-306\rho^7,\\
\tilde{\Phi}_4(\rho) =&\, -27\rho^7+45\rho^8,\\
\tilde{\Phi}_5(\rho) =&\, -10\rho^9+12\rho^{10}.
\end{align*}
Now, we show that
$\tilde{\Phi}_i(\rho),\,i=1,\ldots,5$, is strictly negative for all $0.3\leq\rho\leq 0.5$.
\begin{itemize}
\item $\tilde{\Phi}_1(\rho)$ has one critical point at $\rho_{CP}=4/57(-3+\sqrt{66})\approx 0.36$,
and $\tilde{\Phi}_1(0.3)=-0.379$, $\tilde{\Phi}_1(\rho_{CP})=2/1083(-11463+1408\sqrt{66})\approx -0.045$,
$\tilde{\Phi}_1(0.5)=-2.125$,
\item $\tilde{\Phi}_2(\rho)$ has one critical point at $\rho_{CP}=3/475(86-\sqrt{746})\approx 0.371$,
and $\tilde{\Phi}_2(\rho)/\rho^3_{|\rho=0.3}=-106.65$, $\tilde{\Phi}_2(\rho)/\rho^3_{|\rho=\rho_{CP}}=-27/2375(2206+129\sqrt{746})\approx-3.317$,
$\tilde{\Phi}_2(\rho)/\rho^3_{|\rho=0.5}=-11.25$,
\item $\tilde{\Phi}_3(\rho)$ has one critical point at $\rho_{CP}=1/238(260-\sqrt{21190})\approx 0.481$,
and $\tilde{\Phi}_3(\rho)/\rho^5_{|\rho=0.3}=-144.54$, $\tilde{\Phi}_3(\rho)/\rho^5_{|\rho=\rho_{CP}}=-78/833(-229+5\sqrt{21190})\approx-46.71$,
$\tilde{\Phi}_3(\rho)/\rho^5_{|\rho=0.5}=-37.5$,
\item $\tilde{\Phi}_4(\rho)/\rho^7$ has no root, and $\tilde{\Phi}_4(\rho)/\rho^7_{\rho=0.4}=-9$,
\item $\tilde{\Phi}_5(\rho)/\rho^9$ has no root, and $\tilde{\Phi}_5(\rho)/\rho^9_{\rho=0.4}=-5.2$.
\end{itemize}
This concludes the proof.
\end{proof}
\begin{lemma}[Norm of the second fundamental form]\label{lem:|A| expanding}
$\left.\right.$ \\
Let $F=F^\sigma_2=\sgn(\sigma)\cdot |A|^\sigma$. Then
\begin{align*}
w=\frac{(a-b)^2(a^3+b^3)(a\,b)^\sigma}{(a^2+b^2)^{1/2}(a\,b)^2}
\end{align*}
is a maximum-principle function, if $\sigma\in[-1,0)$.
\end{lemma}
\begin{proof}
It is rather easy to see that $w$ is a MPF except for MPF condition \eqref{IV}.
Here, we need a computer algebra program.
We compute $C(a,b)$, $E(a,b)$, $G(a,b)$ for the given $w$ and $F$.
Then we use a Monte-Carlo method to check for non-positivity of these terms.
Since $C(a,b)$, $E(a,b)$, $G(a,b)$ are homogeneous, $C(a,b)=C(b,a)$, and $E(a,b)=G(b,a)$,
it suffices to check for non-positivity testing only on $(a,b)=(\rho,1)$, where $0<\rho\leq 1$.
In a separate paper, using this MPF we will show convergence to a sphere at $\infty$.
\end{proof}
\begin{corollary}[Norm of the second fundamental form]\label{cor:|A|}
$\left.\right.$ \\
Let $F=F^\sigma_2=\sgn(\sigma)\cdot |A|^\sigma$.
Then we have
\begin{center}
\begin{tabular}{| l | c || c | c |}
\hline
& MPF & & \\
\hline \hline
$\Bigg.\Bigg. \sigma\geq 9.
\begin{small}
\text{89}
\end{small}
$ & non-existent & & Theorem \ref{thm:|A| III} \\
\hline
$\Bigg.\Bigg. \sigma\in(8.
\begin{small}
\text{15}
\end{small}
,9.
\begin{small}
\text{89}
\end{small}
)$ & open problem & & \\
\hline
$ \Bigg.\Bigg. \sigma\in(1,8.
\begin{small}
\text{15}
\end{small}
]$ & $\frac{\left(a-b\right)^2\left(a^2+b^2\right)^\sigma}{\left(a\,b\right)^2}$ & B. Andrews, X. Chen & \cite{ac:surfaces} \\
\hline
$\Bigg.\Bigg. \sigma\in(0,1]$ & non-existent & & Theorem \ref{thm:|A| II} \\
\hline \hline
$\Bigg.\Bigg. \sigma\in[-1,0)$ & $\frac{\left(a-b\right)^2\left(a^3+b^3\right)\left(a\,b\right)^\sigma}{\left(a^2+b^2\right)^{1/2}\left(a\,b\right)^2}$ & & Lemma \ref{lem:|A| expanding} \\
\hline
$\Bigg.\Bigg. \sigma<-1$ & non-existent & & Theorem \ref{thm:|A| I} \\
\hline
\end{tabular}
\end{center}
\end{corollary}
\section{Trace of the second fundamental form, $F^\sigma_\sigma$}
In this chapter, we discuss $F^\sigma_\sigma=\sgn(\sigma)\cdot \tr A^\sigma$
for all $\sigma\in{\mathbb{R}}\setminus\{0\}$.
For $\sigma>1$, the vanishing function
\begin{align*}
v = \frac{(a-b)^2\,F^2}{(a\,b)^2}
\end{align*}
from example \ref{exm:vanishing} is MPF.
This is a result by B. Andrews and X. Chen \cite{ac:surfaces}.
Interestingly, to date, this is the only time
where convergence to a round point can be shown
for a normal velocity homogeneous of
an arbitrarily high degree using a MPF.
For $0<\sigma\leq 1$, we prove the non-existence of MPF by
Lemma \ref{lem:one} and Lemma \ref{lem:zeroone}.
This is Theorem \ref{thm:trA}.
For expanding $F^\sigma_\xi$, i.\,e.\ $\sigma<0$,
the necessary conditions of Theorem \ref{thm:necessary}
do not impose any constraints on the existence of MPF.
We only know that for $\sigma=-1$ a MPF exists.
Since
\begin{align*}
F^{-1}_{-1} =&\, -\tr A^{-1} \\
=&\, -\left(a^{-1}+b^{-1}\right) \\
=&\, -\left(\frac{a+b}{a\,b}\right) \\
=&\, -\frac{H}{K},
\end{align*}
we have a MPF by O. Schn\"urer \cite{os:surfaces}.
\begin{theorem}[Trace of the second fundamental form]\label{thm:trA}
$\left.\right.$ \\
Let $F=F^\sigma_\sigma=\sgn(\sigma)\cdot \tr A^\sigma$.
Then there are no $MPF$ if $0<\sigma\leq 1$.
\end{theorem}
\begin{proof}
This is a direct consequence of Lemma \ref{lem:one} and Lemma \ref{lem:zeroone}.
\end{proof}
\begin{corollary}[Trace of the second fundamental form]\label{cor:trA}
$\left.\right.$ \\
Let $F=F^\sigma_\sigma=\sgn(\sigma)\cdot \tr A^\sigma$.
Then we have
\begin{center}
\begin{tabular}{| l | c || c | c |}
\hline
& MPF & & \\
\hline \hline
$\Bigg.\Bigg. \sigma>1$ & $\frac{(a-b)^2\left(a^\sigma+b^\sigma\right)^2}{(a\,b)^2}$ & B. Andrews, X. Chen & \cite{ac:surfaces} \\
\hline
$\Bigg.\Bigg. \sigma\in(0,1]$ & non-existent & & Theorem \ref{thm:trA} \\
\hline \hline
$\Bigg.\Bigg. \sigma\in(-1,0)$ & open problem & & \\
\hline
$\Bigg.\Bigg. \sigma=-1$ & $\frac{(a-b)^2}{(a\,b)^2}$ & O. Schn\"urer & \cite{os:surfaces} \\
\hline
$\Bigg.\Bigg. \sigma<-1$ & open problem & & \\
\hline
\end{tabular}
\end{center}
\end{corollary}
\section{Outlook}
`The search is over.' \\
At least, this is what had been our second choice for a title of this paper.
It sounded a bit too dramatic, so we decided to go for the first.
However, the MPF Ansatz followed by B. Andrews, O. Schn\"urer, F. Schulze,
and many more, now seems to be exploited to the last drop.
Certainly that is for the Gauss curvature, $F^\sigma_0$,
the mean curvature, $F^\sigma_1$,
and the norm of the second fundamental form, $F^\sigma_2$.
But the search is not over.
In our opinion, MPF condition \eqref{IV} is sufficient,
not necessary, and obviously far too restrictive for many normal velocities.
For example, in case of $F^\sigma_0$, we know that
no MPF exists for any power $\sigma<-2$, but we still have
convergence to a sphere at $\infty$ by C. Gerhardt \cite{cg:non}.
And there are no apparent reasons why convergence to a round point
should stop for powers of the mean curvature greater than 6,
or for powers of the norm of the second fundamental form greater than 11.
We conjecture that we can exploit the MPF Ansatz further,
if we can find a way to weaken MPF condition \eqref{IV}.
Then we also might be able to recycle some of the vanishing functions
that are not MPF according to the present MPF Definition \ref{def:mpf}.
\bibliographystyle{amsplain}
\def\weg#1{} \def\underline{\rule{1ex}{0ex}}} \def$'$} \def\cprime{$'$} \def\cprime{$'${$'${\underline{\rule{1ex}{0ex}}} \def$'$} \def\cprime{$'$} \def\cprime{$'${$'$}
\def$'$} \def\cprime{$'$} \def\cprime{$'${$'$} \def$'$} \def\cprime{$'$} \def\cprime{$'${$'$} \def$'$} \def\cprime{$'$} \def\cprime{$'${$'$}
\providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
\providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR }
\providecommand{\MRhref}[2]{%
\href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2}
}
\providecommand{\href}[2]{#2}
|
\section{Introduction}
The Perseus Cluster is the brightest galaxy cluster in the X-ray sky and has been attracting astronomers attention over the last 40 years. Being the X-ray brightest cluster, Perseus has led to many fundamental cluster discoveries, such as peaked cluster emission \citep{Fab74}, strong FeXXV and FeXXVI lines and thermal emission mechanism of the gas \citep{Mit76}, presence of bubbles of relativistic plasma \citep{Boe93}. The cluster was recently observed deeply with the {\it Chandra} \citep[see e.g.][and references therein]{Fab00,Fab11,Sch02,San07}, {\it XMM-Newton} \citep[see e.g.][and references therein]{Boe02,Chu03,Chu04,Mat13} and {\it Suzaku} \citep[see e.g.][and references therein]{Sim11,Ued13,Wer13,Tam14,Urb14} observatories, as well as by the previous-generation X-ray observatories \citep[see e.g.][]{For72,Mus81,Bra81,Arn94,Fab94,Fur01,Gas04}. These observations unveiled a variety of structures present in the hot gas of the cluster, reflecting richness and complexity of different physical processes occurring in the intracluster plasma. For example, an east-west asymmetry in the X-ray surface brightness (SB) \citep{Sch92}, aligned with a chain of bright galaxies, suggests an ongoing modest merger, which can induce sloshing of the gas and drive gas turbulence \citep{Chu03}, while the central few arcmin are dominated by AGN activity in a form of inflated bubbles of relativistic plasma \citep[see e.g.][]{Boe93,Chu00,Fab00} and weak shocks around them. Rising buoyantly and expanding, the bubbles uplift cooler X-ray gas from the core, producing filamentary structures seen in the optical, far-ultraviolet and soft X-ray, and excite internal waves, energy of which is transported to gas turbulence \citep{Chu01,Omm04,For07,Fab08,Can14,Hil14}.
The truly spectacular statistics accumulated by the {\it Chandra} observations of Perseus (over 85 million counts within 6 arcmin from the center, with 100s and 1000s counts per sq. arcsec) makes this data set an extremely powerful tool to probe structures in the intracluster medium (ICM) on a range of spatial scales. Carefully modeling the Poisson noise, one can probe small-scale fluctuations, which we cannot see in the cluster images by visual inspection. Here, we present statistical analysis of these SB and density fluctuations, using the power spectrum (PS) statistics. We will examine the radial variations of the PS within the Perseus core ($220\times220$ kpc), while its energy dependence will be considered in our future works.
Statistical analysis of the X-ray SB and/or density, pressure fluctuations was done for two clusters so far: the Coma Cluster \citep{Sch04,Chu12} and AWM7 \citep{San12}. It was shown, for example, that the amplitude of density fluctuations in Coma ranges from $5$ to $10$ per cent on scales $30-500$ kpc, leading to nontrivial constraints on perturbations of the cluster gravitational potential, turbulence, entropy variations and metallicity. The fluctuations appear to be correlated, implying suppression of strong isotropic turbulence and conduction \citep{San13}.
The {\it Chandra} data on Perseus allow us to probe fluctuations on similar or even smaller scales, down to few kpc. The main questions we address here include: what is the amplitude of density fluctuations and how does it vary with the distance from the cluster center (Section \ref{sec:den_fluct}); what is the power spectrum of gas turbulent motions (Section \ref{sec:vel_ps}); how clumpy is the gas in the core (Section \ref{sec:clump}). To answer these questions, various instrumental effects (e.g. PSF variations, detector QE fluctuations), contribution of the unresolved point sources, subtraction of the Poisson noise as well as its fluctuations need to be treated carefully. We extensively investigate various systematic uncertainties of the analysis and discuss them in Section \ref{sec:uncert}.
This paper is complementing our series of publications on SB and density fluctuations in the core of the Perseus Cluster. One application of the velocities of gas motions measured here is discussed in \citet{Zhu14b}. There we show that turbulent dissipation in the cluster core produces enough heat to balance locally radiative energy losses from the ICM at each radius. Therefore, turbulence may play an important role in AGN feebdack loop and be the key element in resolving the gas cooling flow problem \citep{Zhu14b}. In our next paper, we will discuss the nature of the fluctuations by measuring cross-spectra of fluctuations in different energy bands (Zhuravleva et al., 2015b, in prep.). Similar analysis for the Virgo and Coma clusters are in progress.
Throughout the paper we adopt $\Lambda$CDM cosmology with $\Omega_m=0.3$, $\Omega_{\Lambda}=0.7$ and $h=0.72$. At the redshift of Perseus, $z=0.01755$, the corresponding angular diameter distance is 71.5 Mpc and 1 arcmin corresponds to a physical scale 20.81 kpc. Position of the cluster center is taken from the NASA/IPAC Extragalactic Database\footnote{https://ned.ipac.caltech.edu/} and corresponds to the position of the central galaxy NGC1275. The total galactic HI column density\footnote{We use the mean value of the Galactic absorption within the field of the Perseus Cluster. Exclusion of soft photons ($0.5-1$ keV) from the SB analysis does not change the resulting amplitude of the SB and density fluctuations. Therefore, even if fluctuations of the Galactic absorption are present, their role in the cluster core is minor.} is 1.36$\cdot$10$^{21}$ cm$^{-2}$. Solar abundances of heavy elements are taken from \citet{And89}. We define the characteristic scale-dependent amplitude $A_{3D}(k)$ through the power spectrum $P_{3D}(k)$ as follows $A^2_{3D}(k)=\int 4\pi P_{3D}(k)k^2dk=4\pi P_{3D}(k) k^3$. A wavenumber $k$ is related to a spatial scale $l$ without a factor $2\pi$, i.e. $k=1/l$.
\section{Initial data processing}
\label{sec:init}
\begin{figure}
\begin{minipage}{0.49\textwidth}
\includegraphics[trim=0 150 0 100,width=1\textwidth]{plot_fluxes_acisi_perseus.pdf}
\end{minipage}
\caption{Emissivity of an optically thin plasma as a function of temperature observed by the ACIS-S detector on {\it Chandra} in two energy bands: $0.5-3.5$ (solid) and $3.5-7.5$ (dashed) keV. Abundance of heavy elements is assumed 0.5 relative to Solar. Shaded region shows a range of gas temperatures in the core of the Perseus Cluster ($3-6.5$ keV). Notice, that for the $0.5-3.5$ keV band, the emissivity is almost independent of the temperature within this region. Therefore, the X-ray surface brightness in this shaded band $I_X$ is just a function of gas electron number density $n_e$, namely, $I_X\sim\int n_e^2 dl$, where $l$ is the length of the line of sight.
\label{fig:fluxes}
}
\end{figure}
For our analysis we use public {\it Chandra} data with the
following ObsIDs: 3209, 4289, 4946, 4947, 4948, 4949, 4950, 4951,
4952, 4953, 6139, 6145, 6146, 11713, 11714, 11715, 11716, 12025,
12033, 12036, 12037. The initial data processing is done using
the latest calibration data and following the standard procedure
described in \citet{Vik05}. For the analysis, we choose $0.5-3.5$ keV energy band. In this band, the gas emissivity has a weak temperature dependence as shown in Fig. \ref{fig:fluxes}, and, therefore, the uncertainty associated with the conversion of PS of SB fluctuations to the PS of density fluctuations is reduced (see Section \ref{sec:den_fluct}).
Correcting for the exposure, vignetting effect and subtracting the background, a mosaic image of the Perseus Cluster is produced, which is shown in Fig. \ref{fig:im_pers}. The combined exposure map is shown in Fig. \ref{fig:psf_exp}. The total exposure of the cleaned data set is $\sim1.4$ Ms. Notice, that the
exposure coverage is inhomogeneous. It varies from few$\cdot
10^6$ s in the center down to few$\cdot 10^5$ s in some distal regions. The uncertainties associated with such inhomogeneous coverage are discussed in Section \ref{sec:uncert}.
\begin{figure*}
\includegraphics[trim=30 120 30 130,width=0.49\textwidth]{pers.pdf}
\includegraphics[trim=30 120 30 130,width=0.49\textwidth]{pers_div.pdf}
\caption{{\bf Left:} {\it Chandra} mosaic image of the Perseus Cluster
in the $0.5-3.5$ keV band. The units are counts/s/pixel. The size of each pixel is 1 arcsec. {\bf
Right:} residual image of the cluster (the initial image divided by
the best-fitting spherically-symmetric $\beta-$model of the surface brightness), which emphasizes the surface
brightness fluctuations present in the cluster. Point sources are
excised from the image. Black circles show the set of annuli used in the analysis of the fluctuations. The width of each annulus is $1.5'$ ($\approx 30$ kpc). The
outermost boundary is at a distance $10.5'$ ($\approx 220$ kpc) from the center. For display purposes, both images are lightly smoothed with a $3''$ Gaussian.
\label{fig:im_pers}
}
\end{figure*}
\begin{figure*}
\includegraphics[trim=30 120 30 130,width=0.49\textwidth]{psf_1.pdf}
\includegraphics[trim=30 120 30 130,width=0.49\textwidth]{exp_1.pdf}
\caption{{\bf Left:} simulated map of the combined {\it Chandra} PSF within the mosaic image of the Perseus Cluster. The random positions of individual
PSFs are used (see Section \ref{sec:init} for details). {\bf Right:} the combined exposure map in seconds (lightly smoothed with a $3''$ Gaussian for display purposes).
\label{fig:psf_exp}
}
\end{figure*}
The Perseus dataset includes observations with different offsets. For the analysis of SB fluctuations, it is essential to know the PSF at each point within the cluster image. We randomly generate positions, where the PSF is checked. The PSF at each position for each observation is determined from the {\it Chandra} PSF libraries \citep{Kar01}. By combining the PSF model images with weights proportional to exposure of each observation, the final PSF map is obtained. These maps are used to correct the PS of the SB fluctuations (Section \ref{sec:sb_fluct}). In order to illustrate the variations of the PSF shape within the cluster mosaic image, we show the PSF map divided by the exposure map in Fig. \ref{fig:psf_exp}. Notice a non-monotonic behavior of the PSF and its elongated (distorted) shape close to the edges of the image.
As the next step, we excise point sources from the image of the cluster. Using the wvdecomp tool, the point source candidates are found as peaks in the image with S/N $> 4$. In order to verify the significance of each source detection, we smooth heavily the image of the cluster with a $20''$ Gaussian, excluding the brightest point sources first. We then subtract the smoothed image from the initial image to remove the cluster contribution and calculate the point source fluxes in small regions around each point source candidate accounting for the combined PSF as $F_{\rm p.s.}=\sum I_{\rm p.s.}PSF/\sum PSF^2$, where $I_{\rm p.s.}$ refers to the residual image. The uncertainties on the fluxes are $\sigma_{F_{\rm p.s.}}=\sqrt{\sum C_iPSF^2}/\sum PSF^2$, where $C_i$ is the number of counts in the initial cluster image. Finally, sources with $F_{\rm p.s.}/\sigma_{F_{\rm p.s.}}>4$ are cut out from the images, using circles with the radius 1.5 times the 90 per cent radius of the combined PSF (see Fig. \ref{fig:im_pers}). Contribution of the unresolved point sources is estimated using sensitivity maps and Log N - Log S distribution of resolved sources. In case of the Perseus core, the contribution of unresolved point sources is negligible.
\section{Spherically-symmetric model}
\label{sec:model}
The spherically-symmetric radial surface brightness profile of the Perseus is shown in Fig. \ref{fig:prof}. Fitting this profile with a $\beta-$model assuming photon counting noise only, gives the core radius $r_c=1.26$ arcmin and the slope of the model $\beta=0.53$ with negligible uncertainties. Allowing for 10 per cent systematic error on the SB in each annulus, the uncertainties become 0.1 and 0.01 for $r_c$ and $\beta$ respectively. Throughout the paper, this model is used as a default model of the unperturbed cluster. Sensitivity of our results to the choice of the model are discussed in Section \ref{sec:uncert}.
The deprojected thermodynamic properties of the cluster, namely, the number electron density $n_e$ and the electron temperature $T_e$, are obtained from projected spectra using a procedure described in \cite{Chu03}. The deprojected spectra are fitted in the broad $0.6-9$ keV band using the XSPEC \citep{Smi01,Fos12} code and APEC plasma model based on ATOMDB version 2.0.1. The abundance of heavy elements is assumed to be either a) constant with radius 0.5 relative to Solar, or b) a free parameter in the plasma model. Fig. \ref{fig:prof} shows the deprojected data and their approximations with continuous, smooth functions. Using these approximations, the sound speed $c_s$ for an ideal monatomic gas is obtained as
\be
c_s=\disp\sqrt{\gamma\frac{k_BT_e}{\mu m_p}},
\ee
where $\gamma=5/3$ is the adiabatic index, $k_B$ is the Boltzmann constant, $\mu=0.61$ is the mean particle weight and $m_p$ is the proton mass. We also show the characteristic mean free path $\lambda_{\rm mfp}$ in the gas, which is numerically defined as \citep[e.g.][]{Sar88}
\be
\lambda_{\rm mfp}=23 {\rm \,kpc} \left(\frac{T_e}{10^8 {\rm \,K}}\right)^2\left(\frac{n_e}{10^{-3} {\rm \,cm^{-3}}}\right)^{-1}.
\ee
Notice, that the mean free path is below kpc in central few tens of kpc and increases with distance, reaching $\sim 10$ kpc at distance 400 kpc from the center.
\begin{figure}
\begin{minipage}{0.49\textwidth}
\includegraphics[trim=20 155 30 90,width=1\textwidth]{perseus_deproj_sb_mfp_cs.pdf}
\end{minipage}
\caption{Radial profiles of thermodynamic properties of the Perseus
Cluster obtained from {\it Chandra} observations in the $0.6-9$ keV band. Orange (filled)
and red (open) points with $1\sigma$ error bars on the left panels show the deprojected gas number electron
density $n_e$, electron temperature $T_e$ and abundance of heavy
elements relative to solar $Z/Z_{\odot}$, assuming the latter one
is constant 0.5 relative to Solar or is a free
parameter in the fitting spectral model of an optically thin plasma
respectively. Analytic approximations of the observed profiles by smooth functions are shown with
black solid curves. Blue points on the top right panel show the
spherically-symmetric X-ray surface brightness profile obtained from the $0.5-3.5$ keV image, while the
black curve shows its best-fitting $\beta-$model with the core radius
$r_c=1.26$ arcmin and the slope $\beta=0.53$. The cluster sound speed
$c_s$ and the mean free path $\lambda_{mfp}$ calculated from the
approximations are shown in the middle and bottom panels on the right.
\label{fig:prof}
}
\end{figure}
\section{Power spectrum of the surface brightness fluctuations}
\label{sec:sb_fluct}
X-ray SB of galaxy clusters is typically peaked towards centers and decreases rapidly with radius. This is especially true for relatively dynamically relaxed clusters like Perseus \citep{Man14}. Therefore, the PS of the initial SB will be dominated by this gradient, leading to a large power on small wavenumbers $k$ and contamination of the power on larger $k$. In order to avoid this, we first remove the global SB gradient and analyze fluctuating part of the SB only. In order to work with dimensionless units of the SB fluctuations, we divide the image of the Perseus Cluster by a simple spherically-symmetric $\beta-$model of the SB\footnote{One can also subtract the model, however in this case additional weights proportional to the global SB profile have to be introduced in order to treat the fluctuations at different radii equally. Clearly, the choice of this underlying model is arguable, but we will return to this problem in Section \ref{sec:uncert}.}. Fig. \ref{fig:im_pers} shows the residual image of the SB fluctuations in Perseus. Notice a variety of structures of different sizes and morphologies present in the cluster, especially within the central $3$ arcmin. Let us quantify these structures statistically, by measuring their power as a function of length scale.
We use a modified $\Delta-$variance method \citep{Are12} to calculate the PS of SB fluctuations. This method is tuned to cope with non-periodic data with gaps, the PS of which are smooth functions. Dashed curves in Fig. \ref{fig:p2d_a2d} show the PS of SB fluctuations in the central $3$ arcmin (central AGN is excluded) and in a broad annulus $3-10.5$ arcmin. The spectra decrease with wavenumber $k$ and reach constant value on small scales, where the signal is dominated by the Poisson noise.
Knowing the number of counts $n_{cnts}$ in each pixel in the cluster image in counts, the contribution of the Poisson noise is evaluated. A white noise has a flat spectrum; however, small deviations are still possible due to imperfections of the $\Delta-$variance method (e.g. due to filter behavior close to the edges of the image or near gaps, see \citet{Are12} for details). We made 100 realizations of the Poisson noise simply multiplying $\sqrt{n_{cnts}}$ by a random number from a normal distribution (with mean 0 and variance 1) in each pixel. For each realization, the PS of the Poisson image is calculated and, finally, the mean and the scatter of the resulting spectrum are obtained. Dotted horizontal regions in Fig. \ref{fig:p2d_a2d} show the level of the Poisson noise with the uncertainties in both considered regions. Subtracting the PS of the noise from the total (SB fluctuations+noise) PS, the PS of SB fluctuations is obtained (hatched region with solid boundaries in Fig. \ref{fig:p2d_a2d}). The uncertainties reflect the photon counting noise at high frequencies ($1\sigma$ uncertainty is shown) and stochasticity of the signal at low frequencies. We do not separate these two types of uncertainties. Multiple realizations of the white noise PS $P_{\rm wn}$ estimate the scatter $\sigma_{p_{\rm wn}}$. These errors are then scaled to the measured PS of SB fluctuations $P$ by a simple multiplication
\be
\sigma_p=\disp\sigma_{p_{\rm wn}}\frac{P}{P_{\rm wn}}.
\ee
Based on experiments with the whole process pipeline, we conservatively added 3 per cent systematic uncertainty to the scatter of the Poisson noise PS, which takes into account subtle numerical and instrumental effects.
Finally, the PS of SB fluctuations have to be corrected for the {\it Chandra} PSF and for the contribution of the unresolved point sources on small scales. Performing multiple realizations of the PSF maps (see Section \ref{sec:init}), and each time calculating the PS of the map within the region of our interest, the mean spectrum of the PSF is obtained. Dash-dotted curves in Fig. \ref{fig:p2d_a2d} show the PSF PS in central $3$ arcmin and in the annulus $3 - 10.5$ arcmin. The deviations of the PSF shape from a $\delta-$function are reflected in a drop of power on the high-$k$ end of the spectrum. Dividing the PS of the SB fluctuations by the PSF power spectrum, the SB power suppressed by the PSF is recovered.
\begin{figure*}
\includegraphics[trim=10 180 30 70,width=0.49\textwidth]{p2d_perseus.pdf}
\includegraphics[trim=10 180 30 70,width=0.49\textwidth]{a2d_perseus.pdf}
\caption{{\bf Left:} power spectra (PS) of the X-ray surface
brightness (SB) fluctuations in the Perseus Cluster in the central $3$ arcmin (purple) and
in the annulus $3-10.5$ arcmin (orange). Dash curves: the initial
PS. Dotted regions: PS of the Poisson noise. Dash-dotted curves: PS
of the combined {\it Chandra} PSF in both analyzed regions. Hatched region with solid boundaries: PS of the SB fluctuations after subtraction of the Poisson noise. Both statistical ($1\sigma$) and stochastic uncertainties are taken into account. Notice, that very high statistics of counts accumulated by
the 1.4 Ms observations allows us to probe fluctuations on a broad range
of length scales, from $>$ 100 kpc down to few kpc (comparable with the mean free path). {\bf Right:} amplitude of the SB fluctuations in the same regions, obtained from the PS
corrected for the PSF (see Section \ref{sec:sb_fluct}). Notice, that the amplitude of fluctuations varies with the radius. For these two regions a maximal difference by a factor of 4 is reached on $\sim 17$ kpc scale.
\label{fig:p2d_a2d}
}
\end{figure*}
The contribution of unresolved point sources is estimated by using the Perseus sensitivity map and by obtaining the shape and normalization of the Log N-Log S distribution of the resolved sources. In case of the Perseus Cluster, the contribution of the unresolved sources is negligible (within the cluster core).
After applying all corrections described above, the characteristic amplitude $A_{2D}(k)$ of the SB fluctuations is calculated, namely $A_{2D}(k)=\sqrt{P_{2D}(k)2\pi k^2}$, where $P_{2D}(k)$ is the measured PS. The amplitude is a more convenient characteristic since its units are the same of the variable in a real space. Fig. \ref{fig:p2d_a2d} shows the amplitude of SB fluctuations in the central $3$ arcmin region and in the outer annulus $3-10.5$ arcmin. High statistics of counts allows us to probe the SB fluctuations on a broad range of length scales - more than an order of magnitude, down to the scales comparable with the mean free path (see Fig. \ref{fig:prof}). The amplitude varies from $\sim 3-14$ per cent on scales $2-50$ kpc in the central $3$ arcmin to $\sim 2-33$ per cent on scales $\sim 14-170$ kpc in the $3-10.5$ arcmin region. The fact that the amplitudes in both regions differ at least by a factor of 3 over a broad range of scales shows that the amplitude of SB fluctuations in Perseus varies with radius. Therefore, instead of global characterization of the fluctuations in the cluster core, we consider below the fluctuations in a set of narrow radial annuli.
\section{Results}
\subsection{Power spectrum of density fluctuations}
\label{sec:den_fluct}
We divided the cluster into a set of radial annuli with a width $1.5'$ ($\approx 31$ kpc) (Fig. \ref{fig:im_pers}) and calculated the PS of SB fluctuations in each annulus, subtracting the Poisson noise and correcting for the unresolved point sources and the PSF. The latter one was done using PSF PS shown in Fig. \ref{fig:psf}.
\begin{figure}
\begin{minipage}{0.49\textwidth}
\includegraphics[trim=0 150 0 75,width=1\textwidth]{psf_mean_v1.pdf}
\end{minipage}
\caption{Power spectrum (PS) of the combined {\it Chandra} PSF obtained in each annulus of the Perseus Cluster in our analysis (Fig. \ref{fig:im_pers}). The combined PSF accounts for the shape of the PSF at different offsets and exposure of each observation (see Section \ref{sec:init}). Each spectrum is averaged over 10 spectra of different realizations of the PSF map. Since the PSF is not a $\delta-$function, the PS deviate significantly from unity (dotted line), especially on scales smaller than 15 kpc ($k\approx0.02$ arcsec$^{-1}$).
\label{fig:psf}
}
\end{figure}
\begin{figure}
\begin{minipage}{0.49\textwidth}
\includegraphics[trim=0 250 0 200,width=1\textwidth]{a3d_obs_200.pdf}
\end{minipage}
\caption{Amplitude of density fluctuations in the inner 200 kpc (radius) region in the simulated relaxed
galaxy cluster (CL21 cluster with a total mass $M_{500c}=6.08 \cdot 10^{14} M_{\odot}$ at $r_{500c}=1215.2$ kpc, see \citet{Zhu14a} for details) obtained from the 3D density information (dash navyblue curve) and recovered
from the image of the X-ray surface brightness (hatched orange/red region).
The width of the hatched region reflects the uncertainty of
the recovered 3D amplitude of density fluctuations due to variations
of the converting geometrical factor with the projected radius. Notice
that the true (from the 3D data) amplitude of density fluctuations is within
the hatched region. This confirms that the observational procedure
of measuring the amplitude of density fluctuations from the X-ray
images recovers reasonably accurate the amplitude of the 3D density
fluctuations.
\label{fig:2dto3d_sim}
}
\end{figure}
The fact that the X-ray SB $I_X\sim \int n_e^2 dl$ for hot clusters with $T_e > 2$ keV (see Fig. \ref{fig:fluxes}) allows us to convert the 2D PS of SB fluctuations into the 3D PS of density fluctuations on scales smaller than the length of the line of sight. The procedure is described in \citet{Chu12}. Namely, knowing the global spherically-symmetric model of the gas emissivity, the conversion factor between the 3D and 2D power spectra at each line of sight $z$ is
\be
\frac{P_{2D}(k)}{P_{3D}(k)}\approx 4 \int |W(k_z)|^2dk_z,
\label{eq:p2dp3d}
\ee
where $W(k_z)$ is the PS of the normalized emissivity distribution along the line of sight.
We tested this procedure using high-resolution cosmological simulations of galaxy clusters \citep{Nag07,Nel14}. Using non-radiative simulations, we obtained the PS of density fluctuations directly from the simulated 3D data for a sample of 6 relatively relaxed clusters and recovered the PS from projected images of the X-ray SB (applying the same procedure we use for the observed X-ray images). The details of the sample and the analysis are described in \citet{Zhu14a}. Direct comparison of both PS (or amplitudes) confirms that the X-ray images can provide reasonably accurate measurements of the PS of density fluctuations. Fig. \ref{fig:2dto3d_sim} shows an example of one of the clusters in our sample.
The accuracy of the recovery of the 3D spectrum in each annulus depends on how steep the emissivity profile is. In case of the Perseus Cluster, the emissivity is strongly peaked towards the center, therefore the conversion factor significantly varies within the radial annuli. The narrower the annulus, the smaller the uncertainty. To do the conversion, we use the mean (within the annulus) value of this factor. In Section \ref{sec:uncert} we will discuss the uncertainties associated with this step of the analysis.
Fig. \ref{fig:a3d_r} shows the amplitude of density fluctuations $A_{3D}$ obtained in each annulus from the PS of SB fluctuations $P_{2D}$ as a function of wavenumber $k$. Scales, which are smaller than the width of each annulus ($\approx 31$ kpc) are plotted. Hatched regions show the amplitude on wavenumbers, over which we deem the measurements are least affected by the uncertainties and more robust against the observational limitations. The choice of the high-$k$ limit is set by the statistical uncertainty or by the PSF distortions of the amplitude. For both cases, the uncertainty is less than 20 per cent in this hatched region. At low $k$, we exclude scales, on which the amplitude flattens or decreases towards smaller $k$. This flattening is most likely determined by the presence of several characteristic length scales (e.g. the distance from the center, scale heights, additional driver of fluctuations) and by the uncertainties in the choice of the underlying cluster model (see Section \ref{sec:uncert}).
\begin{figure*}
\includegraphics[trim=0 150 27 60,width=0.49\textwidth]{a3d_all_perseus_useituseit.pdf}
\includegraphics[trim=-40 125 0 145,width=0.49\textwidth]{a3d_r.pdf}
\caption{{\bf Left:} amplitude of density fluctuations in the Perseus
Cluster versus wavenumber calculated in a set of radial annuli shown in Fig. \ref{fig:im_pers}. Hatched regions show the amplitude on scales where the measurements are least affected by systematic and statistical uncertainties. The width of each curve reflects estimated $1\sigma$ statistical and stochastic
uncertainties. {\bf Right:} radial profiles of the amplitude of density fluctuations measured on
certain scales (see the legend). Notice, that the amplitude decreases
with the radius and increases with the length scale of fluctuations.
\label{fig:a3d_r}
}
\end{figure*}
The amplitude of density fluctuations is less than $15$ per cent, except for the central $\sim 30$ kpc. Table \ref{tab:denvel} summarizes the characteristic values of the amplitude in all considered regions. The amplitude in the innermost region is noticeably higher and the curve itself is flatter than in other regions. This is not surprising, since a significant fraction of the volume of this region is dominated by sharp edges around bubbles of relativistic plasma, shocks around them, filamentary structures and absorption features, flattening the high-$k$ end of the amplitude. Simple tests showed the drop of power on small scales when these features are excluded from the analysis. The radial variations of the amplitude at different scales is shown in the right panel of Fig. \ref{fig:a3d_r}. Notice, that the amplitude decreases with the distance from the center and it increases with the scale (characteristic size) of the fluctuations.
\begin{figure*}
\begin{minipage}{0.49\textwidth}
\includegraphics[trim=0 150 27 60,width=1\textwidth]{v1d_perseus_useituseit_forpap.pdf}
\end{minipage}
\begin{minipage}{0.49\textwidth}
\includegraphics[trim=-45 150 0 40,width=1\textwidth]{v1d_r.pdf}
\end{minipage}
\caption{{\bf Left:} amplitude of one-component velocity of gas
motions versus wavenumber $k=1/l$, measured in a set of radial annuli (see the legend) in the Perseus Cluster. The velocity at each scale is obtained from the amplitude of density fluctuations, shown in Fig. \ref{fig:a3d_r}, using relation \ref{eq:rho_v}. The color-coding and notations are the same as in
Fig. \ref{fig:a3d_r}. The slope of the amplitude for pure Kolmogorov turbulence \citep{Kol41}, $k^{-1/3}$, is shown with dash line. {\bf Right:} radial profiles of one-component velocity amplitude measured on certain length scales written in the legend.
\label{fig:v1d_r}
}
\end{figure*}
\begin{figure}
\begin{minipage}{0.49\textwidth}
\includegraphics[trim=0 150 0 90,width=1\textwidth]{turb_scales_perseus.pdf}
\end{minipage}
\caption{Characteristic length scales present in the Perseus Cluster at different distances $R$ from the center. Colored shaded regions: range of scales, on which our measurements are least affected by systematic and statistical uncertainties (hatched regions in Fig. \ref{fig:a3d_r} and Fig. \ref{fig:v1d_r}). Solid curve: the Ozmidov scale $l_O$ of turbulence in the core of Perseus, obtained assuming local balance between turbulent heating and radiative cooling (relation \ref{eq:ozm}). We do not plot $l_O$ in central $\sim 20$ kpc, since the measured gas entropy is flat there, leading to $l_O\to\infty$. Dash curve: the Kolmogorov (dissipation) scale $l_K$ for unmagnetized plasma (relation \ref{eq:kolm}). Dot-dash curve: the electron mean free path for unmagnetized plasma. See Section \ref{sec:vel_ps} for discussion.
\label{fig:scales}
}
\end{figure}
\begin{table*}
\centering
\caption{The mean values of the amplitude of density fluctuations and the amplitude of one-component velocity in the Perseus Cluster (see Fig. \ref{fig:a3d_r} and \ref{fig:v1d_r}). The values are given on scales $l=1/k$: 10, 20 and 30 kpc.}
\begin{tabular}{@{}lcccccc@{}}
\hline
Annulus & \multicolumn{3}{c}{Amplitude of density fluctuations} & \multicolumn{3}{c}{Amplitude of one-component velocity}\\
& \multicolumn{3}{c} {$A_{3D}=\delta\rho/\rho$ } & \multicolumn{3}{c} {$V_{1,k}$, km/s}\\
\hline
& $l=10$ kpc & $l=20$ kpc & $l=30$ kpc & $l=10$ kpc & $l=20$ kpc & $l=30$ kpc \\
\hline
0\' \,- 1.5\' \,(0 - 31 kpc) & 0.22 & - & - & 200 & - & -\\
1.5\' \,- 3\' \,(31 - 62 kpc) & 0.11 & 0.14 & - & 110 & 140 & -\\
3\' \,- 4.5\' \,(62 - 94 kpc) & 0.08 & - & - & 90 & - & -\\
4.5\' \,- 6\' \,(94 - 125 kpc) & 0.09 & 0.1 & 0.11 & 110 & 130 & 140\\
6\' \,- 7.5\' \,(125 - 156 kpc) & - & 0.09 & 0.12 & - & 120 & 150\\
7.5\' \,- 9\' \,(156 - 187 kpc) & - & 0.08 & 0.09 & - & 100 & 120\\
9\' \,- 10.5\' \,(187 - 219 kpc)& - & - & 0.08 & - & - & 100\\
\hline
\label{tab:denvel}
\end{tabular}
\end{table*}
\subsection{Velocity power spectrum}
\label{sec:vel_ps}
Knowing the amplitude of density fluctuations on each length scale, the amplitude (and the PS) of velocities of gas motions present in the ICM can be obtained. Simple theoretical arguments show that in the stratified atmospheres of relaxed galaxy clusters, the amplitude of density fluctuations $A_{3D}(k)=\disp\frac{\delta\rho_k}{\rho_0}$ and one-component velocity $V_{1,k}$ are proportional to each other at each length scale $l=1/k$ within the intertial range of scales, namely
\be
\frac{\delta\rho_k}{\rho_0}=\eta_1\frac{V_{1,k}}{c_s},
\label{eq:rho_v}
\ee
where $\rho_0$ is the mean gas density and $\eta_1$ is the proportionality coefficient \citep{Zhu14a}. If the injection scale of the turbulence is larger than or comparable with the Ozmidov scale \citep{Ozm92,Bre07}, the coefficient $\eta_1$ is set by gravity-wave physics on large, buoyancy-dominated scales, and is $\sim 1$ in atmospheres of galaxy clusters. It remains the same on smaller scales (within the inertial range) where the density becomes a passive scalar. Cosmological simulations of galaxy clusters confirm the relation \ref{eq:rho_v}, giving the averaged over a sample of relaxed clusters value $\eta_1=1\pm0.3$ \citep{Zhu14a}.
Fig. \ref{fig:v1d_r} shows the amplitude of one-component velocity versus wavenumber obtained from the relation \ref{eq:rho_v} in all seven radial annuli in the Perseus core, as well as the radial profiles of the amplitude on certain scales. Like in Fig. \ref{fig:a3d_r}, hatched regions show the amplitude on scales, where the measurements are least affected by systematic and statistical uncertainties. Notice that the velocity amplitude is higher towards the center, suggesting a power injection from the center. The spectra (amplitudes) show similar dependence on $k$, larger fluctuations are on larger $k$, consistent with cascade turbulence. The slope of the velocity PS is broadly consistent with the slope for canonical Kolmogorov turbulence, accounting for the errors and uncertainties. Within the ``robust'' range of scales, $V_{1,k}$ varies from $\sim 70$ km/s up to $\sim 210$ km/s on scales $\sim 5-30$ kpc (see Table \ref{tab:denvel}). The velocity amplitudes quantitatively match our expectations of typical velocities in the ICM from various observational constraints \citep[see e.g.][and references therein]{Chu04,Sch04,Wer09,San10,San11,deP12,San13,Pin15} and numerical simulations \citep[see e.g.][and references therein]{Nor99,Dol05,Iap08,Lau09,Vaz11,Min14}. Future direct measurements of the velocity of gas motions with X-ray calorimeter on-board {\it Astro-H} observatory \citep{Tak14} will allow us to test the density-velocity relation \ref{eq:rho_v}. Recently-proposed observational strategy for the Perseus Cluster maps the cluster core in radial and azimuthal directions, enabling us to do the calibration \citep{Kit14}.
\begin{figure*}
\begin{minipage}{0.49\textwidth}
\includegraphics[trim=0 200 -70 150,width=1\textwidth]{nobub.pdf}
\end{minipage}
\begin{minipage}{0.49\textwidth}
\includegraphics[trim=0 170 30 80,width=1\textwidth]{v1d_perseus_useituseit_forpap_nobub.pdf}
\end{minipage}
\caption{{\bf Left:} residual image of the Perseus Cluster in the $0.5-3.5$ keV band (the same as Fig. \ref{fig:im_pers}) with excluded bubbles of relativistic plasma taken from \citet{Fab11} (black ovals). {\bf Right:} one-component velocity amplitude versus wavenumber obtained from the image on the left. The color-coding and labels are the same as in Fig. \ref{fig:v1d_r}.
\label{fig:nobub}
}
\end{figure*}
It was recently shown that in the cores of Perseus and Virgo clusters, where the cooling time is shorter than the Hubble time, the heating of the gas due to dissipation of turbulence is sufficient to offset radiative cooling losses \citep{Zhu14b}. Assuming that the dissipation rate is matching the cooling rate, it is straightforward to estimate the Ozmidov scale $l_O$ of the turbulence using only thermodynamic properties of the Perseus Cluster. Namely,
\be
l_O=N^{-3/2}\varepsilon^{1/2}=N^{-3/2}\disp\left(\frac{Q_{cool}}{\rho}\right)^{1/2},
\label{eq:ozm}
\ee
where $N=\disp\sqrt\frac{g}{\gamma H_s}$ is the Brunt-V$\rm\ddot a$is$\rm\ddot a$l$\rm\ddot a$ frequency in the cluster atmosphere, $g$ is the acceleration of gravity, $H_s=\disp\left(\frac{d{\rm ln}S}{dr}\right)^{-1}$ is the entropy scale height, and $\varepsilon$ is the dissipation rate per unit mass. Here we assume $\rho\varepsilon\sim$ the cooling rate $Q_{cool}=n_e n_i \Lambda_n(T)$, where $n_e$ and $n_i$ are the number densities of electrons and ions, respectively, and $\Lambda_n(T)$ is the normalized gas cooling function \citep{Sut93}.
Fig. \ref{fig:scales} shows the radial profile of the Ozmidov scale $l_O$ and a range of scales we are probing in each annulus (hatched regions in Fig. \ref{fig:v1d_r}). One can see that $l_O$ is within the interval of scales we are probing at each distance from the center within the cluster core. This means that the main assumption behind the derivation in \citet{Zhu14a} is satisfied, namely the Ozmidov scale is smaller than the injection scale of turbulence (assuming that we are probing velocity PS within the inertial range). Notice, that we do not show $l_O$ at $R<20$ kpc since the measured gas entropy is flat towards the center, leading to $l_O\to\infty$.
It is also interesting to compare the scales we are probing with the Kolmogorov (dissipation) scale
\be
l_K=\disp\frac{\nu_{kin}^{3/4}}{(Q_{cool}/\rho)^{1/4}},
\label{eq:kolm}
\ee
where $\nu_{kin}=\disp\frac{\nu_{dyn}}{\rho}$ is the kinematic viscosity, which is obtained through the dynamic viscosity $\nu_{dyn}$ for an ionized plasma without magnetic field. Fig. \ref{fig:scales} shows the Kolmogorov scale $l_K$ as well as the mean free path for comparison. The Kolmogorov scale is significantly below the scales we are probing, which justifies even better our assumption about the inertial range of scales. In reality, the situation can be more complicated, e.g., due to the presence of magnetic fields. Here, we neglect these complications, following the simplest approach as a good starting point.
We cannot claim, of course, that the structures in the SB are due to turbulence only. For example, there are many bubbles of relativistic plasma in the Perseus core \citep[see e.g.][]{Boe93,Chu00,Fab00}, which may contribute to the signal. The question is whether their contribution is dominant in the considered regions. Various sharp structures (e.g. bubble edges) would give the slope of the amplitude $k^{-1/2}$, which is hard to discriminate from the Kolmogorov slope $k^{-1/3}$ accounting for the uncertainties of our measurements and the assumptions used. However the contribution of the bubbles to the signal can be easily seen if we repeat the analysis excluding the known bubbles from the image of Perseus. Fig. \ref{fig:nobub} shows the one-component velocity amplitude calculated from such image. Notice, that the velocity amplitude decreases by a factor $\sim 1.6-1.2$ depending on scale in the central $1.5$ arcmin. This is expected since bubbles in this region are particularly prominent and occupy a substantial fraction of the volume. In the $1.5-3$ arcmin annulus, this factor is $\sim 1.15$ over the whole range of scales. Outside $3$ arcmin, the exclusion of bubbles from the analysis does not change the measured velocity amplitude.
Sound waves \citep{Fab06,Ste09}, mergers and gas sloshing \citep{Mar07} might also contribute to the observed spectrum of the density fluctuations. Unsharp masking of the Perseus Cluster revealed quasi-spherical structures (``ripples'') in the SB, which have been interpreted as isothermal sound waves \citep{Fab00,Fab03}. An alternative interpretation is stratified turbulence, which arises naturally in the cluster atmosphere, where rough estimates give Froude number $Fr\sim 0.3-1$ \citep{Zhu14a}. In this case, the radial size of each ``ripple'' $\Delta r$ is determined by $HV/c_s$, where $H$ is the characteristic scale height and $V$ is the velocity amplitude. For example, in the $1.5-3$ arcmin annulus in Perseus, the scale height $H$ is $\sim 40-70$ kpc, the sound speed $c_s\sim 900$ km/s and the velocity measured from the SB fluctuations is $\sim 100-140$ km/s. This gives us typical radial sizes of the fluctuations $\sim 5-10$ kpc, which are consistent with those seen in the X-ray image of the cluster once the large-scale asymmetry in SB is removed. Certainly, we cannot claim that the observed fluctuations are associated with the stratified turbulence only. The question on the nature of the fluctuations will be addressed in our future work. At the very least, the measured turbulent velocities can be treated as an upper limit.
\subsection{Gas clumping}
\label{sec:clump}
\begin{figure}
\includegraphics[trim=0 170 0 80,width=0.45\textwidth]{clumping_paper.pdf}
\caption{Radial profile of the clumping factor $C-1$ in the core of the Perseus Cluster obtained from the relation (\ref{eq:clk}) using the measurements of density fluctuations power spectra. {\bf Red points}: the measured density spectrum is fitted by a power-law function with the Kolmogorov slope $-11/3$ and extended to small and large scales, which we do not probe in our analysis. {\bf Blue points:} the slope of the power-law function is assumed $-3.1$ and $-4$ for the innermost annulus $0-1.5$ and for other six regions respectively. The slope $-4$ is obtained from the power spectrum of density fluctuations in the broad annulus $1.5-10.5$ arcmin. {\bf Dash points}: $k_{min}=1/R$, $k_{max}=\infty$. {\bf Solid points:} $k_{min}=0.01$ arcmin$^{-1}$ (inverse width of each annulus), $k_{max}=\infty$. {\bf Gray dotted points:} clumping factor obtained from the measured power spectrum of density fluctuation directly, avoiding any approximations with the power-laws. Measurements within the range of scales least affected by the uncertainties (hatched regions in Fig. \ref{fig:a3d_r}) are used. The corresponding scales are written below each point. Notice, that the clumping factor is less than $7-8$ per cent outside the central $\sim 30$ kpc.
\label{fig:clumping}
}
\end{figure}
The term ``gas clumping'' refers to any deviations of gas density isosurfaces from equipotential surfaces and, therefore, includes both large-scale inhomogeneities in the gas and small-scale clumps\footnote{Less general definitions are often used in astrophysics. For example, clumping can be referred to gas clumps and subhalos only, which are notably denser and cooler than the ambient gas.}. It is difficult to determine or even define unambiguously the surfaces of constant potential in clusters. Therefore, in practice, we measure clumping relative to a model, which describes a global gas distribution in the cluster potential well. Namely, the gas clumping factor is usually defined as
\be
C=\disp\frac{\langle (\rho/\rho_0)^2\rangle}{\langle \rho/\rho_0\rangle ^2},
\label{eq:cl}
\ee
where $\langle\rangle$ denotes the mean or median inside a spherical shell and $\rho_0$ is the global density model, which accounts for the density gradient within each shell.
Gas clumping leads to an overestimate of the gas density and, as a consequence, affects the gas mass and total hydrostatic mass measurements of galaxy clusters \citep[see e.g.][]{Lau09} as well as the SZ measurements of the Compton parameter \citep[e.g.][]{Khe13}. Numerical simulations predict clumping $< 1.05$ in central $0.5r_{500}$, $\sim 1.1-1.4$ at $r_{500}$ and up to $2$ at $r_{200}$ \citep[see e.g.][]{Mat99,Nag11,Ron13,Zhu13,Vaz13}. The clumping factor obtained from X-ray observations of cluster outskirts is slightly higher and varies from $\sim 2$ to $\sim 3$ at $r_{200}$ \citep[see e.g.][and references therein]{Urb14,Wal12,Mor14}. The main challenges of the clumping measurements from the X-ray observations are a very low X-ray SB in cluster outskirts, where clumping is expected to be higher, and, in contrast, a low value of the clumping factor, less than 5 per cent according to numerical simulations, in the bright cluster cores.
\begin{figure*}
\includegraphics[trim=40 270 45 250,width=1\textwidth]{imaged.pdf}
\caption{{\bf Top:} underlying models of ``unperturbed'' surface brightness in the Perseus Cluster. From left to right: spherically-symmetrical $\beta-$model (default choice), patched $\beta-$models with $\sigma=80$ arcsec (removes large-scale asymmetry), $\sigma=30$ arcsec and $\sigma=10$ arcsec (see Section \ref{sec:und_mod} for the details). {\bf Bottom:} residual images of the SB fluctuations in Perseus obtained from the initial image divided by the underlying model (top panels). The smaller the $\sigma$ the smaller the structures included to the model and the less structures remain in the residual image.
\label{fig:imaged}
}
\end{figure*}
Presenting the density $\rho$ through unperturbed $\rho_0$ and fluctuating $\delta\rho/\rho_0$ parts, $\rho=\rho_0(1+\delta\rho/\rho_0)$, the gas clumping factor definition (\ref{eq:cl}) can be re-written through the PS of density fluctuations as
\be
C=1+\left(\disp\frac{\delta\rho}{\rho_0}\right)^2=1+\int\limits_{k_{min}}^{k_{max}}4\pi P_{3D}(k)k^2dk,
\label{eq:clk}
\ee
where the amplitude of density fluctuations $\disp\frac{\delta\rho}{\rho_0}$ is defined through the PS as $\disp\left(\frac{\delta\rho}{\rho_0}\right)^2=\int 4\pi P_{3D}(k)k^2dk$ and the integration is over all wavenumbers, i.e. from $k_{min}=0$ till $k_{max}=\infty$. Using this relation and our measurements of the PS of density fluctuations in the Perseus core, we estimate the gas clumping factor, which is shown in Fig. \ref{fig:clumping}. For each annulus we obtain $k_{min}$ and $k_{max}$ as the smallest and the largest wavenumbers within the range of wavenumbers least affected by uncertainties (hatched regions in Fig. \ref{fig:a3d_r}), see dotted gray points in Fig. \ref{fig:clumping}. The gas clumping is very low, less than 2 per cent at $R>\sim 30$ kpc, which is expected since we are integrating over a narrow range of wavenumbers. In order to estimate the total clumping, we approximated the PS of density fluctuations with power-law functions and extended them up to $k_{max}=\infty$ and down to $k_{min}=0.01$ arcmin$^{-1}$ (inverse width of each annulus, solid points in Fig. \ref{fig:clumping}) or down to $k_{min}=1/R$ (dash points in Fig. \ref{fig:clumping}), where $R$ is the distance from the center to each annulus. Red points in Fig. \ref{fig:clumping} show the gas clumping in case when we fit a power-law with the Kolmogorov slope $-11/3$ to the measured spectra, varying only its normalization in each annulus. For the blue points, we assume the slope being consistent with the slope of the global density spectrum obtained in the broad $1.5-10.5$ arcmin annulus. Namely, for annuli at $R>30$ kpc ($1.5$ arcmin), the slope of the spectrum is $-4$, while for the innermost annulus, the slope is $-3.1$. Notice, that the gas clumping is dominated by the large scales; it decreases with the distance from the center if we do not account for fluctuations on scales larger than the width of each annulus (it simply reflects the fact that the amplitude of fluctuations decreases with the radius, see Fig. \ref{fig:a3d_r}); outside the central $\sim 30$ kpc, the gas clumping is lower than $7-8$ per cent, which leads to a density bias $\sim (C-1)/2$ less than $3-4$ per cent.
\section{Uncertainties in the analysis}
\label{sec:uncert}
In this Section, we examine various systematic uncertainties in the analysis. Namely, we consider uncertainties associated with
\begin{itemize}
\item the choice of the unperturbed cluster density model;
\item the bias of the $\Delta-$variance method used for the PS calculations;
\item inhomogeneous exposure coverage;
\item conversion $P_{2D}$ to $P_{3D}$;
\item velocity measurements.
\end{itemize}
Even though the list of uncertainties is, admittedly, long, a conservative estimate of the total uncertainty on the amplitude of density fluctuations (one-component velocity) is less than $50$ ($60$) per cent.
\subsection{Underlying model of the surface brightness}
\label{sec:und_mod}
\begin{figure}
\begin{minipage}{0.49\textwidth}
\includegraphics[trim=0 170 0 90,width=1\textwidth]{a3d_perseus_3_45_smor.pdf}
\end{minipage}
\caption{Amplitude of density fluctuations in the $3-4.5$ arcmin annulus in the Perseus Cluster obtained by using spherically-symmetrical $\beta-$model as the underlying one (red) and more flexible models shown in Fig. \ref{fig:imaged}. Notice, that the removal of the global cluster asymmetry (blue hatched region) does not affect the amplitude measurements relative to the default model (red) on scales we are probing. Also, notice that even if the amplitude suppression is present on scales, which are included to the underlying model (as expected), the amplitude on smaller scales, which are not affected by the model, remains almost the same. This means that the measured amplitude is indeed due to the presence of fluctuations of these scales and not due to the power leakage from the large scales.
\label{fig:ampl_bmod}
}
\end{figure}
Decomposition of the cluster image into ``perturbed'' and ``unperturbed'' components is ambiguous. Physically motivated, the underlying model of the ``unperturbed'' component should reflect the global potential of the cluster in equilibrium, i.e. when the isosurfaces of the gas density and temperature are aligned with the equipotential surfaces. Our default choice of the model - $\beta-$model for the azimuthally averaged SB (or just mean SB) - seems to be a reasonable choice of such ``unperturbed'' model at least for (or close to) relaxed galaxy clusters. However, one can see the large-scale, west - east asymmetry in the SB of the Perseus Cluster (Fig. \ref{fig:im_pers}, left), which might question our most simple and conservative choice. To check how the global cluster asymmetry affects the density amplitude measurements, we repeated the analysis, using two other $\beta-$models of the SB averaged in $180^{\circ}$ sectors. The best-fitting parameters of the $\beta-$model in the east ($90^{\circ}\div270^{\circ}$) and the west ($-90^{\circ}\div90^{\circ}$) sectors are $r_c=1.76$ and $0.89$ arcmin, $\beta=0.61$ and $0.48$, respectively. Choosing these underlying models, the maximal deviations of the density amplitude from the default case are on the largest scales in each annulus (typically $\sim 30-20$ kpc) and are less than $10$ per cent in the central 6 arcmin, less than $17$ per cent in the $6-7.5$ arcmin annulus and less than $30$ per cent at the distance $7.5-10.5$ arcmin from the cluster center.
Going beyond this simple spherically-symmetric model implies that we believe that the underlying cluster potential is more intricate. It is not clear to what degree of complexity of the model we should go. There is always a danger that some of the structures unrelated to the cluster gravitational potential are removed. We made a number of tests, introducing more flexibility to the $\beta-$model, by patching it on large scales to account for possible complexity of the potential. Namely, our patched model is defined as $I_{pm}=I_{\beta} S_{\sigma}[I_{X}/I_{\beta}]$, where $I_{\beta}$ is the $\beta-$model of azimuthally-averaged SB, $S_{\sigma}[\cdot]$ denotes Gaussian smoothing with the smoothing window size $\sigma$ and $I_X$ is the clusters X-ray SB. Varying the size of the smoothing window, the model changes from most conservative symmetrical $\beta-$model (large $\sigma$) to most complicated (and clearly implausible) one, which accounts for structures on all scales (small $\sigma$) \footnote{An alternative way to remove large-scale structure is used in the analysis of SB fluctuations in the AWM7 cluster \citep{San12}. The cluster is modeled by fitting ellipses to contours of SB, spaced logarithmically.}. Examples of such models and the corresponding residual images of the SB fluctuations are shown in Fig. \ref{fig:imaged}. Models with large $\sigma$ (e.g. $\sim 80$ arcsec) remove the large-scale (east-west) asymmetry of the cluster, while those with a smaller $\sigma$ absorb more features of the image on smaller scales. Fig. \ref{fig:ampl_bmod} shows the amplitude of density fluctuations in $3-4.5$ arcmin annulus measured from the residual images in Fig. \ref{fig:imaged}. Notice, that the removal of the large-scale asymmetry (blue hatched region) does not change the amplitude of density fluctuations relative to spherically-symmetric $\beta-$model (red hatched region) over the range of scales probed in our analysis ($< 30$ kpc). As expected, the more complex models (with the smaller $\sigma$), the broader range of scales, on which the amplitude is suppressed, and the stronger the suppression of the amplitude on the largest scales.
One can estimate the scales, on which the amplitude is expected to be suppressed depending on the size of the window function $\sigma$ used in the patched underlying models. Assuming that the PS of density fluctuations is a power-law $k^{-\alpha}$, the convolution of the image PS with the Mexican Hat filter will give \citep[see relation A8 in][]{Are12}
\be
P_{\rm{mh}}(k_r)\propto \int k^{-\alpha}\left(\frac{k}{k_r}\right)^4 e^{-2(k/k_r)^2} d^nk,
\label{eq:fil1}
\ee
where $k_r$ is the characteristic wavenumber. The convolution of the PS of the image, with removed large-scale part smoothed with a Gaussian, is
\be
P_{\rm{mh,\sigma}}(k_r)\propto\int k^{-\alpha}\left(\frac{k}{k_r}\right)^4e^{-2(k/k_r)^2} \left(1-e^{-2\pi ^2 k^2 \sigma^2}\right)^2 d^nk.
\label{eq:fil2}
\ee
The ratio of both will give us an estimate of a wavenumber, on which the amplitude will be suppressed for any size $\sigma$ of the window function used for the underlying model of the SB. For the Kolmogorov PS, $k^{-11/3}$, the suppression of the amplitude obtained using the patched $\beta-$model relative to the amplitude measured using the spherically-symmetric model is $\sim 20$ per cent on $k\approx 0.8/\sigma$. This is roughly consistent with what we see in Fig. \ref{fig:ampl_bmod}. For example, the underlying patched model with $\sigma=20$ arcsec gives a 20 per cent difference in the amplitudes of density fluctuations on $k_{\rm char}\approx0.04$ arcsec$^{-1}$. On $k<k_{\rm char}$ the ``patched'' amplitude is strongly suppressed, while on $k>k_{\rm char}$ the amplitudes remain almost the same. This means that at $k>k_{\rm char}$ the measured amplitude of density fluctuations is due to the presence of fluctuations on these scales and not due to the leakage of power from the large scales. The net conclusion is that the amplitude of density fluctuations measured on scales $< 30$ kpc is almost not affected by the choice of the underlying model, unless the cluster potential is very disturbed.
\subsection{Bias in the $\Delta-$variance method}
The normalization of the PS obtained through the $\Delta-$variance method may be biased slightly, depending on the slope of the PS \citep[see Appendix B in][]{Are12}. The approximations of the initial PS of SB fluctuations with a power-law functions give the slopes $\sim -3 - -3.8$ depending on the distance from the cluster center. Therefore, the normalization of the measured amplitude of density fluctuations is on average overestimated by $\sim 20-30$ per cent and slightly lower, by $\sim 10$ per cent, in the innermost ($0-1.5$ arcmin) and outermost ($9-10.5$ arcmin) annuli.
\subsection{Inhomogeneous exposure coverage}
\label{sec:inhom_exp}
The exposure map is not uniform and the brightness of the cluster itself varies across each annulus as seen in Fig. \ref{fig:im_pers} and \ref{fig:psf_exp}. Both arguments may bias the measured amplitude of density fluctuations. By default, when calculating the amplitude by taking RMS of fluctuations present in filtered images, we use most uniform weighting scheme, $w=1$, i.e. we treat all pixels in the image with the same weight. The measured amplitude is then more sensitive to fluctuations close to the outer edge of each annulus. The least uniform scheme, but at the same time most optimal for the reduction of the Poisson noise requires weights to be $w_1\propto t_{exp}I_{mod}$, where $t_{exp}$ is the exposure map and $I_{mod}$ is the underlying model of the SB. In this case, those parts of the cluster that have higher number of counts would have larger weights. We also experimented with the weight $w_2\propto t_{exp}$, which is more sensitive to the deepest-exposure parts of the image. We find $<15$ per cent ($<30$ per cent) higher (lower) value of the amplitude of density fluctuations on scales $< 20$ kpc if weights $w_1$ and $w_2$ are used. On larger scales, $\sim 30$ kpc, the amplitude can be $< 60$ per cent lower. As an example, Fig. \ref{fig:syst_w_tw2} shows the amplitude in $1.5-3$ arcmin annulus measured using three different weighting schemes.
\begin{figure}
\begin{minipage}{0.49\textwidth}
\includegraphics[trim=0 160 -10 60,width=1\textwidth]{a3d_uncert.pdf}
\end{minipage}
\caption{Amplitude of density fluctuations in $1.5-3$ arcmin annulus in the Perseus Cluster measured varying the weight (top panel, see Section \ref{sec:inhom_exp}) and the conversion factor $f_{2D\to3D}$ (bottom panel, see Section \ref{sec:con2d3d}). The default choice is shown with navyblue hatched region.
\label{fig:syst_w_tw2}
}
\end{figure}
\subsection{Conversion from 2D to 3D power spectrum}
\label{sec:con2d3d}
We convert the two-dimensional PS of the SB fluctuations $P_{2D}$ into the three-dimensional PS of density fluctuations $P_{3D}$ using relation (\ref{eq:p2dp3d}). The conversion factor $f_{2D\to 3D}=P_{2D}/P_{3D}$ depends on a distribution of fluctuations along the line of sight and its value varies with projected distance from the cluster center. If radial profile of the SB is steep, the uncertainty can be large unless narrow annuli are considered. We use the value of $f_{2D\to 3D}$ evaluated at the mean radius of each annulus. This factor is different in inner and outer edges of each annulus, leading to a maximal uncertainty on the measured amplitude $<17$ per cent. Fig. \ref{fig:syst_w_tw2} shows the amplitude in $1.5-3$ arcmin annulus obtained, accounting for the variations of the conversion factor $f_{2D\to 3D}$ within the annulus.
\subsection{Uncertainties in velocity measurements}
The proportionality between the amplitude of density fluctuations and velocity with the coefficient $\eta\sim 1$ is based on the simplest approach which is a good approximation (as confirmed by numerical simulations) that captures the key physics. The derivation neglects gas heating, cooling, heat fluxes, magnetic fields and assume that all perturbations are small. Magnetic fields on large scales should not significantly modify buoyancy physics since $\beta=8\pi nkT/B^2 >>1$. On small scales, the density fluctuations would behave as a passive scalar \citep{Sch09}. Therefore, we expect the linear relation to hold even in MHD case, however the proportionality coefficient is likely to change in this case. The {\it Astro-H} X-ray observatory will allow us to verify and calibrate the relation. Any deviation from it will indicate the importance of the neglected physics.
The intermittency of density fluctuations and velocity can be admittedly large in the Perseus Cluster. For example, analyzing the fluctuations in small patches within the $3-4.5$ arcmin annulus, we noticed that the amplitude of density fluctuations varies in a relatively broad range from 3 to 10 per cent. In order to achieve statistical convergence, we perform our measurements in relatively wide annuli. The amplitude obtained in the twice broader annuli is consistent with the amplitudes measured in two individual annuli.
\section{Conclusions}
\label{concl}
We performed detailed analysis of the X-ray SB and gas density fluctuations in a set of radial annuli within the core (central $r\sim 220$ kpc) of the Perseus Cluster, using deep {\it Chandra} observations.
To summarize our findings:
\begin{itemize}
\item The characteristic amplitude of the density fluctuations varies from $8$ to $12$ per cent on scales $\sim 10-30$ kpc within $30-160$ kpc annulus and from $ 9$ to $7$ per cent on scales $\sim 20-30$ kpc in the outer annuli, $160-220$ kpc. The amplitude in the innermost $30$ kpc is higher, up to $20-22$ per cent on scales $\sim 5-15$ kpc. The higher amplitude in this region reflects the presence of bubbles, shocks and sound waves around them, filaments and absorption feature, which occupy a large area of the considered region. The smallest scale we probe varies from $\sim 5$ kpc in the central $30$ kpc to $\sim 25$ kpc at distance $\sim 200$ kpc from the center (while the mean free path for unmagnetized plasma is $\sim 0.1$ and $5$ kpc, respectively).
\item Given stratification and gas entropy gradient in the atmosphere of the cluster, we use linear relation between the amplitude of density fluctuations and velocity of gas motions to evaluate the characteristic velocity amplitude on different scales. The typical amplitude of the one-component velocity outside the central 30 kpc region is $\sim 90-140$ km/s on $\sim 20-30$ kpc scale and $\sim 70-100$ km/s on smaller scales $\sim 7-10$ kpc. These measurements match our expectations of typical velocities in the ICM from numerical simulations and various observational constraints. Measured velocity spectra suggest power injection from the center (e.g. from the central AGN in Perseus). Spectra are consistent with the cascade turbulence. Their slopes are broadly consistent with the slope for the canonical Kolmogorov turbulence. It was previously shown that the heating of the gas due to dissipation of such motions balances the gas radiative cooling in the cluster core \citep{Zhu14b}.
\item The gas clumping estimated from the PS of the density fluctuations is lower than $7-8$ per cent at distance from $30$ to $220$ kpc from the center, which gives a density bias less than $3-4$ per cent in the cluster core. The clumping factor is dominated by fluctuations on large scales.
\item Systematic uncertainties in the analysis were analyzed. Conservative estimates of the final uncertainty on the amplitude of density fluctuations is $\sim 50$ per cent and on the velocity amplitude is $\sim 60$ per cent.
Future direct measurements of the velocities of gas motions with the X-ray microcalorimeters on-board {\it Astro-H}, {\it Athena} and {\it Smart-X} will allow us to better calibrate the statistical relation between the amplitude of density fluctuations and the velocity amplitude. Any strong deviations from the proportionality coefficient $\sim 1$ would indicate the importance of the neglected physics in the ICM.
\end{itemize}
\section{Acknowledgements}
Support for this work was provided by the NASA through Chandra award number AR4-15013X issued by the Chandra X-ray Observatory Center, which is operated by the Smithsonian Astrophysical Observatory for and on behalf of the NASA under contract NAS8-03060. S.W.A. acknowledges support from the US Department of Energy under contract number DE-AC02-76SF00515. I.Z. and N.W. are partially supported from Suzaku grants NNX12AE05G and NNX13AI49G. P.A. acknowledges financial support from FONDECYT grant 1140304. E.C. and R.S. are partly supported by grant No. 14-22-00271 from the Russian Scientific Foundation.
|
\section{Introduction}
Turbulence is ubiquitous in astrophysical plasmas, ranging
from coronal mass ejections and
stellar winds \cite{1995ARA&A..33..283G},
through star formation in molecular clouds \cite{2004RvMP...76..125M},
to the gas in the interstellar \cite{2004ARA&A..42..211E} and
intracluster medium.
While experimental setups \cite{2014PhPl...21a3505C} become increasingly
more realistic, they are
still far away from the regime acting in such extreme conditions.
For numerical simulations it is computationally too expensive
(if even possible) to capture
the entire range of physical processes from plasma kinetics to
the integral scales of turbulence.
In an astrophysical context, one has to further contend
with the additional complications brought about by high compressibility
and the accompanying supersonic and super-Alfvenic motion.
Possible ways to circumvent the infeasibility of direct numerical simulations
are the
use of calculations based on mean-field theories or
large-eddy simulations \cite{2014arXiv1404.2483S}.
These simulations only resolve the energy containing large scale dynamics and
require a subgrid-scale (SGS) model to account for unresolved effects.
While a lot of research has been successfully carried out in the realm of
hydrodynamics \cite{2009lesc.book.....G}, compressible magnetohydrodynamic SGS
closures are essentially unexplored.
Previous research is mainly based on the concept of
turbulent dissipation in incompressible flows
\cite{Mueller2002,1994PhPl....1.3016T,Chernyshov2014,Miki2008}.
They expand the idea of a turbulent eddy-viscosity to an additional
eddy-resistivity in the induction equation and propose different phenomenological
models.
Even though these models are then evaluated \textit{a posteriori}\xspace, a general verification and
justification \textit{a priori}\xspace has so far only been considered
for a single incompressible dataset \cite{2010arXiv1010.5759B}.
Thus, our objective is to establish the validity of closures for the filtered,
compressible MHD equations by coarse-graining multiple datasets from
high-resolution simulations of statistically homogeneous, forced MHD turbulence.
In general, the effect of finite resolution in numerical simulations can be
mimicked by applying a low-pass filter to the standard, ideal MHD equations.
This is achieved by convolving the equations with a suitable filter kernel $G$.
See e.g. Garnier \etal \cite{2009lesc.book.....G} for details on the properties of low-pass
filtering and the conditions that $G$ needs to satisfy.
For a homogeneous, isotropic, stationary kernel, under periodic boundary
conditions \cite{1983PhFl...26.2851F} the equations take the following form
\begin{eqnarray*}
\pd{\overline{\rho}}{t}+\div {\overline{\rho} \fav{\v{u}}} = 0,\\
\pd{\flt{\rho} \fav{\v{u}}}{t}
+ \div{\flt{\rho} \fav{\v{u}} \otimes \fav{\v{u}}
- \flt{\v{B}} \otimes \flt{\v{B}}}
+ \grad{\overline{P} + \frac{\flt{B}^2}{2}} =&- \nabla \cdot \tau,\\
\pd{\flt{\v{B}}}{t} - \curl\bra{\fav{\v{u}} \times \flt{\v{B}}} =
\curl{{\bm{\mathcal{E}}}}.
\end{eqnarray*}
The units of the magnetic field $\flt{\v{B}}$ incorporate $1/\sqrt{4\pi}$.
An overbar $\flt{\Box}$ denotes\footnote{Throughout the paper the symbol $\protect\Box$ is
used as a generic placeholder for variables.
$\protect\widehat{\protect\Box}$ designates closure expressions. Furthermore,
we employ Einstein summation convention and
$\protect\Box_{i,k}$ is identified with the $k$-th partial derivative of the
$i$-th component of $\protect\Box$.}
filtered and a tilde $\fav{\Box}$ mass-weighted filtered
quantities \cite{1983PhFl...26.2851F}.
For instance, the filtered density field is given by $\overline{\rho}=G\ast\rho$,
while the mass-weighed filtered velocity field is $\fav{\v{u}} = \flt{\v{u}\rho}/\overline{\rho}$.
In this formalism, all filtered primary quantities, density $\overline{\rho}$,
velocity $\fav{\v{u}}$, magnetic field $\flt{\v{B}}$, and thermal pressure $\overline{P}$
are presumed to be known and directly accessible.
Due to the introduction of mass-weighted filtering the only remaining terms
that require closure are the SGS stress $\tau$
and the electromotive force (EMF), ${\bm{\mathcal{E}}}$.
They are analytically expressed \cite{Miki2008} as
\begin{eqnarray}
{\bm{\mathcal{E}}} &= \flt{\v{u}\times\v{B}} - \fav{\v{u}} \times
\flt{\v{B}}\label{eq:EMFdata}
\quad \mathrm{and} \\
\label{eq:tau_def}\tau_{ij} &= \tuijdata - \tbijdata +
\bra{\flt{B^2} - \flt{B}^2}\frac{\delta_{ij}}{2}, \quad \mathrm{with} \nonumber \\
\tuijdata &\equiv \flt{\rho} \bra{\fav{u_i u_j} - \fav{u}_i \fav{u}_j}
\qquad \mathrm{and} \qquad
\tbijdata \equiv \bra{\flt{B_i B_j} - \flt{B}_i~\flt{B}_j}\;. \label{eq:tudata}
\end{eqnarray}
The SGS stress tensor can be decomposed into the well-known turbulent
Reynolds stress $\tau^{\mathrm{u}}$,
a turbulent Maxwell stress $\tau^{\mathrm{b}}$ and
a magnetic pressure term.
Furthermore, the definitions of the SGS energies are obtained from applying the
filter to the total filtered energy density $\flt{E}$, which can be decomposed into
the contribution due to resolved fields only and a remainder, designated as SGS energy
\begin{eqnarray*}
\flt{E} = \underbrace{\frac{1}{2} \flt{\rho} \fav{u}^2 + \frac{1}{2}\flt{B}^2}_
{\mathrm{(resolved)}}
+ \underbrace{\frac{1}{2}\flt{\rho} \bra{\fav{u^2} - \fav{u}^2} + \frac{1}{2} \bra{\flt{B^2} - \flt{B}^2}}_
{= E^{\mathrm{u}}_{\mathrm{sgs}} + E^{\mathrm{b}}_{\mathrm{sgs}} \equiv E_{\mathrm{sgs}} \; \mathrm{(unresolved)}}.
\end{eqnarray*}
It is important to point out that in general the filtering operator is not a Reynolds operator,
in particular $\flt{\flt{\Box}}\neq\flt{\Box}$.
It follows that SGS terms, like $E_{SGS}$, carry information not only about the interactions
between unresolved fields but also about cross-scale interactions between unresolved and resolved fields.
In addition to this,
the turbulent magnetic pressure is identical to the magnetic
SGS energy $E^{\mathrm{b}}_{\mathrm{sgs}}$
and both kinetic and magnetic SGS energies are
directly given by $2E^{\mathrm{u}}_{\mathrm{sgs}} = \Tr\bra{\tau^{\mathrm{u}}}$ and
$2E^{\mathrm{b}}_{\mathrm{sgs}} = \Tr\bra{\tau^{\mathrm{b}}}$,
i.e. they constitute the isotropic parts of the respective SGS tensors.
Following the general tensor decomposition, the deviatoric, traceless
parts are then
given by $\tau^{\Box *}_{ij} = \tau^{\Box}_{ij} - \frac{1}{3}\delta_{ij}\tau^{\Box}_{kk}$.
\section{Traditional closures}
\label{sec:TradClosures}
In hydrodynamics, the traceless part of the SGS stress tensor
is commonly closed
by means of the eddy-viscosity hypothesis
$\tu[*] = - 2 \nu^\mathrm{u} \flt{\rho} \fav{\mathcal{S}_{ij}^*}$
in analogy to the molecular viscosity term in the momentum equation, where
$\fav{\mathcal{S}_{ij}} \equiv \frac{1}{2}\bra{\fav{u}_{i,j} + \fav{u}_{j,i}}$
is the filtered kinetic rate-of-strain tensor.
This introduces a turbulent kinetic eddy-viscosity
$\nu^\mathrm{u}~=~C^\mathrm{u}_\nu \Delta \bra{E^{\mathrm{u}}_{\mathrm{sgs}}/\flt{\rho}}^{1/2}$ which is proportional
to a characteristic velocity, commonly given by the kinetic SGS energy,
and a characteristic length scale $\Delta$.
This closure has already been applied directly to MHD
\cite{Mueller2002,Petro2007} by neglecting the magnetic contribution
$\tb[*]$ in the momentum equation.
A turbulent magnetic viscosity
$\nu^\mathrm{b} = C^\mathrm{b}_\nu \Delta \bra{E^{\mathrm{b}}_{\mathrm{sgs}}}^{1/2}$ was used
in \cite{Miki2008} with the closure
\begin{eqnarray*}
\tb[*] = - 2 \nu^\mathrm{b} \flt{\mathcal{M}_{ij}} \;,
\end{eqnarray*}
where $\flt{\mathcal{M}_{ij}} \equiv \frac{1}{2}\bra{\flt{B}_{i,j} + \flt{B}_{j,i}}$
is the filtered magnetic rate-of-strain tensor.
The SGS energies can either be determined by individual evolution equations,
where several terms again require closure, or by an instantaneous closure.
Smagorinsky \cite{Smagorinsky1963} introduced such an instantaneous closure
in pure incompressible hydrodynamics ($\v{B} = 0$) by assuming the SGS energy
flux to be in equilibrium
with the rate of dissipation
\begin{eqnarray}
\label{eq:SmagU}
\widehat{E}^{\mathrm{u}}_{\mathrm{sgs}}= C^\mathrm{u}_\mathrm{E} \Delta^2 \rho |\fav{\mathcal{S}}^*|^2 \;.
\end{eqnarray}
Here, $|\fav{\mathcal{S}}^*| \equiv \sqrt{2 \fav{\mathcal{S}_{ij}^*} \fav{\mathcal{S}_{ij}^*}}$
denotes the rate-of-strain magnitude.
Finally, the EMF is commonly modeled
(e.g. \cite{Mueller2002,1994PhPl....1.3016T,Petro2007,Miki2008})
by variations of
\cite{Yoshizawa1990}
\begin{eqnarray*}
\widehat{\bm{\mathcal{E}}} = \alpha \flt{\v{B}} - \beta \flt{\v{J}} + \gamma \fav{\v{\Omega}},
\end{eqnarray*}
with resolved
current $\flt{\v{J}}=\curl \flt{\v{B}}$ and
vorticity $\fav{\v{\Omega}} = \curl \fav{\v{u}}$.
The coefficients $\alpha$, $\beta$, and $\gamma$ are typically related
to the $\alpha$-effect, turbulent resistivity and turbulent cross helicity,
respectively.
The commonly used closures for these coefficients,
\begin{eqnarray*}
\alpha = C^\mathcal{E}_{\alpha} t_{\mathrm{turb}} H, \qquad &
\beta = C^\mathcal{E}_{\beta} t_{\mathrm{turb}} E_{\mathrm{sgs}}/\flt{\rho}, \qquad &
\gamma = C^\mathcal{E}_{\gamma} t_{\mathrm{turb}} W, \qquad
\end{eqnarray*}
are based on dimensional arguments, with turbulent
cross helicity $W = \flt{\v{u}\cdot \v{B}} - \fav{\v{u}}\cdot \flt{\v{B}}$,
residual helicity
$H \sim \bra{\flt{\v{J} \cdot \v{B}}-\flt{\v{J}} \cdot \flt{\v{B}}} - \flt{\rho}\bra{\flt{\v{\Omega} \cdot \v{u}}-\fav{\v{\Omega}} \cdot
\fav{\v{u}}}$,
and timescale $t_{\mathrm{turb}}~=~\Delta~\bra{E_{\mathrm{sgs}}/\overline{\rho}}^{-1/2}$.
\section{Nonlinear closures}
\label{sec:NonLinClosures}
In our new approach we adopt the compressible hydrodynamic nonlinear closure for the kinematic deviatoric
stress tensor $\tudata[*]$
from \cite{Schmidt2011}, similar to the incompressible one from~\cite{Woodward2006}.
We propose the straightforward extension to MHD with
\begin{eqnarray}
\label{eq:tuStar}
\tu[*] &= 2 C^\mathrm{u}_\mathrm{nl} E^{\mathrm{u}}_{\mathrm{sgs}} \left (
\frac{\fav{u}_{i,k} \fav{u}_{j,k}}{\fav{u}_{l,s} \fav{u}_{l,s}}
- \frac{1}{3} \delta_{ij} \right), \\
\label{eq:tbStar}
\tb[*] &= 2 C^\mathrm{b}_\mathrm{nl} E^{\mathrm{b}}_{\mathrm{sgs}} \left (
\frac{\flt{B}_{i,k} \flt{B}_{j,k}}{\flt{B}_{l,s} \flt{B}_{l,s }}
- \frac{1}{3} \delta_{ij} \right) \;.
\end{eqnarray}
The tensorial structure, e.g. $\fav{u}_{i,k} \fav{u}_{j,k}$,
can be obtained by a Taylor expansion
discarding terms with $2^\mathrm{nd}$ and higher order gradients of the resolved fields.
The overall normalisation with the subgrid-scale energies comes from the constraint that the SGS stresses
vanish in laminar flows.
Applying the nonlinearity idea to the EMF
generalizes the closure proposed by \cite{2010arXiv1010.5759B} to
the compressible
regime
\begin{eqnarray}
\widehat{\mathcal{E}}_{i} = \varepsilon_{ijk} C^\mathcal{E}_\mathrm{nl} \Delta^2 \fav{u}_{j,s} \flt{B}_{k,s} \;.
\end{eqnarray}
The closure explicitly preserves the anti-symmetry
between velocity and magnetic field in ${\bm{\mathcal{E}}}$, which in turn helps in
capturing their relative geometry.
Finally, to complete the set of nonlinear closure equations,
we use the Smagorinsky expression
for the turbulent kinetic energy \eref{eq:SmagU}
and propose an analogous extension to the magnetic part
\begin{eqnarray}
\label{eq:SmagB}
\widehat{E}^{\mathrm{b}}_{\mathrm{sgs}} = C^\mathrm{b}_\mathrm{E} \Delta^2 |\flt{\mathcal{M}}|^2 \;.
\end{eqnarray}
Here, the turbulent magnetic energy
is proportional to the magnetic rate-of-strain magnitude
$|\flt{\mathcal{M}}| \equiv \sqrt{2 \flt{\mathcal{M}_{ij}}\, \flt{\mathcal{M}_{ij}}}$.
There are two advantages of closing $E^{\mathrm{u}}_{\mathrm{sgs}}$ and $E^{\mathrm{b}}_{\mathrm{sgs}}$
separately and not jointly via the total SGS energy.
First, there is no additional need to close the often neglected turbulent
magnetic pressure, as it is given by $E^{\mathrm{b}}_{\mathrm{sgs}}$.
Second, the individual energies provide closures to the isotropic parts of
the turbulent stress tensors $\tudata$ and $\tbdata$.
\section{Validation Method}
\label{sec:Validation}
In order to evaluate
the proposed closures, we perform an \textit{a priori}\xspace
comparison using simulation data obtained
from two, grid-based MHD codes (\textsc{Enzo}\xspace \cite{Enzo2013} and \textsc{FLASHv4}\xspace~\cite{Fryxell2000}).
This way the results are less likely to hinge on the particulars of the numerical implementation.
In both cases we follow the evolution of a compressible, isothermal fluid in
a cubic box with resolution of $512^3$ grid cells and periodic boundary
conditions, starting from uniform initial conditions.
In \textsc{Enzo}\xspace we use an ideal equation of state with adiabatic exponent
$\kappa = 1.001$ in order to approximate isothermal gas.
\textsc{Enzo}\xspace is a finite-volume code, i.e. the evolution equations are evaluated in integral form by solving
a Riemann problem for the mass, momentum and energy flux through cell walls. This allows for the conservation of MHD invariants (e.g. energy)
to machine precision.
We use a MUSCL-Hancock scheme \cite{Waagan2011} (a second-order accurate Godunov extension) with second-order Runge-Kutta time integration and
Harten-Lax-van Leer (HLL) Riemann solver (a two-wave, three-state solver) to solve the ideal MHD equations.
The \textsc{FLASHv4}\xspace code is similar to \textsc{Enzo}\xspace (second-order accurate in space and time), but uses the positive-definite
HLL3R Riemann solver \cite{Waagan2011}. Another difference is that \textsc{FLASHv4}\xspace uses a polytropic equation of state to keep the gas
exactly isothermal. Moreover, explicit kinematic viscosity and magnetic resistivity terms are included in the momentum,
energy, and induction equations. We set the kinematic and magnetic Reynolds
numbers to $\Re=\mathrm{Rm}=3780$.
Consequently, the magnetic Prandtl number $\mathrm{Pm}=\mathrm{Rm}/\Re$ is unity. For details on the numerical methods used in \textsc{FLASHv4}\xspace,
including viscous and resistive dissipation, see \cite{Federrath2011} and \cite{Federrath2014}.
Both codes employ divergence cleaning \cite{Dedner2002} to maintain
$\nabla \cdot \v{B} = 0$.
A state of homogeneous and isotropic turbulence is reached by supersonic
stochastic driving in the momentum equation (given by an Ornstein-Uhlenbeck process) at small
wave-numbers, similar to \cite{Schmidt2009,2010A&A...512A..81F}.
Thus, the forcing field is evolving in time and space.
The associated large auto-correlation time-scale $T$ of the forcing
translates to the eddy turnover time of the largest, energy-containing eddies.
It is therefore the chosen unit of time in the following.
We explore a range of parameters.
The initial strength of the magnetic field is set by the plasma $\beta_\mathrm{p}$
-- the ratio of thermal to magnetic pressure.
The final sonic Mach number $\mathrm{M_s}$ is determined by the forcing amplitude.
For the \textsc{Enzo}\xspace simulations we have initial $\beta_\mathrm{p} = 0.25, 2.5, 25$,
with $\mathrm{M_s} \approx 2.5$ after $t \approx 2T$ turnover times.
The \textsc{FLASHv4}\xspace simulations reach $\mathrm{M_s} \approx 4, 2$ for initial $\beta_\mathrm{p} = 1, 5$,
keeping instead constant Alfvenic Mach number $\mathrm{M_a} \approx 3$.
We discard all initial data affected by transients (before a simulation time
of $t=2T$) and analyze consequent snapshots taken approximately in intervals of
$0.15T$ and $0.1T$ for the \textsc{Enzo}\xspace and \textsc{FLASHv4}\xspace datasets, respectively.
The analysis begins with the application of a low-pass Gaussian (test) filter
to the equations of motion.
In the context of the closures we investigate, filtered quantities
(i.e. density, velocity, and magnetic field)
are interpreted as resolved, while the remainders represent the
unresolved small scales.
We can then compute $\tau$ and ${\bm{\mathcal{E}}}$ both
directly from \eref{eq:EMFdata} and \eref{eq:tudata},
and from their respective traditional and nonlinear closures.
The determination of the length-scale of the filter bears some consideration.
It needs to fall within an intermediate range of length scales, away from
the particular effects of both the large-scale forcing and the small-scale dissipation.
The largest scale of the system is the full box size $L$ and corresponds to the wavenumber
$n=1$, while the smallest scale is given by the Nyquist wavenumber $n_\mathrm{Nyq}=N/2=256$ for the linear
numerical resolution $N=512$ grid cells. The turbulence injection wavenumber is $n_\mathrm{inj} = 2$
(corresponding to half the box size $L/2$) in both codes, which is why the
energy spectra,
as illustrated in \fref{fig:spectrum}, peak there.
The figure shows the mean kinetic and total (kinetic plus magnetic) energy
spectra as a function of wavenumber.
Since the stochastic forcing is implemented only in the momentum equation both for \textsc{Enzo}\xspace and \textsc{FLASHv4}\xspace,
the kinetic energy spectrum exhibits the most direct imprint of the forcing itself. Conversely, the total energy
spectrum carries the overall effect of the small-scale dissipation through both kinetic and magnetic channels.
\Fref{fig:spectrum} demonstrates that our simulations produce approximate power-law scaling
within a narrow range of wavenumbers, which is indicative of self-similar
turbulent fluctuations \cite{Federrath2013a}.
This can be interpreted as inertial range dynamics, although the nature of the inertial range
in compressible MHD turbulence is still not fully understood
\cite{Galtier2011, Banerjee2013, Aluie2011, Aluie2013}.
Furthermore, this range separates the forcing scale and the dissipation scales and is not affected by
numerical diffusion in the absence of a bottleneck effect as demonstrated by \cite{Kitsionas2009}.
The vertical dotted line in \fref{fig:spectrum} indicates our chosen filter length scale,
corresponding to $\Delta = 16$ grid cells
or wavenumber $n_\mathrm{filter} =N/(2\Delta) = 16$.
This filter scale falls within the range of
the self-similar power-law range for both the kinetic and total energy spectra.
This is why we use this ideal scale for our filter in the following analysis.
\begin{figure}[!ht]
\centering
\hfill
\subfigure[kinetic energy\label{fig:kin spectrum}]{
\includegraphics[width=0.4\textwidth]{fig1a}
}
\hfill
\subfigure[kinetic $+$ magnetic energy\label{fig:tot spectrum}]{
\includegraphics[width=0.4\textwidth]{fig1b}
}
\hfill
\caption{
Kinetic (a) and total (b) energy spectrum for each dataset, averaged over the time between $2T$ and $5T$.
The kinetic energy is calculated from the Fourier transform of $\sqrt{\rho}\v{u}$.
}
\vspace{-1em}
\label{fig:spectrum}
\end{figure}
Additionally, this provides the motivation to treat data from both simulations on equal footing,
even though \textsc{FLASHv4}\xspace has explicit viscosity and diffusivity while \textsc{Enzo}\xspace solves the ideal MHD equation,
subject to numerical dissipation only.
In order to incorporate coordinate independence, a scalar field is chosen for
comparison, the SGS energy flux $\Sigma$, i.e. the term responsible for
the transfer of SGS energy between resolved and unresolved scales.
Its components associated with the Reynolds and Maxwell stresses and the EMF are
$\Sigma^u=\tau^u_{ij} \fav{S}_{ij}$, $\Sigma^b=\tau^b_{ij} \fav{S}_{ij}$ and
$\Sigma^{{\mathcal{E}}}={\bm{\mathcal{E}}} \cdot \flt{\v{J}}$, respectively
(see appendix of \cite{Miki2008} for the detailed SGS energy equation).
Here, we substitute \eref{eq:EMFdata} and \eref{eq:tudata} to obtain
exact fluxes $\Sigma^\Box$ and match these to model fluxes
$\widehat{\Sigma}^\Box$ that employ the corresponding closures.
For example, in the case of the eddy-viscosity closure for the deviatoric
turbulent Reynolds stress tensor we compare the exact flux
\begin{eqnarray}
\Sigma^{\mathrm{u*}} =
\tau_{ij}^{\mathrm{u*}} \fav{\mathcal{S}_{ij}} =
\flt{\rho} \bra{ \bra{\fav{u_i u_j} - \fav{u}_i \fav{u}_j}
- \frac{1}{3} \delta_{ij}
\bra{\fav{u_k u_k} - \fav{u}_k \fav{u}_k}}
\fav{\mathcal{S}_{ij}}
\end{eqnarray}
with the model flux
\begin{eqnarray}
\widehat{\Sigma}^{\mathrm{u*}} =
\tu[*] \fav{\mathcal{S}_{ij}} =
- \nu^\mathrm{u} \flt{\rho} |\fav{\mathcal{S}^*}|^2
= - C^\mathrm{u}_\nu \Delta \bra{\flt{\rho} E^{\mathrm{u}}_{\mathrm{sgs}} }^{1/2}
|\fav{\mathcal{S}^*}|^2 \;.
\label{eq:eddyViscFlux}
\end{eqnarray}
On the one hand, the comparison involves the determination of the
constant (in space and time), dimensionless closure coefficients
$C_{\Box}^{\Box}$.
They are computed individually for each snapshot
by minimizing the error between $\Sigma^\Box$ and $\widehat{\Sigma}^\Box$
in the least-square sense.
This allows to further test the constancy of the coefficients with respect to
time and plasma parameters.
On the other hand, the general performance of the closure is gauged by computing the
Pearson correlation coefficient of $\Sigma^\Box$ and $\widehat{\Sigma}^\Box$, where
the obtained closure coefficients are substituted in.
Several assumptions should be pointed out concerning this validation technique.
Firstly, the simulation data we have available for comparison fall short of realistic
astrophysical parameters, e.g. with regards to Reynolds numbers and resolution.
In that sense, it would be interesting to use higher resolution direct numerical simulation data or
three-dimensional observations or experimental results. The problem is that
experimental data for supersonic compressible turbulent plasmas are not available
and obtaining realistic Reynolds numbers is computationally challenging
for astrophysical parameters.
However, as seen from \figref{fig:spectrum}, the data we have are sufficiently well resolved for
our analysis.
Secondly, in choosing the SGS energy flux $\Sigma$, as a diagnostic variable,
we implicitly assume that in the context of homogeneous and isotropic turbulence
the turbulent transport (encoded by terms of the form
$ \div {\fav{u}\cdot \tau}$ and $\div{\flt{B}\times {\bm{\mathcal{E}}}}$)
averages out to zero on subgrid scales.
This assumption can nevertheless be easily relaxed by incorporating further diagnostic variables.
Finally, we have focused on the SGS energy since it increases monotonically with the strength of turbulence
regardless of the type of turbulence (e.g. compressive or solenoidal, weak or strong, etc.).
As an extension, the other two quadratic MHD invariants -- the magnetic helicity and cross-helicity,
may further highlight distinct turbulence properties present in particular flow configurations.
These should be kept in mind as further avenues of investigation, once a preferred closure
has been identified by the described validation technique.
\section{Results}
\begin{table}
\caption{\label{tab:valueoverview}
Model coefficient overview -- coefficient value and energy flux correlation:
median and bounds of the central 90\% interval across all datasets.}
\begin{indented}
\item[]
\lineup
\begin{tabular}{@{}lcll}
\br
\multicolumn{1}{c}{Model} & Coefficient & Value & Corr[$\Sigma^\Box,\,\widehat{\Sigma}^\Box$]\\
\mr
\multirow{2}{*}{Smagorinsky}
& $C^\mathrm{u}_\mathrm{E}$ & $0.056^{+0.016}_{-0.015}$ & $0.90^{+0.024}_{-0.04}$ \\
& $C^\mathrm{b}_\mathrm{E}$ & $0.075^{+0.034}_{-0.007}$ & $0.91^{+0.021}_{-0.04}$ \\
\mr
\multirow{2}{*}{eddy-viscosity}
& $C^\mathrm{u}_\nu$ & $0.061^{+0.045}_{-0.019}$ & $0.70^{+0.13}_{-0.11}$ \\
& $C^\mathrm{b}_\nu$ & $-0.002^{+0.029}_{-0.03}$ & $0.06^{+0.14}_{-0.06}$ \\
\mr
\multirow{3}{*}{nonlinear}
& $C^\mathrm{u}_\mathrm{nl}$ & $0.68^{+0.09}_{-0.09}$ & $0.94^{+0.04}_{-0.04}$ \\
& $C^\mathrm{b}_\mathrm{nl}$ & $0.77^{+0.08}_{-0.12}$ & $0.90^{+0.04}_{-0.07}$ \\
& $C^\mathcal{E}_\mathrm{nl}$ & $0.12^{+0.013}_{-0.024}$ & $0.79^{+0.07}_{-0.17}$\\
\mr
\multirow{3}{*}{$\alpha-\beta-\gamma$}
& $C^\mathcal{E}_{\alpha}$ & $0.0007^{+0.0010}_{-0.0016}$ & \multirow{3}{*}{\Bigg\} $0.58^{+0.06}_{-0.16}$ }\\
& $C^\mathcal{E}_{\beta}$ & $0.020^{+0.009}_{-0.005}$ & \\
& $C^\mathcal{E}_{\gamma}$ & $-0.005^{+0.067}_{-0.045}$ & \\
\br
\end{tabular}
\end{indented}
\end{table}
The fitting results for the SGS stress tensors' energy flux
are given in \figref{fig:Esgs} for the isotropic components and \figref{fig:tauStar}
for the deviatoric components.
The isotropic parts of $\tudata$ and $\tbdata$ are given
by the SGS energies $\tau^\Box_{ii} = \frac{2}{3} E^\Box_{\mathrm{sgs}}$ from
\eref{eq:SmagU} and \eref{eq:SmagB}.
\begin{figure}[!ht]
\centering
\hfill
\subfigure[kinetic SGS energy\label{fig:EsgsU}]{
\includegraphics[width=0.4\textwidth]{fig2a}
}
\hfill
\subfigure[magnetic SGS energy\label{fig:EsgsB}]{
\includegraphics[width=0.4\textwidth]{fig2b}
}
\hfill
\caption{
Model coefficient values (top panels) and correlations (bottom panels with
--- median) from fitting model energy flux
$\widehat{\Sigma}^\Box_{\mathrm{E}} = \widehat{\tau}^\Box_{ij}\fav{\mathcal{S}_{ij}}$
to exact flux for the isotropic parts of the SGS stress tensors.
These are given by the respective energy model in the trace elements
($\tau^\Box_{ii} = \frac{2}{3} E^\Box_{\mathrm{sgs}}$).
Each panel contains the joint data of all
simulations and each snapshot is represented by a marker.
Values are given in \tref{tab:valueoverview}.}
\label{fig:Esgs}
\end{figure}
Both, the coefficient values of kinetic part (\figref{fig:EsgsU} top panel)
and the magnetic part (\figref{fig:EsgsB} top panel), have a small spread within
a factor of two across time and all simulations.
Furthermore, closure and data are highly correlated (bottom panels) with a
median correlation coefficient of 0.90 and 0.91, respectively.
More detailed numerical values of these and all following coefficients and
correlations are listed in \tref{tab:valueoverview}.
\begin{figure}[!ht]
\centering
\subfigure[traceless kinetic SGS stress tensor\label{fig:tauUstar}]{
\includegraphics[width=0.45\textwidth]{fig3a}
}
\hfill
\subfigure[traceless magnetic SGS stress tensor\label{fig:tauBstar}]{
\includegraphics[width=0.49\textwidth]{fig3b}
}
\caption{
Model coefficient values (top panels) and correlations (bottom panels with --- median)
from fitting model energy flux
$\widehat{\Sigma}^\Box = \widehat{\tau}^\Box_{ij}\fav{\mathcal{S}_{ij}}$
to exact flux for the nonlinear closure (left panels) and eddy-viscosity closure (right panels).
Each panel contains the joint data of all
simulations and each snapshot is represented by a marker.
Values are given in \tref{tab:valueoverview}.}
\label{fig:tauStar}
\end{figure}
The differences in the deviatoric parts
$\tunoij[*]$ in \figref{fig:tauUstar} and $\tbnoij[*]$ in \figref{fig:tauBstar}
between the nonlinear and the eddy-viscosity closures
are apparent.
While our nonlinear closure exhibits approximately constant coefficient values
and correlations over time in all simulations, the
kinetic eddy-viscosity closure shows a correlation weaker by $\approx 0.2$ and
bigger spread in coefficient values.
Moreover, the magnetic eddy-viscosity closure is effectively uncorrelated with
the simulation data and
the coefficients can even switch sign at different times.
\begin{figure}[!h]
\centering
\includegraphics{fig4}
\vspace{-1em}
\caption{
Representative snapshot (\textsc{Enzo}\xspace sim. with
$\beta_\mathrm{p} = 2.5$ at $t = 4.44T$) of the energy flux
$\Sigma^\Box = \tau^\Box_{ij}\mathcal{S}_{ij}$
distribution within the simulation box.
}
\vspace{-1em}
\label{fig:fluxDistribution}
\end{figure}
The performance of the different closures can be understood from
\figref{fig:fluxDistribution}, where we plot the energy flux distributions
$\Sigma^\mathrm{u*}$ and $\Sigma^\mathrm{b*}$ for a single snapshot.
A negative flux $\Sigma^{\mathrm{u}*}<0$ corresponds to a forward energy cascade
--
the transfer of energy from resolved to subgrid scales, because
$\Sigma^{\mathrm{u}*}$ appears
as a sink term in the SGS kinetic and magnetic energy evolution equations and
as a source term in the respective resolved energy equations.
Conversely, a positive flux corresponds to an inverse energy cascade,
i.e. transport of energy from subgrid to resolved spatial scales.
The general distribution of the actual fluxes in \figref{fig:fluxDistribution}
is representative for all snapshots.
The kinetic SGS energy fluxes are globally almost 1:1 in both directions of the
turbulent cascade with a slight tendency towards the forward cascade.
However, the forward cascade is about 3-10 times stronger depending on the parameters
as indicated by the position of the peaks in the distribution.
For this reason, the kinetic eddy-viscosity closure shows a moderate
correlation even though it captures only the forward energy cascade --
from large to small scales.
In fact, since under the eddy-viscosity hypothesis
the kinetic SGS energy flux has the form
$\widehat{\Sigma}^{\mathrm{u*}}=-\nu^\mathrm{u} \flt{\rho}|\fav{S}^*|^2$,
see \eref{eq:eddyViscFlux},
any model in which the eddy-viscosity $\nu^\mathrm{u}$ has a definite signature
with respect to space cannot reproduce a bi-directional energy cascade
that is well represented by the nonlinear closure.
In contrast to the kinetic SGS energy flux, the global magnetic flux clearly has
a preferred direction.
Depending on the parameters, between 60\% and 80\% of cells have a positive SGS magnetic
flux
indicating energy transfer from large to small scales.
Nevertheless, the difference in strength is less pronounced as the overall
forward flux is
about 2 times stronger than the backward one.
Again, these properties are well captured by the proposed nonlinear closure whereas
the magnetic eddy-viscosity closure is poorly correlated in both strength and magnitude.
Finally, it should be noted, that the nonlinear closures also work
very well with the Smagorinsky energy closure.
Exchanging the exact expressions $E_{\mathrm{sgs}}^\Box$ in \eref{eq:tuStar} and
\eref{eq:tbStar} with $\widehat{E}_{\mathrm{sgs}}^\Box$ only slightly reduces the correlations
(max. 5\%) and the coefficients remain constant up to the second
significant figure (not plotted here).
Moving on to the EMF,
the nonlinear closure outperforms the
traditional $\alpha-\beta-\gamma$ closure
in almost all datasets, maintaining a constant coefficient
with a median correlation of 0.79 (\figref{fig:EMF_corr}).
\begin{figure}[ht!]
\centering
\includegraphics{fig5}
\caption{
Model coefficient values (top panels) normalized to the sample
median (---) and the corresponding Pearson
correlation coefficients (bottom panels) with 90\% central interval (- -)
for the nonlinear closure
(left panels) and the reference closure (right panels) from
energy flux fitting for $\widehat{\bm{\mathcal{E}}}$. Each panel contains the joint data of all
simulations and each snapshot is represented by a marker.
Values are listed in \tref{tab:valueoverview}.
}
\label{fig:EMF_corr}
\end{figure}
The traditional closure exhibits consistently weaker correlations,
despite the increased flexibility of three free coefficients.
Only $C^\mathcal{E}_{\beta}$, related to the turbulent resistivity term in the EMF,
is approximately constant, whereas $C^\mathcal{E}_{\alpha}$ and $C^\mathcal{E}_{\gamma}$
fail to maintain steady values or consistent signature.
The reason for the consistently better correlations of the nonlinear closure
is hinted at in \figref{fig:EMF_ali}.
\begin{figure}[ht!]
\centering
\includegraphics{fig6}
\caption{
Probability density plot of the local (cell-by-cell) alignment between ${\bm{\mathcal{E}}}$ and $\widehat{\bm{\mathcal{E}}}$. The lines designate the median PDF across all times
and datasets for the nonlinear closure (red ---) and
the $\alpha-\beta-\gamma$ closure (gray - -).
The shaded regions designate the central 90\% interval across all
times and datasets.
}
\label{fig:EMF_ali}
\end{figure}
This probability density plot of the local alignment between ${\bm{\mathcal{E}}}$ and $\widehat{\bm{\mathcal{E}}}$
demonstrates that the traditional closure is almost randomly aligned
(flat distribution) whereas the nonlinear closure approaches the desired
$\delta$-distribution at $0^\circ$.
\section{Conclusions and outlook}
In summary, we have proposed a set of constant coefficient closures
for the SGS stress and EMF in the filtered MHD equations and conducted
\textit{a priori}\xspace tests.
The tests we performed do show that the new nonlinear closures perform significantly better
than traditional, phenomenological
closures with respect to both structural and functional diagnostics. The tests consist of
filtering \textsc{Enzo}\xspace and \textsc{FLASHv4}\xspace simulations of homogeneous, isotropic turbulence
and comparing the resulting SGS terms to their respective closures
(dependent only on the filtered fields).
All quantities are compared via their contributions to the SGS energy
flux $\Sigma^{\Box}$,
where the closure coefficients are computed by individual least-square
fitting.
In addition, the alignment for the EMF vector is investigated.
All new coefficients correlate well with the data.
They are constant over time and as a direct consequence the proposed closures
may be implemented in large-eddy simulations without the need for a computationally expensive
dynamical procedure which computes the coefficient values at run time.
In addition, the coefficients remain constant across simulation runs from two different codes and
a wide range of plasma parameters, suggesting that the proposed closures capture
an underlying physical mechanism at work in highly compressible turbulent plasma flows.
Moreover, the new closures successfully represent the turbulent magnetic pressure,
reproduce the bi-directional energy cascade and are well aligned with the EMF.
We recognize the slightly lower correlation of the nonlinear closure in the
EMF than in the SGS stress counterpart, suggesting small room for improvement.
Nevertheless, the performance improvement over the traditional closures
already supports the implementation and validation of the new closures in an SGS model
for large-eddy simulations of compressible turbulent plasma flows.
These simulations would then allow us to infer the effect of the proposed model on the large scale flow in practice.
Potential applications include accretion disks \cite{Clark2011b}, star-forming magnetized clouds
\cite{Greif2008, Greif2012a}
and plasmas on cosmological scales \cite{Agertz2013, Devriendt2010,Vazza2014,Schleicher2013a, Latif2013, Latif2014}.
\ack
PG worked on section \ref{sec:NonLinClosures} and conducted the \textsc{Enzo}\xspace simulations.
DV worked on section \ref{sec:TradClosures}.
The analysis framework is a joint development by PG and DV.
PG acknowledges financial support by the
\textit{International Max Planck Research School for Solar System
Science at the University of Göttingen}.
DV, WS and DS acknowledge research funding by the
\textit{Deutsche Forschungsgemeinschaft (DFG)} under grant
\textit{SFB 963/1, project A15}.
CF thanks for funding provided by the ARC (grants DP130102078 and DP150104329).
We gratefully acknowledge the J\"ulich Supercomputing Centre (grant hhd20), the Leibniz
Rechenzentrum and the Gauss Centre for Supercomputing (grant pr32lo), the Partnership for
Advanced Computing in Europe (PRACE grant pr98mu), and the Australian National Computing
Infrastructure (grant ek9). The software used in this work was in part developed by the
DOE-supported Flash Center for Computational Science at the University of Chicago.
\section*{References}
\bibliographystyle{iopart-num}
|
\section{I. Normal Phase }
A normal mode with time-dependent frequency modulation [cf.~Eq.~\eqref{ham-nm} in the main text] can be generically described by the Hamiltonian
for a parametric oscillator:
\begin{equation}\label{ham1}
H= \frac{p^2}{2m} + \frac12 [\omega_0^2 +\mu(t)] x^2\,.
\end{equation}
For a classical oscillator with sinusoidal modulation, the above Hamiltonian describes the well known Mathieu parametric oscillator \cite{{mclachlan1964,nayfehnonlinear}}. The quantum parametric Hamiltonian can be rewritten as
\begin{align} \label{aham-1}
H& = [\omega_0 + \frac{\mu(t)}{2 \omega_0}] a^\dagger a + \frac{\mu(t)}{4 \omega_0} (a^2 + a^{\dagger 2})\nonumber\\
&\equiv \sqrt{\omega^2_0 + \mu(t)} {\bar a}^\dagger {\bar a}\,,
\end{align}
where the ladder operators $a, a^\dagger$ are defined with respect to the time independent model (i.e., $\mu(t)=0$).
\begin{figure}[htbp]
\centering
\includegraphics[width=\columnwidth]{normalcurv.pdf}
\caption{\label{gl_curves}Characteristic numerical time-integration plots of Eq.~\eqref{gl_eq} in different regions of the phase diagram. All curves are with $\lambda_0=0.4\Omega$, $\epsilon=0.5\Omega$, and $\kappa=\eta=0.1\Omega$. We see that for (i) $\omega_0=0.5\Omega$, mode $A_1$ is stable whereas $A_2$ is not, (ii) $\omega_0=0.7\Omega$, mode $A_1$ is unstable whereas $A_2$ is stable, and (iii) $\omega_0=0.9\Omega$, both modes $A_m$ are stable.}
\end{figure}
We consider a coupling to an external bath which is in the rotating wave approximation. Hence, the solutions to the Heisenberg equations of motion, in the presence of dissipation, for the operators $a$ and $a^\dagger$ take the form~\cite{tan2011}
\begin{eqnarray}
a(t) &=& G(t) a(0) + L^*(t) a^\dagger (0) + F(t)\,, \\
a^\dagger (t)&=& G^*(t) a^\dagger(0) + L(t) a (0)+ F^\dagger(t)\,,
\label{soln}
\end{eqnarray}
where $G$ and $L$ are time dependent functions obeying the initial conditions $G(0)= G^*(0)=1$ and $L(0)=L^*(0)=0$.
$F$ is an operator term which stems from the dissipation and also depends on the functions $G$ and $L$. It satisfies
the condition $F(0)=F^\dagger(0)=0$. The functions $G,L$ obey the integro-differential equations
\begin{align} \label{int-diff1}
\resizebox{.85\hsize}{!}{$\displaystyle{\dot G}(t) = -i (\omega_0 + \frac{\mu(t)}{2 \omega_0}) G - i \frac{\mu(t)}{2 \omega_0} L -\int_0^t ds K(t -s) G(s)\,,$}\\
\resizebox{.85\hsize}{!}{$\displaystyle{\dot L}(t)= i (\omega_0 + \frac{\mu(t)}{2 \omega_0}) L + i \frac{\mu(t)}{2 \omega_0} G -\int_0^t ds K(t -s) L(s)\,,$}
\label{int-diff2}
\end{align}
where the dissipative kernel $K(t) = \int d\omega J(\omega) e^{-i \omega t}$ and $J(\omega)$ is the spectral density of
the dissipative bath.
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{normal.pdf}
\caption{\label{gl_stability}The numerical stability diagram of the normal modes $A_m$ [cf.~Eq.~\eqref{ham-nm} in the main text] for $\lambda_0=0.4\Omega$ and $\kappa=\eta=0.1\Omega$. As the drive in the Dicke model affects each mode differently, we obtain two ``Arnold tongue'' stability diagrams that are shifted and scaled with respect to each other. Superposing the zones where both modes are stable leads to the stability of the normal phase (NP) [cf.~Fig.~\ref{fig:diagram} in the main text].}
\end{figure}
These coupled first order integro-differential equations are rather difficult to solve and numerical
solutions are needed. However, for standard Markovian dissipation, induced by cavity leakage in the rotating frame or coupling to an ohmic bath, $K(t) = \gamma \delta(t)$ where $\gamma$ is the damping rate. Substituting this in Eqs.~\eqref{int-diff1} and \eqref{int-diff2}, we see that they become a set of linear ODEs.
Observables and correlation functions can easily be obtained from these solutions. For example,
\begin{equation}
\langle a(t) \rangle = G(t) \langle a(0) \rangle + L^*(t) \langle a^\dagger(0) \rangle + \langle F(t) \rangle\,,
\end{equation}
where the expectation values are with respect to the initial density matrix.
Assuming initial conditions such that $\langle F(t)\rangle =0$, which are expected for baths with no particular ordering, we obtain
\begin{align}
\label{gl_eq}
\langle x(t) \rangle &={\rm Tr}\left\{\rho(t)\left(a+a^\dagger\right)\right\} \\
&= {\rm Re}[G(t)+L(t)] \langle x(0) \rangle - {\rm Im}[G(t)+L(t)] \frac{\langle p(0) \rangle }{m \omega_0}\,.\nonumber
\end{align}
For arbitrary initial conditions, the stability of the oscillator is dictated by whether the pre-factors $\left[G(t)+L(t)\right]$ grow with time as one approaches the asymptotic state. For the parametric oscillator, we expect it to become exponentially unstable as the strength of the driving is increased \cite{zerbe1994}. Choosing $\mu(t) = g \cos(2\Omega t)$, the zones of stability can be traced in the $\omega_0-g$ plane. The resulting stability diagram is the same as that for the classical Mathieu oscillators, which can also be extracted from the classical equations of motions [cf.~Eq.~\eqref{cla-eom} in the main text].
To obtain the full NP stability diagram of the Dicke model [see Fig~\ref{fig:diagram} in the main text], we study the stability of both normal modes $A_m$, $m=1,2$, with frequencies [cf.~Eq.~\eqref{ham-nm} in the main text]
\begin{eqnarray}
\Omega_1^2(t) &=& \omega_0^2 + 2 \lambda_0 \omega_0 + 2 \omega_0 \epsilon \cos(2\Omega t)\,, \\
\Omega_2^2(t) &=& \omega_0^2 - 2 \lambda_0 \omega_0 - 2 \omega_0 \epsilon \cos(2\Omega t)\,.
\end{eqnarray}
To simplify our calculation, we also assume that the baths the two modes couple to, have identical spectral densities~\cite{galve2010}. Though
relaxing this condition would lead to more technical complexity, it should not have any nontrivial physical consequence in the limit of
weak dissipation studied here.
We solve the corresponding Eqs.~(\ref{int-diff1})-(\ref{int-diff2}) and in Fig.~\ref{gl_curves}, we present characteristic numerical time-integration plots of Eq.~\eqref{gl_eq} for the normal modes $A_m$. Repeating this procedure for different $\omega_0$ and $\epsilon$, and by checking for converging/diverging solutions [cf.~Fig.~\ref{gl_curves}] we find the stability diagram for both modes $A_m$, see Fig.~\ref{gl_stability}.
The superposition of the two stability diagrams then yields the stability of the normal phase shown in Fig.~\ref{fig:diagram}.
\section{II. Mean-field analysis}
\setcounter{enumi}{2}
\setcounter{figure}{0}
\renewcommand{\thefigure}{\Roman{enumi}.\arabic{figure}}
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{fft.pdf}
\caption{\label{mf_curves}Characteristic numerical analysis of the mean-field equations [cf.~Eqs.~\eqref{amean}-\eqref{ymean} in the main text] leading to the displayed phase diagram [cf.~Fig.~\ref{fig:diagram} in the main text]. All plots are with $\lambda_0=0.4\Omega$, $\epsilon=0.5\Omega$, and $\kappa=\eta=0.1\Omega$. (a) and (c) are characteristic numerical time-integration plots of the mean-field equations with $\omega_0=0.6\Omega$ and $\omega_0=0.65\Omega$, respectively. (b) and (d) are the corresponding Fast Fourier Transform (FFT). We see that in the super-radiant phase (SP) region [Figs.~(a) and (b)], all order parameters are oscillating around a non-zero mean with relatively small oscillations amplitudes. In the dynamical normal phase (D-NP) region [Figs.~(c) and (d)], apart from $z$, all order parameters have a zero mean value, but their oscillations are large taking the full length of the central spin in its $x$ and $y$ components.}
\end{figure}
In the main text, we obtained a set of coupled non-autonomous mean-field equations [cf.~Eqs.~\eqref{amean}-\eqref{ymean} in the main text] for the mean field parameters $\alpha, x ~{\rm and} ~y$. These equations were solved numerically with a variety of ODE solvers. Characteristic numerical time-integration plots of the solutions to these equations are shown in Figs.~\ref{mf_curves} (a) and (b). Note that the solutions converge to the asymptotic regime for times $t\Omega \sim 500$ and are oscillatory. The time scale for reaching the asymptotic regime varies with the parameters. Repeating this procedure as a function of $\omega_0$ and $\epsilon$, the time-average of the order parameters over the steady-state behaviour (the last one-third of the integrated time) is presented in Figs.~\ref{fig:diagram}(a) in the main text.
We find that typically the solutions show sinusoidal oscillations characterized by the frequency of the parametric drive. However,
in certain parameter regimes, the order parameters oscillate strongly with a complex beat structure involving multiple frequencies. To analyze all these solutions in a systematic manner, we Fast Fourier Transform (FFT) the steady-state signals, see Figs.~\ref{mf_curves}(c) and (d). The largest amplitude of a finite frequency in such plots serves as a measure for the extent of the oscillation that the order paramaters undergo. This amplitude is plotted as a function of $\omega_0$ and $\epsilon$ in Figs.~\ref{fig:diagram}(b) in the main text.
\end{document} |
\section{Introduction}
The support recovery problem consists of determining a sparse subset of a set of variables that is relevant in producing a set of observations, and arises frequently in disciplines such as group testing \cite{Mal13, Ati12}, compressive sensing (CS) \cite{Fou13}, and subset selection in regression \cite{Mil02}. The observation models can vary significantly among these disciplines, and it is of considerable interest to consider these in a unified fashion. This can be done via probabilistic models relating the sparse vector $\beta \in \RR^p$ to a single observation $Y \in \RR$ in the following manner:
\begin{equation}
(Y|S=s,X=x,\beta=b) \sim P_{Y|X_S\beta_S}(\,\cdot\,|x_s,b_s), \label{eq:intr_model}
\end{equation}
where $S \subseteq \{1,\dotsc,p\}$ represents the set of relevant variables, $X \in \RR^{p}$ is a measurement vector, $X_S$ (respectively, $\beta_S$) is the subvector of $X$ (respectively, $\beta_S$) containing the entries indexed by $S$, and $P_{Y|X_S\beta_S}$ is a given probability distribution. Given a collection of measurements $\Yv\in\RR^{n}$ and the corresponding measurement matrix $\Xv \in \RR^{n \times p}$ (with each row containing a single measurement vector), the goal is to find the conditions under which the support $S$ can be recovered either \emph{perfectly} or \emph{partially}. In this paper, we study the information-theoretic limits for this problem, characterizing the number of measurements $n$ required in terms of the sparsity level $k$ and ambient dimension $p$ regardless of the computational complexity. Such studies are useful for assessing the performance of practical techniques and determining to what extent improvements are possible.
Before proceeding, we state some important examples of models that are captured by \eqref{eq:intr_model}.
\subsubsection*{Linear Model}
The linear model \cite{Wai09,Wai09a} is ubiquitous in signal processing, statistics, and machine learning, and in itself covers an extensive range of applications. Each observation takes the form
\begin{equation}
Y = \langle X, \beta \rangle + Z, \label{eq:intr_linear}
\end{equation}
where $\langle\cdot,\cdot\rangle$ denotes the inner product, and $Z$ is additive noise. An important quantity in this setting is the signal-to-noise ratio (SNR) $\frac{\EE[\langle X, \beta \rangle^2]}{\EE[Z^2]}$, and in the context of support recovery, the smallest non-zero absolute value $\beta_{\mathrm{min}}$ in $\beta$ has also been shown to play a key role \cite{Wai09,Wan10,Rad11}.
\subsubsection*{Quantized Linear Models}
Quantized variants of the linear model are of significant interest in applications with hardware limitations. An example that we will consider in this paper is the 1-bit model \cite{Bou08}, given by
\begin{equation}
Y = \sign\big(\langle X, \beta \rangle + Z\big),
\end{equation}
where the $\sign$ function equals $1$ if its argument is non-negative, and $-1$ if it is negative.
\subsubsection*{Group Testing}
Studies of group testing problems began several decades ago \cite{Dor43,Mal78}, and have recently regained significant attention \cite{Ati12,Cha11}, with applications including medical testing, database systems, computational biology, and fault detection. The goal is to determine a small number of ``defective'' items within a larger subset of items. The items involved in a single test are indicated by $X \in \{0,1\}^p$, and each observation takes the form
\begin{equation}
Y = \openone\bigg\{ \bigcup_{i \in S} \{X_i = 1\} \bigg\} \oplus Z,
\end{equation}
with $S$ representing the defective items, $Y$ indicating whether the test contains at least one defective item, and $Z$ representing possible noise (here $\oplus$ denotes modulo-2 addition). In this setting, one can think of $\beta$ as deterministically having entries equaling one on $S$, and zero on $S^c$.
The above examples highlight that \eqref{eq:intr_model} captures both discrete and continuous models. Beyond these examples, several other non-linear models are captured by \eqref{eq:intr_model}, including the logistic, Poisson, and gamma models.
\subsection{Previous Work and Contributions} \label{sec:PREVIOUS_WORK}
Numerous previous works on the information-theoretic limits of support recovery have focused on the linear model \cite{Wai09,Wan10,Rad11,Fle09,Aks13,Ree12,Ree13,Jin11,Aer10,Tan10,Akc10,Tul13,Sca13e}. The main aim of these works, and of that the present paper, is to develop necessary and sufficient conditions for which an ``error probability'' vanishes as $p \to \infty$. However, there are several distinctions that can be made, including:
\begin{itemize}
\item Random measurement matrices \cite{Wai09,Wan10,Fle09,Rad11} vs.~arbitrary measurement matrices \cite{Aer10,Ree13,Tan10};
\item Exact support recovery \cite{Wai09,Wan10,Fle09,Rad11} vs.~partial support recovery \cite{Ree12,Ree13,Akc10};
\item Minimax characterizations for $\beta$ in a given class \cite{Wai09,Wan10,Fle09,Rad11} vs.~average performance bounds for random $\beta$ \cite{Aks13,Tul13,Ree13}.
\end{itemize}
Perhaps the most widely-studied combination of these is that of minimax characterizations for exact support recovery with random measurement matrices. In this setting, within the class of vectors $\beta$ whose non-zero entries have an absolute value exceeding some threshold $\Bmin$, necessary and sufficient conditions on $n$ are available with matching scaling laws \cite{Wan10,Rad11}. See also \cite{Wu10,Wu12} for information-theoretic studies of the linear model with a mean square error criterion.
Compared to the linear model, research on the information-theoretic limits of support recovery for non-linear models is relatively scarce. The system model that we have adopted follows those of a line of works seeking mutual information characterizations of sparsity problems \cite{Mal78,Ati12,Aks13,Tan14}, though we make use of significantly different analysis techniques. Similarly to these works, we focus on random measurement matrices and random non-zero entries of $\beta$. Other works considering non-linear models have used vastly different approaches such as regularized $M$-estimators \cite{Lee15,Li14} and approximate message passing \cite{Tan14d}.
{\bf High-level Contributions:} We consider an approach using thresholding techniques akin to those used in information-spectrum methods \cite{Han03}, thus providing a new alternative to previous approaches based on maximum-likelihood decoding and Fano's inequality. Our key contributions and the advantages of our framework are as follows:
\begin{enumerate}
\item Considering both exact and partial support recovery, we provide non-asymptotic performance bounds applying to general probabilistic models, along with a procedure for applying them to specific models (\emph{cf.}~Section \ref{sec:APPLYING}).
\item We explicitly provide the constant factors in our bounds, allowing for more precise characterizations of the performance compared to works focusing on scaling laws (e.g., see \cite{Wai09,Akc10,Rad11}). In several cases, the resulting necessary and sufficient conditions on the number of measurements coincide up to a multiplicative $1+o(1)$ term, thus providing \emph{exact} asymptotic thresholds (sometimes referred to as \emph{phase transitions} \cite{Ame14,Wu12}) on the number of measurements.
\item As evidenced in our examples outlined below, our framework often leads to such exact or near-exact thresholds for significantly more general scalings of $k$, SNR, etc.~compared to previous works.
\item The majority of previous works have developed converse results using Fano's inequality, leading to necessary conditions for $\PP[\mathrm{error}]\to0$. In contrast, our converse results provide necessary conditions for $\PP[\mathrm{error}]\not\to1$. The distinction between these two conditions is important from a practical perspective: One may not expect a condition such as $\PP[\mathrm{error}] \ge 10^{-10}$ to be significant, whereas the condition $\PP[\mathrm{error}]\to1$ is inarguably so.
\end{enumerate}
\addtolength{\abovedisplayskip}{-2ex}
\addtolength{\belowdisplayskip}{-2ex}
\begin{table*}
\begin{centering}
\begin{tabular}{|>{\centering}m{1.5cm}|>{\centering}m{0.8cm}|>{\centering}m{2.3cm}|>{\centering}m{1.6cm}|>{\centering}m{4cm}|>{\centering}m{3.8cm}|}
\hline
Model & Result & Parameters & Distributions & Sufficient $n$ for $\PP[\mathrm{error}] \to 0$ & Necessary $n$ for $\PP[\mathrm{error}] \not\to 1$\tabularnewline
\hline
\multirow{2}{1.5cm}{\vspace*{-0.7cm}\centering Linear} & Cor. \ref{cor:linear_genK} & $k=o(p)$ & Discrete $\beta_{S}$
Gaussian $\mathbf{X}$ &
\[
\max_{\ell=1,\dotsc k}\frac{(1+\Delta_{\ref*{cor:linear_genK}})\log{p-k \choose \ell}}{f_{\ref*{cor:linear_genK}}(\ell)}
\]
{\medskip\scriptsize $(\Delta_{\ref*{cor:linear_genK}}\to0$ for various scalings)} &
\[
\max_{\ell=1,\dotsc k}\frac{\log{p-k+\ell \choose \ell}}{f_{\ref*{cor:linear_genK}}(\ell)}
\]
\tabularnewline
\cline{2-6}
& Cor. \ref{cor:linear_partial} & $k\to\infty,k=o(p)$
Partial recovery of proportion $1-\alpha^*$ & Gaussian $\beta_{S}$
Gaussian $\mathbf{X}$ &
\[
\max_{\alpha\in[\alpha^*,1]}\frac{\alpha k\log\frac{p}{k}}{f_{\ref*{cor:linear_partial}}(\alpha)}
\]
&
\[
\max_{\alpha\in[\alpha^*,1]}\frac{(\alpha-\alpha^*)k\log\frac{p}{k}}{f_{\ref*{cor:linear_partial}}(\alpha)}
\]
\tabularnewline
\hline
& Cor. \ref{cor:1bit_fixedK} & $k=\Theta(1)$
Low SNR & Discrete $\beta_{S}$
Gaussian $\mathbf{X}$ &
\[
\max_{\ell=1,\dotsc k}\frac{\ell\log p}{f_{\ref*{cor:1bit_fixedK}}(\ell)}
\]
{\scriptsize \medskip (within a factor $\frac{\pi}{2}$ of linear model)} &
\[
\max_{\ell=1,\dotsc k}\frac{\ell\log p}{f_{\ref*{cor:1bit_fixedK}}(\ell)}
\]
{\scriptsize \medskip (within a factor $\frac{\pi}{2}$ of linear model)}\tabularnewline
\cline{2-6}
1-bit & Cor. \ref{cor:1bit_genK} & $k=\Theta(p)$
High SNR & Fixed $\beta_{S}$
Gaussian $\mathbf{X}$ & - & \smallskip $\Omega(p\sqrt{\log p})$
{\scriptsize(compared to $\Theta(p)$ for linear model)}\tabularnewline
\cline{2-6}
& Cor. \ref{cor:1bit_partial} & $k \to \infty, k=o(p)$
Partial recovery of proportion $1-\alpha^*$ & Gaussian $\beta_{S}$
Gaussian $\mathbf{X}$ &
\[
\max_{\alpha\in[\alpha^*,1]}\frac{\alpha k\log\frac{p}{k}}{f_{\ref*{cor:1bit_partial}}(\alpha)}
\]
&
\[
\max_{\alpha\in[\alpha^*,1]}\frac{(\alpha-\alpha^*)k\log\frac{p}{k}}{f_{\ref*{cor:1bit_partial}}(\alpha)}
\]
\tabularnewline
\hline
\multirow{3}{1.5cm}{ {\vspace*{-1.3cm}\centering Group~testing}} & Cor. \ref{cor:gt} & $k=\Theta(p^{\theta})$ & Fixed $\beta_{S}$
Bernoulli $\Xv$ &
\[
\frac{k\log\frac{p}{k}}{f_{\ref*{cor:gt}}(\theta)}
\]
{\scriptsize ($f_{\ref*{cor:gt}}(\theta) = \log2$ for $\theta \le \frac{1}{3}$)}
&
\[
\frac{k\log\frac{p}{k}}{\log 2}
\]
\tabularnewline
\cline{2-6}
& Cor. \ref{cor:gt_noisy} & $k=\Theta(p^{\theta})$
Noisy (crossover probability $\rho$) & Fixed $\beta_{S}$
Bernoulli $\mathbf{X}$ &
\[
\frac{k\log\frac{p}{k}}{f_{\ref*{cor:gt_noisy}}(\theta)}
\]
{\scriptsize ($f_{\ref*{cor:gt_noisy}}(\theta) = \log2 - H_2(\rho)$ for small $\theta$)}
&
\[
\frac{k\log\frac{p}{k}}{\log 2 - H_2(\rho)}
\]
\tabularnewline
\cline{2-6}
& Cor. \ref{cor:gt_partial} & $k\to\infty,k=o(p)$
Partial recovery of proportion $1-\alpha^*$ & Fixed $\beta_{S}$
Bernoulli $\mathbf{X}$ &
\[
\frac{k\log\frac{p}{k}}{\log 2 - H_2(\rho)}
\]
&
\[
\frac{(1-\alpha^{*})\big(k\log\frac{p}{k}\big)}{\log 2 - H_2(\rho)}
\]
\tabularnewline
\hline
General discrete observations & Cor. \ref{cor:gen_finite} & Arbitrary & Arbitrary & - &
\[
\max_{\ell=1,\dotsc k}\frac{\log{p-k+\ell \choose \ell}}{f_{\ref*{cor:gen_finite}}(\ell)+\Delta_{\ref*{cor:gen_finite}}}
\]
\tabularnewline
\hline
\end{tabular}
\par
\end{centering}
\caption{Overview of main results for exact or partial support recovery under various observation models. In the necessary and sufficient number of measurements, asymptotically negligible terms have been omitted. All quantities are defined precisely in Section \ref{sec:EXAMPLES}. } \label{tbl:overview}
\end{table*}
\addtolength{\abovedisplayskip}{2ex}
\addtolength{\belowdisplayskip}{2ex}
{\bf Contributions for Specific Models:} An overview of our bounds for specific models is given in Table \ref{tbl:overview}, where we state the derived bounds with the asymptotically negligible terms omitted. All of the models and their parameters are defined precisely in Section \ref{sec:EXAMPLES}; in particular, the functions $f_1,\dotsc,f_{\ref*{cor:gen_finite}}$ and the remainder terms $(\Delta_{\ref*{cor:linear_genK}},\Delta_{\ref*{cor:gen_finite}})$ are given explicitly, and are easy to evaluate. We proceed by discussing these contributions in more detail, and comparing them to various existing results in the literature:
\begin{enumerate}
\item \emph{(Linear model)} In the case of exact recovery, we recover the exact thresholds on the required number of measurements given by Jin \emph{et al.} \cite{Jin11}, as well as handling a broader range of scalings of $\beta_{\mathrm{min}} := \min\{|\beta_i| \,:\, \beta_i \ne 0\}$ (see Section \ref{sec:GAUSSIAN_DISC} for details) and strengthening the converse by considering the more stringent condition $\PP[\mathrm{error}] \to 1$. Our results for partial recovery provide near-matching necessary and sufficient conditions under scalings with $k=o(p)$, thus complementing the extensive study of the scaling $k=\Theta(p)$ by Reeves and Gastpar \cite{Ree12,Ree13}.
\item \emph{(1-bit model)} We provide two surprising observations regarding the 1-bit model: Corollary \ref{cor:1bit_fixedK} provides a low-SNR setting where the quantization only increases the asymptotic number of measurements by a factor of $\frac{\pi}{2}$, whereas Corollary \ref{cor:1bit_genK} provides a high-SNR setting where the scaling law is strictly worse than the linear model. Similar behavior will be observed for partial recovery (Corollaries \ref{cor:linear_partial} and \ref{cor:1bit_partial}) by numerically comparing the bounds for various SNR values.
\item \emph{(Group testing)} Asymptotic thresholds for group testing with $k=\Theta(1)$ were given previously by Malyutov \cite{Mal78} and Atia and Saligrama \cite{Ati12}. However, for the case that $k\to\infty$, the sufficient conditions of \cite{Ati12} that introduced additional logarithmic factors. In contrast, we obtain matching $\Theta\big(k\log\frac{p}{k}\big)$ scaling laws for any sublinear scaling of the form $k=O(p^\theta)$ ($\theta\in(0,1)$). Moreover, for sufficiently small $\theta$ we obtain exact thresholds. In particular, for the noiseless setting we show that $n \approx k\log_2\frac{p}{k}$ measurements are both necessary and sufficient for $\theta\le\frac{1}{3}$. This is in fact the same threshold as that for adaptive group testing \cite{Bal13}, thus proving that non-adaptive Bernoulli measurement matrices are \emph{asymptotically optimal} even when adaptivity is allowed; this was previously known only in the limit as $\theta\to0$ \cite{Ald14}. For the noisy case, we prove an analogous claim for sufficiently small $\theta$. A shortened and simplified version of this paper focusing exclusively on group testing can be found in \cite{Sca15b}.
\item \emph{(General discrete observations)} Our converse for the case of general discrete observations (Corollary \ref{cor:gen_finite}) recovers that of Tan and Atia \cite{Tan14} for the case that $\beta_S$ is fixed, strengthens it due to a smaller remainder term $\Delta_{\ref*{cor:gen_finite}}$, and provides a generalization to the case that $\beta_S$ is random.
\end{enumerate}
\subsection{Structure of the Paper}
In Section \ref{sec:SETUP}, we introduce our system model. In Section \ref{sec:GENERAL}, we present our main non-asymptotic achievability and converse results for general observation models, and the procedure for applying them to specific problems. Several applications of our results to specific models are presented in Section \ref{sec:EXAMPLES}. The proofs of the general bounds are given in Section \ref{sec:PROOFS}, and conclusions are drawn in Section \ref{sec:CONCLUSION}.
\subsection{Notation} \label{sec:NOTATION}
We use upper-case letters for random variables, and lower-case variables for their realizations. A non-bold character may be a scalar or a vector, whereas a bold character refers to a collection of $n$ scalars (e.g., $\Yv \in \RR^n$) or vectors (e.g., $\Xv \in \RR^{n\times p}$). We write $\beta_{S}$ to denote the subvector of $\beta$ at the columns indexed by $S$, and $\Xv_{S}$ to denote the submatrix of $\Xv$ containing the columns indexed by $S$. The complement with respect to $\{1,\dotsc,p\}$ is denoted by $(\cdot)^c$.
The symbol $\sim$ means ``distributed as''. For a given joint distribution $P_{XY}$, the corresponding marginal distributions are denoted by $P_{X}$ and $P_{Y}$, and similarly for conditional marginals (e.g., $P_{Y|X}$). We write $\PP[\cdot]$ for probabilities, $\EE[\cdot]$ for expectations, and $\var[\cdot]$ for variances. We use usual notations for the entropy (e.g., $H(X)$) and mutual information (e.g., $I(X;Y)$), and their conditional counterparts (e.g., $H(X|Z)$, $I(X;Y|Z)$). Note that $H$ may also denote the differential entropy for continuous random variables; the distinction will be clear from the context. We define the binary entropy function $H_2(\rho) := -\rho\log\rho - (1-\rho)\log(1-\rho)$, and the Q-function $Q(x) := \PP[W \ge x]$ ($W \sim N(0,1)$).
We make use of the standard asymptotic notations $O(\cdot)$, $o(\cdot)$, $\Theta(\cdot)$, $\Omega(\cdot)$ and $\omega(\cdot)$. We define the function $[\cdot]^+ = \max\{0,\cdot\}$, and write the floor function as $\lfloor\cdot\rfloor$. The function $\log$ has base $e$.
\section{Problem Setup} \label{sec:SETUP}
\subsection{Model and Assumptions}
Recall that $p$ denotes the ambient dimension, $k$ denotes the sparsity level, and $n$ denotes the number of measurements. We let $\Sc$ be the set of subsets of $\{1,\dotsc,p\}$ having cardinality $k$. The key random variables in our setup are the support set $S \in \Sc$, the data vector $\beta\in\RR^p$, the measurement matrix $\Xv\in\RR^{n \times p}$, and the observation vector $\Yv \in \RR^{n}$.\footnote{Extensions to more general alphabets beyond $\RR$ are straightforward.}
The support set $S$ is assumed to be equiprobable on the $p \choose k$ subsets within $\Sc$. Given $S$, the entries of $\beta_{S^c}$ are deterministically set to zero, and the remaining entries are generated according to some distribution $\beta_S \sim P_{\beta_S}$. We assume that these non-zero entries follow the same distribution for all of the $p \choose k$ possible realizations of $S$, and that this distribution is permutation-invariant.
The measurement matrix $\Xv$ is assumed to have i.i.d.~values on some distribution $P_{X}$. We write $P_{X}^{n \times p}$, to denote the corresponding i.i.d.~distributions for matrices, and we write $P_{X}^k$ as a shorthand for $P_{X}^{k \times 1}$. Given $S$, $\Xv$, and $\beta$, each entry of the observation vector $\Yv$ is generated in a conditionally independent manner, with the $i$-th entry $Y^{(i)}$ distributed according to
\begin{equation}
(Y^{(i)}|S=s,X^{(i)}=x^{(i)},\beta=b) \sim P_{Y|X_S\beta_S}(\,\cdot\,|x^{(i)}_s,b_s), \label{eq:single_output}
\end{equation}
for some conditional distribution $P_{Y|X_{S}\beta_S}$. We again assume symmetry with respect to $S$, namely, that $P_{Y|X_{S}\beta_S}$ does not depend on the specific realization, and that the distribution is invariant when the columns of $X_S$ and the entries of $\beta_S$ undergo a common permutation.
Given $\Xv$ and $\Yv$, a decoder forms an estimate $\hat{S}$ of $S$. Similarly to previous works studying information-theoretic limits on support recovery, we assume that the decoder knows the system model. We consider two related performance measures. In the case of exact support recovery, the error probability is given by
\begin{equation}
\pe := \PP[\hat{S} \ne S], \label{eq:pe}
\end{equation}
and is taken with respect to the realizations of $S$, $\beta$, $\Xv$, and $\Yv$; the decoder is assumed to be deterministic. We also consider a less stringent performance criterion requiring that only $k - \dmax$ entries of $S$ are successfully recovered, for some $\dmax \in \{1,\dotsc,k-1\}$. Following \cite{Ree12,Ree13}, the error probability is given by
\begin{equation}
\pe(\dmax) := \PP\big[|S \backslash \hat{S}| > \dmax \cup |\hat{S} \backslash S| > \dmax\big]. \label{eq:pe_partial}
\end{equation}
Note that if both $S$ and $\hat{S}$ have cardinality $k$ with probability one, then the two events in the union are identical, and hence either of the two can be removed.
For clarity, we formally state our main assumptions as follows:
{ \setenumerate[1]{label=[A\arabic*]}
\begin{enumerate}
\item The support set $S$ is uniform on the ${p \choose k}$ subsets of $\{1,\dotsc,p\}$ of size $k$, and the measurement matrix $\Xv$ is i.i.d.~on some distribution $P_X$.
\item The non-zero entries $\beta_S$ are distributed according to $P_{\beta_S}$, and this distribution is permutation-invariant and the same for all realizations of $S$.
\item The observation vector $\Yv$ is conditionally i.i.d.~according to $P_{Y|X_S\beta_S}$, and this distribution is the same for all realizations of $S$, and invariant to common permutations of the columns of $X_S$ and entries of $\beta_S$.
\item The decoder is given $(\Xv,\Yv)$, and also knows the system model including $k$, $P_{Y|X_S\beta_S}$, and $P_{\beta_S}$.
\end{enumerate} }
Our main goal is to derive necessary and sufficient conditions on $n$ and $k$ (as functions of $p$) such that $\pe$ or $\pe(\dmax)$ vanishes as $p\to\infty$. Moreover, when considering converse results, we will not only be interested in conditions under which $\pe \not\to 0$, but also conditions under which the stronger statement $\pe\to1$ holds.
In particular, we introduce the terminology that the \emph{strong converse} holds if there exists a sequence of values $n^{*}$, indexed by $p$, such that for all $\eta > 0$, we have $\pe\to0$ when $n \ge n^{*} (1+\eta)$, and $\pe\to1$ when $n \le n^{*} (1-\eta)$. This is related to the notion of a \emph{phase transition} \cite{Ame14,Wu12}. More generally, we will refer to conditions under which $\pe\to1$ as \emph{strong impossibility results}, not necessarily requiring matching achievability bounds. That is, the strong converse conclusively gives a sharp threshold between failure and success, whereas a strong impossibility result may not.
It will prove convenient to work with random variables that are implicitly conditioned on a fixed value of $S$, say $s=\{1,\dotsc,k\}$. We write $P_{\beta_s}$ and $P_{Y|X_{s}\beta_s}$ in place of $P_{\beta_S}$ and $P_{Y|X_{S}\beta_S}$ to emphasize that $S=s$. Moreover, we define the corresponding joint distribution
\begin{equation}
P_{\beta_s X_s Y}(b_s,x_s,y) := P_{\beta_s}(b_s)P_{X}^{k}(x_s)P_{Y|X_s\beta_s}(y|x_s,b_s), \label{eq:distr_sl}
\end{equation}
and its multiple-observation counterpart
\begin{equation}
P_{\beta_s\Xv_s\Yv}(b_s,\xv_s,\yv) := P_{\beta_s}(b_s)P_X^{n \times k}(\xv_s) P^{n}_{Y|X_s\beta_s}(\yv|\xv_{s},b_s). \label{eq:vector_distr}
\end{equation}
where $P^{n}_{Y|X_s\beta_s}(\cdot|\cdot,b_s)$ is the $n$-fold product of $P_{Y|X_s\beta_s}(\cdot|\cdot,b_s)$.
Except where stated otherwise, the random variables $(\beta_s,X_s,Y)$ and $(\beta_s,\Xv_s,\Yv)$ appearing throughout this paper are distributed as
\begin{align}
(\beta_s,X_s,Y) &\sim P_{\beta_s X_s Y} \label{eq:joint_dist} \\
(\beta_s,\Xv_s,\Yv) &\sim P_{\beta_s\Xv_s\Yv}, \label{eq:joint_dist_n}
\end{align}
with the remaining entries of the measurement matrix being distributed as $\Xv_{s^c} \sim P_X^{n \times (p-k)}$, and with $\beta_{s^c}=0$ deterministically. That is, we condition on a fixed $S=s$ except where stated otherwise.
For notational convenience, the main parts of our analysis are presented with $P_{\beta_s}$, $P_X$ and $P_{Y|X_s\beta_s}$ representing probability mass functions (PMFs), and with the corresponding averages written using summations. However, except where stated otherwise, our analysis is directly applicable to case that these distributions instead represent probability density functions (PDFs), with the summations replaced by integrals where necessary. The same applies to mixed discrete-continuous distributions.
\subsection{Information-Theoretic Definitions}
Before introducing the required definitions for support recovery, it is instructive to discuss thresholding techniques in channel coding studies. These commenced in early works such as \cite{Fei54,Sha57}, and have recently been used extensively in information-spectrum methods \cite{Ver94, Han03}.
\subsubsection{Channel Coding}
We first recall the mutual information, which is ubiquitous in information theory:
\begin{equation}
I(X;Y) := \sum_{x,y} P_{XY}(x,y)\log\frac{P_{Y|X}(y|x)}{P_{Y}(y)}.
\end{equation}
In deriving asymptotic and non-asymptotic performance bounds, it is common to work directly with the logarithm,
\begin{equation}
\imath(x;y) := \log\frac{P_{Y|X}(y|x)}{P_{Y}(y)},
\end{equation}
which is commonly known as the \emph{information density}. The thresholding techniques work by manipulating probabilities of events of the form $\sum_{i=1}^{n} \imath(X_i;Y_i) \le \gamma$ and $\sum_{i=1}^{n} \imath(X_i;Y_i) > \gamma$. For the former, one can perform a \emph{change of measure} from the conditional distribution $\Yv$ given $\Xv$ to the unconditional distribution of $\Yv$, with a multiplicative constant $e^{-\gamma}$. For the latter, one can similarly perform a change of measure from $\Yv$ to $(\Yv|\Xv)$. Hence, in both cases, there is a simple relation between the conditional and unconditional probabilities of the output sequences.
Using these methods, one can get upper and lower bounds on the error probability such that the dominant term is
\begin{equation}
\PP\bigg[ \frac{1}{n}\sum_{i=1}^{n} \imath(X_i;Y_i) \le I(X;Y) + \zeta_n \bigg]
\end{equation}
for some $\zeta_n = o(1)$. Assuming that $\{(X_i,Y_i)\}_{i=1}^n$ has some form of i.i.d.~structure, one can analyze this expression using tools from probability theory. The law of large numbers yields the channel capacity $C=\max_{P_X}I(X;Y)$, and refined characterizations can be obtained using variations of the central limit theorem \cite{Pol10}.
Among the channel coding literature, our analysis is most similar to that of mixed channels \cite[Sec.~3.3]{Han03}, where the relation between the input and output sequences is not i.i.d., but instead conditionally i.i.d.~given another random variable. In our setting, $\beta_s$ will play the role of this random variable. See Figure \ref{fig:channel_diagram} for a depiction of this connection.
\begin{figure}
\begin{centering}
\includegraphics[width=1\columnwidth]{channel_diagram.pdf}
\par
\end{centering}
\caption{ Connection between support recovery and coding over a mixed channel. \label{fig:channel_diagram}}
\end{figure}
\subsubsection{Support Recovery}
As in \cite{Ati12,Aks13}, we will consider partitions of the support set $s \in \Sc$ into two sets $\sdif \ne \emptyset$ and $\seq$. As will be seen in the proofs, $\seq$ will typically correspond to an overlap between $s$ and some other set $\sbar$ (i.e., $s \cap \sbar$), whereas $\sdif$ will correspond to the indices in one set but not the other (e.g., $s \backslash \sbar$). There are $2^k - 1$ ways of performing such a partition with $\sdif \ne \emptyset$.
For fixed $s\in\Sc$ and a corresponding pair $(\sdif,\seq)$, we introduce the notation
\begin{align}
P_{Y|\Xv_{\sdif}\Xv_{\seq}}(\yv|\xv_{\sdif},\xv_{\seq}) &:= P_{\Yv|\Xv_s}(\yv|\xv_s) \label{eq:p_split1} \\
P_{Y|X_{\sdif}X_{\seq}\beta_s}(y|x_{\sdif},x_{\seq},b_s) &:= P_{Y|X_s\beta_s}(y|x_s,b_s), \label{eq:p_split2}
\end{align}
where $P_{\Yv|\Xv_s}$ is the marginal distribution of \eqref{eq:vector_distr}. While the left-hand sides of \eqref{eq:p_split1}--\eqref{eq:p_split2} represent the same quantities for any such $(\sdif,\seq)$, it will still prove convenient to work with these in place of the right-hand sides. In particular, this allows us to introduce the marginal distributions
\begin{align}
& P_{\Yv|\Xv_{\seq}}(\yv|\xv_{\seq}) \nonumber \\
&:= \sum_{\xv_{\sdif}} P_X^{n \times \ell}(\xv_{\sdif})P_{Y|\Xv_{\sdif}\Xv_{\seq}}(\yv|\xv_{\sdif},\xv_{\seq}) \\
& P_{Y|X_{\seq}\beta_s}(y|x_{\seq},b_s) \nonumber \\
&:= \sum_{x_{\sdif}} P_X^{\ell}(x_{\sdif})P_{Y|X_{\sdif}X_{\seq}\beta_s}(y|x_{\sdif},x_{\seq},b_s),
\end{align}
where $\ell := |\sdif|$. Using the preceding definitions, we introduce two information densities. The first contains probabilities averaged over $\beta_s$,
\begin{equation}
\imath(\xv_{\sdif}; \yv | \xv_{\seq} ) := \log\frac{ P_{\Yv|\Xv_{\sdif}\Xv_{\seq}}(\yv|\xv_{\sdif},\xv_{\seq}) }{ P_{\Yv|\Xv_{\seq}}(\yv|\xv_{\seq}) }, \label{eq:idens_multi}
\end{equation}
whereas the second conditions on $\beta_s=b_s$:
\begin{equation}
\imath^n(\xv_{\sdif}; \yv | \xv_{\seq}, b_s ) := \sum_{i=1}^{n} \imath(x^{(i)}_{\sdif}; y^{(i)} | x^{(i)}_{\seq}, b_s), \label{eq:idens_bn}
\end{equation}
where the single-letter information density is
\begin{equation}
\imath(x_{\sdif}; y | x_{\seq}, b_s) := \log\frac{ P_{Y|X_{\sdif}X_{\seq}\beta_s}(y|x_{\sdif},x_{\seq},b_s) }{ P_{Y|X_{\seq}\beta_s}(y|x_{\seq},b_s) }. \label{eq:idens_b}
\end{equation}
As mentioned above, we will generally work with discrete random variables for clarity of exposition, in which case the ratio is between two PMFs. In the case of continuous observations the ratio is instead between two PDFs, and more generally this can be replaced by the Radon-Nikodym derivative as in the channel coding setting \cite{Pol10}.
Averaging \eqref{eq:idens_b} with respect to the random variables in \eqref{eq:joint_dist} conditioned on $\beta_s = b_s$ yields a conditional mutual information, which we denote by
\begin{equation}
I_{\sdif,\seq}(b_s) := I(X_{\sdif};Y|X_{\seq},\beta_s=b_s). \label{eq:condMI}
\end{equation}
This quantity will play a key role in our bounds, which will typically have the form
\begin{equation}
\pe \approx \PP\bigg[ n \le \max_{(\sdif,\seq)}\frac{{p \choose |\sdif|}}{ I_{\sdif,\seq}(\beta_s) } \bigg],
\end{equation}
as will be made more precise in the subsequent sections.
\section{General Achievability and Converse Bounds} \label{sec:GENERAL}
In this section, we provide general results holding for arbitrary models satisfying the assumptions given in Section \ref{sec:SETUP}. Each of the results for exact recovery has a direct counterpart for partial recovery. For clarity, we focus on the former throughout Sections \ref{sec:ACHIEVABILITY} and \ref{sec:APPLYING}, and then proceed with the latter in Section \ref{sec:PARTIAL}.
\subsection{Initial Non-Asymptotic Bounds} \label{sec:ACHIEVABILITY}
Here we provide our main non-asymptotic upper and lower bounds on the error probability. These bounds bear a strong resemblance to analogous bounds from the channel coding literature \cite{Han03}; in each case, the dominant term involves tail probabilities of the information density given in \eqref{eq:idens_bn}. The mean of the information density is the mutual information in \eqref{eq:condMI}, which thus arises naturally in the subsequent necessary and sufficient conditions on $n$ upon showing that the deviation from the mean is small with high probability. The procedure for doing this given a specific model will be given in Section \ref{sec:APPLYING}.
We start with our achievability result. Here and throughout this section, we make use of the random variables defined in \eqref{eq:joint_dist_n}.
\begin{thm} \label{thm:ach1}
For any constants $\delta_1 > 0$ and $\gamma$, there exists a decoder such that
\begin{multline}
\pe \le \PP\bigg[ \bigcup_{(\sdif,\seq)\,:\,\sdif\ne\emptyset} \bigg\{ \imath^n(\Xv_{\sdif}; \Yv | \Xv_{\seq}, \beta_s ) \\ \le \log {{p-k} \choose |\sdif|} + \log\bigg(\frac{k^2}{\delta_1^2}{k \choose |\sdif|}^2\bigg) + \gamma \bigg\} \bigg] + P_0(\gamma) + 2\delta_1, \label{eq:thresh_sl}
\end{multline}
where
\begin{equation}
P_0(\gamma) := \PP\bigg[ \log \frac{P_{\Yv|\Xv_s,\beta_s}(\Yv|\Xv_s,\beta_s)}{P_{\Yv|\Xv_{s}}(\Yv|\Xv_{s})} > \gamma \bigg]. \label{eq:P_0}
\end{equation}
\end{thm}
\begin{proof}
See Section \ref{sec:PROOF_GEN_ACH}.
\end{proof}
\begin{rem} \label{rem:P_0}
The probability in the definition of $P_0(\gamma)$ is not an i.i.d.~sum, and the techniques for ensuring that $P_0(\gamma) \to 0$ vary between different settings. The following approaches will suffice for all of the applications in this paper:
\begin{enumerate}
\item In the case that $P_{\beta_S}$ is discrete, $P_{\Yv|\Xv_{s}}(\yv|\xv_{s}) = \sum_{b_s} P_{\beta_s}(b_s) P_{\Yv|\Xv_s,\beta_s}(\yv|\xv_s,b_s)$, and it follows that
\begin{equation}
\gamma = \log\frac{1}{\min_{b_s}P_{\beta_s}(b_s)} \implies P_0(\gamma) = 0. \label{eq:discrete_gamma}
\end{equation}
Moreover, this can be strengthened by noting from the proof of Theorem \ref{thm:ach1} that $\gamma$ may depend on $\beta_s$, and choosing $\gamma(b_s) = \log\frac{1}{P_{\beta_s}(b_s)}$ accordingly.
\item Defining
\begin{align}
I_0 &:= I(\beta_s;\Yv|\Xv_{s}) \label{eq:I_0} \\
V_0 &:= \var\bigg[\log \frac{P_{\Yv|\Xv_s,\beta_s}(\Yv|\Xv_{s},\beta_s)}{P_{\Yv|\Xv_{s}}(\Yv|\Xv_{s})}\bigg], \label{eq:V_0}
\end{align}
we have for any $\delta_0 > 0$ that
\begin{equation}
\gamma = I_0 + \sqrt{\frac{V_0}{\delta_0}} \implies P_0(\gamma) \le \delta_0. \label{eq:gamma_cheby}
\end{equation}
This follows directly from Chebyshev's inequality.
\item Defining
\begin{equation}
I_{0,+} := \EE\bigg[\bigg[\log \frac{P_{\Yv|\Xv_s\beta_s}(\Yv|\Xv_{s},\beta_s)}{P_{\Yv|\Xv_{s}}(\Yv|\Xv_{s})}\bigg]^+\bigg], \label{eq:I_0+}
\end{equation}
we have for any $\delta_0 > 0$ that
\begin{equation}
\gamma = \frac{I_{0,+}}{\delta_0} \implies P_0(\gamma) \le \delta_0. \label{eq:gamma_markov}
\end{equation}
This follows directly from Markov's inequality.
\end{enumerate}
\end{rem}
The proof of Theorem \ref{thm:ach1} is based on a decoder the searches for a unique support set $s$ such that
\begin{equation}
\imath(\xv_{\sdif}; \yv | \xv_{\seq} ) > \gamma_{|\sdif|} \label{eq:decoder_cond}
\end{equation}
for some $\{\gamma_{\ell}\}_{\ell=1}^k$ and all $2^{k} - 1$ partitions $(\sdif,\seq)$ of $s$ with $\sdif\ne\emptyset$. Since the numerator in \eqref{eq:idens_multi} is the likelihood of $\yv$ given $(\xv_{\sdif},\xv_{\seq})$, this decoder can be thought of a weakened version of the maximum-likelihood (ML) decoder. Like the ML decoder, computational considerations make its implementation intractable.
The following theorem provides a general non-asymptotic converse bound.
\begin{thm} \label{thm:conv}
Fix $\delta_1 > 0$, and let $(\sdif(b_s),\seq(b_s))$ be an arbitrary partition of $s=\{1,\dotsc,k\}$ (with $\sdif\ne\emptyset$) depending on $b_s\in\RR^k$. For any decoder, we have
\begin{multline}
\pe \ge \PP\bigg[ \imath^n(\Xv_{\sdif(\beta_s)};\Yv|\Xv_{\seq(\beta_s)},\beta_s) \\ \le \log {{p-k+|\sdif(\beta_s)|} \choose |\sdif(\beta_s)|} + \log\delta_1 \bigg] - \delta_1. \label{eq:conv_sl}
\end{multline}
\end{thm}
\begin{proof}
See Section \ref{sec:PROOF_GEN_CONV}.
\end{proof}
The proof of Theorem \ref{thm:conv} is based on Verd\'u-Han type bounding techniques \cite{Ver94}.
\subsection{Techniques for Applying Theorems \ref{thm:ach1} and \ref{thm:conv}} \label{sec:APPLYING}
The bounds presented in the preceding theorems do not directly reveal the number of measurements required to achieving a vanishing error probability. In this subsection, we present the steps that can be used to obtain such conditions. We provide examples in Section \ref{sec:EXAMPLES}.
The idea is to use a concentration inequality to bound the first term in \eqref{eq:thresh_sl} (or \eqref{eq:conv_sl}), which is possible due to the fact that each summation $\imath^n$ is conditionally i.i.d.~given $\beta_s$. We proceed by providing the details of these steps separately for the achievability and converse. We start with the former.
\begin{enumerate}
\item Observe that, conditioned on $\beta_s=b_s$, the mean of $\imath^n(\Xv_{\sdif}; \Yv | \Xv_{\seq}, \beta_s )$ is $nI_{\sdif,\seq}(b_s)$, where $I_{\sdif,\seq}(b_s)$ is defined in \eqref{eq:condMI}.
\item Fix $\delta_2 \in (0,1)$, and suppose that for a fixed value $b_s$ of $\beta_s$, we have for all $(\sdif,\seq)$ that
\begin{multline}
\log{{p-k} \choose |\sdif|} + \log\bigg(\frac{k^2}{\delta_1^2}{k \choose |\sdif|}^2\bigg) + \gamma \\ \le n(1 - \delta_2) I_{\sdif,\seq}(b_s), \label{eq:B_cond_ach0}
\end{multline}
and
\begin{multline}
\hspace*{-6ex}\PP\Big[ \imath^n(\Xv_{\sdif}; \Yv | \Xv_{\seq}, b_s) \le n(1 - \delta_2)I_{\sdif,\seq}(b_s) \,\big|\, \beta_s = b_s \Big] \\ \le \psi_{|\sdif|}(n,\delta_2) \label{eq:psi_ach}
\end{multline}
for some functions $\{\psi_\ell\}_{\ell=1}^{k}$ (e.g., these may arise from Chebyshev's inequality or Bernstein's inequality \cite[Ch.~2]{Bou13}). Combining these conditions with the union bound, we obtain
\begin{multline}
\hspace*{-3ex} \PP\bigg[ \bigcup_{(\sdif,\seq)\,:\,\sdif\ne\emptyset} \bigg\{ \imath^n(\Xv_{\sdif}; \Yv | \Xv_{\seq}, \beta_s ) \\ \le \log {{p-k} \choose |\sdif|} + \log\bigg(\frac{k^2}{\delta_1^2}{k \choose |\sdif|}^2\bigg) + \gamma \bigg\} \,\Big|\,\beta_s=b_s\bigg] \\ \le \sum_{\ell=1}^k{k \choose \ell} \psi_\ell(n,\delta_2).
\end{multline}
\item Observe that the condition in \eqref{eq:B_cond_ach0} can be written as
\begin{equation}
n \ge \frac{ \log{{p-k} \choose |\sdif|} + \log\Big(\frac{k^2}{\delta_1^2}{k \choose |\sdif|}^2\Big) + \gamma }{ I_{\sdif,\seq}(b_s) (1-\delta_2) }. \label{eq:final_ach}
\end{equation}
\end{enumerate}
\noindent We summarize the preceding findings in the following.
\begin{thm} \label{thm:simplified_ach}
For any constants $\delta_1>0$, $\delta_2\in(0,1)$ and $\gamma$, and functions $\{\psi_{\ell}\}_{\ell=1}^{k}$ ($\psi_\ell : \ZZ\times\RR\to\RR$), define the set
\begin{multline}
\Bc(\delta_1,\delta_2,\gamma) := \big\{ b_s \,:\, \text{\eqref{eq:psi_ach} and \eqref{eq:final_ach} hold for all } \\ (\sdif,\seq)\text{ with }\sdif\ne\emptyset\big\}. \label{eq:setB}
\end{multline}
Then we have
\begin{equation}
\pe \le \PP\big[ \beta_s \notin \Bc(\delta_1,\delta_2,\gamma)\big] + \sum_{\ell=1}^k{k \choose \ell}\psi_\ell(n,\delta_2) + P_0(\gamma) + 2\delta_1. \label{eq:nonasymp_ach}
\end{equation}
\end{thm}
\begin{rem} \label{rem:delta2}
The preceding arguments remain unchanged when $\delta_2$ also depends on $\ell = |\sdif|$. We leave this possible dependence implicit throughout this section, since a fixed value will suffice for all but one of the models considered in Section \ref{sec:EXAMPLES}.
\end{rem}
In the case that \eqref{eq:psi_ach} holds for all $b_s$ (or more generally, within a set whose probability under $P_{\beta_s}$ tends to one) and the final three terms in \eqref{eq:nonasymp_ach} vanish, the overall upper bound approaches the probability, with respect to $P_{\beta_s}$, that \eqref{eq:final_ach} fails to hold. In many cases, the second logarithm in the numerator therein is dominated by the first. It should be noted that the condition that the second term in \eqref{eq:nonasymp_ach} vanishes can also impose conditions on $n$. For most of the examples presented in Section \ref{sec:EXAMPLES}, the condition in \eqref{eq:final_ach} will be the dominant one; however, this need not always be the case, and it depends on the concentration inequality used in \eqref{eq:psi_ach}.
The application of Theorem \ref{thm:conv} is done using similar steps, so we provide less detail. Fix $\delta_2 > 0$, and suppose that, for a fixed value $b_s$ of $\beta_s$, the pair $(\sdif,\seq)=(\sdif(b_s),\seq(b_s))$ is such that
\begin{equation}
\log{{p-k+|\sdif|} \choose |\sdif|} - \log \delta_1 \ge n(1 + \delta_2)I_{\sdif,\seq}(b_s), \label{eq:B_cond_conv}
\end{equation}
and
\begin{multline}
\hspace*{-2ex}\PP\Big[ \imath^n(\Xv_{\sdif}; \Yv | \Xv_{\seq}, b_s) \le n(1 + \delta_2)I_{\sdif,\seq}(b_s) \,\big|\, \beta_s = b_s \Big] \\ \ge 1 - \psi'_{|\sdif|}(n,\delta_2) \label{eq:psi_conv}
\end{multline}
for some function $\psi'_{|\sdif|}$. Combining these conditions, we see that the first probability in \eqref{eq:conv_sl}, with an added conditioning on $\beta_s = b_s$, is lower bounded by $1 - \psi'_{|\sdif|}(n,\delta_2)$. In the case that $\psi'_{\ell}$ is defined for multiple $\ell$ values corresponding to different values of $b_s$, we can further lower bound this by $1 - \max_{\ell}\psi'_\ell(n,\delta_2)$.
Next, we observe that \eqref{eq:B_cond_conv} holds if and only if
\begin{equation}
n \le \frac{\log {{p-k+|\sdif|} \choose |\sdif|} - \log\delta_1 }{ I_{\sdif,\seq}(b_s)(1 + \delta_2) }. \label{eq:final_conv}
\end{equation}
Recalling that the partition $(\sdif,\seq)$ is an arbitrary function of $\beta_s$, we can ensure that this coincides with
\begin{equation}
n \le \max_{(\sdif,\seq) \,:\,\sdif\ne\emptyset} \frac{\log {{p-k+|\sdif|} \choose |\sdif|} - \log\delta_1 }{ I_{\sdif,\seq}(b_s) (1 + \delta_2) } \label{eq:final_conv_max}
\end{equation}
by choosing each pair $(\sdif,\seq)$ as a function of $b_s$ to achieve this maximum.
Finally, we note that the maximum over $\ell$ in the above-derived term $1 - \max_{\ell}\psi'_\ell(n,\delta_2)$ may be restricted to any set $\Lc \subseteq \{1,\dotsc,k\}$ provided that $|\sdif|$ is constrained similarly in \eqref{eq:final_conv_max}; one simply chooses the partition $(\sdif(b_s),\seq(b_s))$ so that $\ell = |\sdif|$ always lies in this set. Putting everything together, we have the following.
\begin{thm} \label{thm:simplified_conv}
For any set $\Lc \subseteq \{1,\dotsc,k\}$, constants $\delta_1>0$ and $\delta_2 > 0$, and functions $\{\psi'_{\ell}\}_{\ell\in\Lc}$ ($\psi'_\ell : \ZZ\times\RR\to\RR$), define the set
\begin{multline}
\Bc'(\delta_1,\delta_2) := \big\{ b_s \,:\, \text{\eqref{eq:psi_conv} and \eqref{eq:final_conv} hold for all } \\ (\sdif,\seq)\text{ with }|\sdif| \in \Lc \big\}. \label{eq:setB'}
\end{multline}
Then we have
\begin{equation}
\pe \ge \PP\big[ \beta_s \in \Bc'(\delta_1,\delta_2) \big]\Big(1 - \max_{\ell \in \Lc}\psi'_\ell(n,\delta_2)\Big) - \delta_1. \label{eq:nonasymp_conv}
\end{equation}
\end{thm}
If the pair $(\sdif,\seq)$ had been fixed in Theorem \ref{thm:conv}, as opposed to being a function of $\beta_s$, then we would have only obtained a weaker result with the statement ``for all $(\sdif,\seq)$'' in \eqref{eq:setB'} replaced by a fixed pair. Assuming that the remainder terms in \eqref{eq:nonasymp_conv} are insignificant, this weaker result is of the form $\pe \gtrsim \max_{(\sdif,\seq)} \PP\big[n \le f(\sdif,\seq,\beta_s)\big]$ rather than $\pe \gtrsim \PP\big[n \le \max_{(\sdif,\seq)} f(\sdif,\seq,\beta_s)\big]$. This can lead to significantly different bounds on the sample complexity, and the distinction is crucial in our applications in Section \ref{sec:EXAMPLES}. As described in the proof in Section \ref{sec:PROOFS}, the key to obtaining this difference is in applying a refined version of an argument based on a genie.
The general steps in applying Theorems \ref{thm:simplified_ach} and \ref{thm:simplified_conv} to specific problems are outlined in Procedure \ref{pr:procedure}.
\begin{algorithm}
\caption*{ \manuallabel{pr:procedure}{1} \textbf{Procedure 1:} Steps for Obtaining Necessary and Sufficient Conditions on $n$ from Theorems \ref{thm:simplified_ach} and \ref{thm:simplified_conv} }
\begin{enumerate}
\item \emph{(Identify a Typical Set)} Construct a sequence of ``typical'' sets $\Tc_{\beta} \subseteq \RR^k$ of non-zero entries, indexed by $p$, such that $\PP\big[\beta_s \in \Tc_{\beta}\big] \to 1$, thus restricting the vectors $b_s$ for which $\imath(X_{\sdif};Y|X_{\seq},b_s)$ needs to be characterized.
\item \emph{(Bound the Information Density Tail Probabilities)} Using a concentration inequality for i.i.d.~summations (e.g., Chebyshev, Bernstein), bound the tail probabilities in \eqref{eq:psi_ach} and \eqref{eq:psi_conv} for each $(\sdif,\seq)$ and $b_s \in \Tc_{\beta}$, with a fixed constant $\delta_2$. Upon making these dependent on $(\sdif,\seq,b_s)$ only through $\ell := |\sdif|$, the bounds are denoted by $\psi_{\ell}(n,\delta_2)$ and $\psi'_{\ell}(n,\delta_2)$.
\item \emph{(Control the Remainder Terms)} By suitable rearrangements, find conditions on $n$ under which the terms $\sum_{\ell}{k \choose \ell}\psi_\ell(n,\delta_2)$ and $\max_{\ell \in \Lc}\psi'_\ell(n,\delta_2)$ in \eqref{eq:nonasymp_ach} and \eqref{eq:nonasymp_conv} vanish, thus ensuring that their contribution is negligible. Similarly, choose $\delta_1$ to vanish with $p$ so that its contribution is negligible, and for the achievability part, choose $\gamma$ such that the remainder term $P_0(\gamma)$ vanishes (\emph{cf.}~Remark \ref{rem:P_0}).
\item \emph{(Combine and Simplify)} Combine the previous steps as follows:
\begin{enumerate}
\item Construct the set of non-zero entries $\Bc(\delta_1,\delta_2,\gamma) \subseteq \RR^{k}$ (respectively, $\Bc'(\delta_1,\delta_2)$) in \eqref{eq:setB} (respectively, \eqref{eq:setB'});
\item Deduce from \eqref{eq:nonasymp_ach} (respectively, \eqref{eq:nonasymp_conv}) and Step 3 that $\pe \le \PP[\beta_s\notin\Bc(\delta_1,\delta_2,\gamma)] + o(1)$ (respectively, $\pe \ge \PP[\beta_s\in\Bc'(\delta_1,\delta_2)] + o(1)$);
\item From the properties of the typical set $\Tc_{\beta}$ in Steps 1--2, deduce that $\pe\to0$ (respectively, $\pe\to1$) when $n$ satisfies \eqref{eq:final_ach} (respectively, \eqref{eq:final_conv}) for all $b_s \in \Tc_{\beta}$;
\item Augment this condition on $n$ with Step 3.
\end{enumerate}
\end{enumerate}
\end{algorithm}
In our experience, the choice of $\Tc_{\beta}$ in the first step of Procedure \ref{pr:procedure} usually comes naturally given the specific model. On the other hand, it is often less straightforward to find a sufficiently powerful concentration inequality in Step 2. A simple choice is Chebyshev's inequality, which expresses $\psi_{\ell}$ and $\psi'_{\ell}$ in terms of $I_{\sdif,\seq}(b_s)$ (see \eqref{eq:condMI}) and the corresponding variances of the information densities. This choice is usually effective for the converse, wheres the achievability part typically requires sharper concentration inequalities such as Bernstein's inequality, due to the combinatorial terms in \eqref{eq:nonasymp_ach}.
\subsection{Extensions to Partial Recovery} \label{sec:PARTIAL}
We now turn to the partial support recovery criterion in \eqref{eq:pe_partial}. The changes in the analysis required to generalize Theorems \ref{thm:ach1} and \ref{thm:conv} to this setting are given in Section \ref{sec:PROOF_PARTIAL}; rather than repeating each of these, we focus our attention on the resulting analogues of Theorems \ref{thm:simplified_ach} and \ref{thm:simplified_conv}.
\begin{thm} \label{thm:simplified_ach_p}
For any constants $\delta_1>0$, $\delta_2\in(0,1)$ and $\gamma>0$, and functions $\{\psi_{\ell}\}_{\ell=\dmax+1}^{k}$ ($\psi_\ell : \ZZ\times\RR\to\RR$), define the set
\begin{multline}
\Bc(\delta_1,\delta_2,\gamma) := \big\{ b_s \,:\, \text{\eqref{eq:psi_ach} and \eqref{eq:final_ach} hold for all }\\ (\sdif,\seq)\text{ with }|\sdif|\in\{\dmax+1,\dotsc,k\} \big\}. \label{eq:setB_p}
\end{multline}
Then we have
\begin{multline}
\pe(\dmax) \le \PP\big[ \beta_s \notin \Bc(\delta_1,\delta_2,\gamma)\big] \\ + \sum_{\ell=\dmax+1}^{k} {k \choose \ell}\psi_\ell(n,\delta_2) + P_0(\gamma) + 2\delta_1, \label{eq:nonasymp_ach_p}
\end{multline}
where $P_0$ is defined in \eqref{eq:P_0}.
\end{thm}
For the converse part, \eqref{eq:final_conv} is replaced by
\begin{equation}
n \ge \frac{\log {{p-k+|\sdif|} \choose |\sdif|} - \log\sum_{d=0}^{\dmax}{{p-k} \choose d}{|\sdif| \choose d} - \log\delta_1 }{ I_{\sdif,\seq}(b_s) }, \label{eq:final_conv_p}
\end{equation}
and we have the following analog of Theorem \ref{thm:simplified_conv}.
\begin{thm} \label{thm:simplified_conv_p}
For any set $\Lc \subseteq \{\dmax+1,\dotsc,k\}$, constants $\delta_1>0$ and $\delta_2\in(0,1)$, and functions $\{\psi'_{\ell}\}_{\ell\in\Lc}$ ($\psi'_\ell : \ZZ\times\RR\to\RR$), define the set
\begin{multline}
\Bc'(\delta_1,\delta_2) := \big\{ b_s \,:\, \text{\eqref{eq:psi_conv} and \eqref{eq:final_conv_p} hold for all } \\ (\sdif,\seq)\text{ with }|\sdif|\in\Lc \big\}. \label{eq:setB'_p}
\end{multline}
Then we have
\begin{equation}
\pe(\dmax) \ge \PP\big[ \beta_s \in \Bc'(\delta_1,\delta_2) \big]\Big(1 - \max_{\ell\in\Lc}\psi'_\ell(n,\delta_2)\Big) - \delta_1. \label{eq:nonasymp_conv_p}
\end{equation}
\end{thm}
The applications of Theorems \ref{thm:simplified_ach_p} and \ref{thm:simplified_conv_p} follow identical steps to Procedure \ref{pr:procedure}. However, it will be seen that the restriction $|\sdif| > \dmax$ can in fact considerably simplify these steps, since it removes the need to obtain concentration inequalities for smaller values of $|\sdif|$.
\subsection{Comparison to Fano's Inequality} \label{sec:FANO}
Most previous works on the information-theoretic limits of sparsity recovery have made use of Fano's inequality \cite[Sec.~2.11]{Cov01}. For this reason, we provide here a discussion on the relative merits of this approach and our approach. To this end, we consider the following bound, which can be obtained by combining the analysis of \cite{Ati12,Aks13} with our refined genie argument:
\begin{equation}
\pe \ge \sum_{b_s}P_{\beta_S}(b_s)\max\bigg\{0, 1 - \frac{nI_{\sdif(b_s),\seq(b_s)}(b_s) + 1}{\log{ p-k+|\sdif(b_s)| \choose |\sdif(b_s)| }}\bigg\}
\end{equation}
in the notation of Theorem \ref{thm:conv}. By analyzing this bound similarly to Section \ref{sec:APPLYING}, we obtain for any $\delta_2 > 0$ that
\begin{equation}
\pe \ge \delta_2\, \PP\big[\beta_s \in \Bcfano(\delta_2)\big] - \frac{1}{\log(p-k+1)}, \label{eq:pe_Fano}
\end{equation}
where
\begin{multline}
\Bcfano(\delta_2) := \bigg\{ b_s \,:\, n \le \frac{\log {{p - k + |\sdif|} \choose |\sdif|}}{I_{\sdif,\seq}(b_s)}(1-\delta_2) \\ \text{ for all }(\sdif,\seq)\text{ with }\sdif\ne\emptyset\bigg\}. \label{eq:pe_Fano2}
\end{multline}
A similar result for partial recovery can also be derived by incorporating the arguments from \cite{Ree13} and the present paper.
As discussed in the introduction, the key advantage of Theorem \ref{thm:simplified_conv} is that it provides a more precise characterization of how far the error probability is from zero, and in particular, the conditions under which $\pe \to 1$ (strong impossibility results). On the other hand, the bound on $\pe$ in \eqref{eq:pe_Fano} is always bounded away from one for fixed $\delta_2$, and becomes increasingly weak for small $\delta_2$.
The advantage of Fano's inequality is that it only requires the mutual information to be computed, whereas our approach also requires the application of a concentration inequality. This, in turn, typically requires the variance of the information density to be characterized, which is not always straightforward. However, as discussed following Procedure \ref{pr:procedure}, the main difficulty associated with these concentration inequalities is typically in finding one which is sufficiently powerful for the \emph{achievability} part. Thus, the added difficulty in the converse may not add to the overall difficulty in deriving matching achievability and converse bounds.
\section{Applications to Specific Models} \label{sec:EXAMPLES}
In this section, we present applications of Theorems \ref{thm:simplified_ach}--\ref{thm:simplified_conv_p} to the linear \cite{Wai09}, 1-bit \cite{Bou08}, and group testing \cite{Ati12} models, and to more general models with discrete observations \cite{Tan14}. Throughout the section, we make use of general concentration inequalities given in Appendix \ref{sec:CONCENTRATION}. We also make use of the following variance quantity:
\begin{equation}
V_{\sdif,\seq}(b_s) := \var\big[ \imath(X_{\sdif}; Y | X_{\seq}, b_s) \,\big|\, \beta_s = b_s \big]. \label{eq:Vb}
\end{equation}
\subsection{Linear Model with Discrete $\beta_s$} \label{sec:GAUSSIAN_DISC}
Here we consider the linear model, where each observation takes the form
\begin{equation}
Y = \langle X, \beta \rangle + Z, \label{eq:linear_model}
\end{equation}
where $Z \sim N(0,\sigma^2)$ for some $\sigma > 0$.
Without loss of generality, we consider the fixed support set $s=\{1,\dotsc,k\}$. Following the setup of \cite{Jin11}, we let $\beta_s$ be a uniformly random permutation of a fixed vector $(b_1,\dotsc,b_k)$, and we choose $P_X \sim N(0,1)$. Since both the measurement matrices and the noise are Gaussian, the mutual information in \eqref{eq:condMI} is given by \cite[Ch.~10]{Cov01}
\begin{equation}
I_{\sdif,\seq}(b_s) = \frac{1}{2}\log\Big(1 + \frac{1}{\sigma^2} \sum_{i \in \sdif}b_i^2\Big). \label{eq:I_linear}
\end{equation}
Throughout this subsection, we denote $\bmin := \min_{i} |b_i|$ and $\bmax := \max_{i} |b_i|$. We assume that $\sigma^2 = \Theta(1)$, and that $\bmin = \Theta(\bmax)$ and $0<\bmin = O(1)$; note that $\bmin = o(1)$ is allowed. The steps of Procedure \ref{pr:procedure} are as follows.
\subsubsection*{Step 1}
We trivially choose the typical set $\Tc_{\beta}$ to contain all vectors on the support of $P_{\beta_S}$.
\subsubsection*{Step 2}
We make use of the following concentration inequality based on Bernstein's inequality.
\begin{prop} \label{prop:conc_linear2}
Under the preceding setup for the linear model, we have for all $(\sdif,\seq)$ and $b_s$ that
\begin{multline}
\PP\bigg[ \big|\imath^n(\Xv_{\sdif}; \Yv | \Xv_{\seq}, b_s) - nI_{\sdif,\seq}(b_s) \big| \ge n\delta \,\Big|\, \beta_s = b_s \bigg] \\ \le 2\exp\bigg(-\frac{\delta^2 n}{2(4\alpha_{\sdif}^2+\delta\alpha_{\sdif})} \bigg),
\end{multline}
where
\begin{equation}
\alpha_{\sdif} := \frac{2\siglv(\sigma + \siglv)}{\sigma^2 + \siglv^2} \label{eq:alpha_linear}
\end{equation}
with $\siglv^2 := \sum_{i \in \sdif} b_i^2$.
\end{prop}
\begin{proof}
See Appendix \ref{sec:LINEAR_PROOFS}.
\end{proof}
Setting $\delta = \delta_2 I_{\sdif,\seq}(b_s)$, it follows that in \eqref{eq:psi_ach} and \eqref{eq:psi_conv} we can set
\begin{multline}
\psi_{\ell}(n,\delta_2) = \psi'_{\ell}(n,\delta_2) = 2\max_{(\sdif,\seq,b_s)\,:\,|\sdif|=\ell} \\ \exp\bigg(- \frac{ ( \delta_2 I_{\sdif,\seq}(b_s) )^2 n}{2(4\alpha_{\sdif}+\delta_2 I_{\sdif,\seq}(b_s))\alpha_{\sdif} } \bigg). \label{eq:linear_psi}
\end{multline}
\subsubsection*{Step 3}
In accordance with the first item of Remark \ref{rem:P_0}, we set $\gamma$ as in \eqref{eq:discrete_gamma} so that $P_0(\gamma) = 0$.
We focus on the conditions on $n$ under which the term $\sum_{\ell=1}^k{k \choose \ell}\psi_\ell(n,\delta_2)$ in \eqref{eq:nonasymp_ach} vanishes; the term containing $\psi'_{\ell}$ in \eqref{eq:nonasymp_conv} can be handled in a similar yet simpler fashion. By the assumptions $\sigma^2 = \Theta(1)$ and $\bmax = \Theta(\bmin)$, we readily obtain $I_{\sdif,\seq}(b_s) = \Theta(\log(1+\ell\bmin^2))$ and $\alpha_{\sdif}^2 = \Theta(\min\{1,\ell\bmin^2\})$ using \eqref{eq:I_linear} and \eqref{eq:alpha_linear}, where $\ell = |\sdif|$. Using these growth rates and upper bounding the summation in \eqref{eq:nonasymp_ach} by $k$ times the corresponding maximum, we see that $\sum_{\ell=1}^k{k \choose \ell}\psi_\ell(n,\delta_2) \to 0$ provided that the following holds for some sufficiently small constant $\zeta$ (depending on $\delta_2$):
\begin{multline}
\frac{ n \log^2(1+\ell\bmin^2)}{\min\{1,\ell\bmin^2\} + \log(1+\ell\bmin^2) \sqrt{\min\{1,\ell\bmin^2\}} }\zeta \\ - \ell \log\frac{k}{\ell} - \log k \to \infty \label{eq:lin_n_cond1}
\end{multline}
for all $\ell$. We now treat two cases separately:
\begin{itemize}
\item If $\ell\bmin^2 = o(1)$, the first term in \eqref{eq:lin_n_cond1} behaves as $\Theta(n \ell\bmin^2)$; by rearranging, we conclude that it suffices that $n\to\infty$ and $n = \Omega\big(\frac{\log k}{\bmin^2}\big)$ with a sufficiently large implied constant.
\item If $\ell\bmin^2 = \Omega(1)$, the first term in \eqref{eq:lin_n_cond1} behaves as $\Omega(n)$, and it thus suffices that $n\to\infty$ and $n=\Omega(k)$ with a sufficiently large implied constant.
\end{itemize}
Thus, the overall condition that we require is $n\to\infty$ and
\begin{equation}
n = \Omega\Big( \frac{\log k}{\bmin^2} \Big) \quad \text{ and } \quad n = \Omega(k), \label{eq:linear_n_cond}
\end{equation}
with sufficiently large implied constants. For the converse, the analogous condition to \eqref{eq:lin_n_cond1} contains only the first term on the left-hand side (the difference being due to the fact that the combinatorial term in \eqref{eq:nonasymp_ach} is not present in \eqref{eq:nonasymp_conv}), and a similar argument reveals that it suffices that $n = \omega\big( \frac{1}{\bmin^2} \big)$.
\subsubsection*{Step 4}
Combining the preceding steps and applying asymptotic simplifications, we obtain the following.
\begin{cor} \label{cor:linear_genK}
Under the preceding setup for the linear model with $\sigma^2 = \Theta(1)$, $\bmin=\Theta(\bmax)$, $\bmin^2 = O(1)$, $k=o(p)$, and $m_{\beta}$ distinct elements in $(b_1,\dotsc,b_k)$, we have $\pe \to 0$ as $p\to\infty$ provided that
\begin{equation}
n \ge \max_{\sdif\ne\emptyset} \frac{ \log {{p-k} \choose |\sdif|} }{ \frac{1}{2}\log\Big(1 + \frac{1}{\sigma^2} \sum_{i \in \sdif}b_i^2 \Big)} (1 + \eta), \label{eq:linear2_ach}
\end{equation}
under any one of the following additional conditions: (i) $k = \Theta(1)$; (ii) $k=o(\log p)$ and $m_{\beta} = \Theta(1)$; (iii) $k = O((\log p)^{\theta})$ for some $\theta>0$, and $m_{\beta} = 1$; (iv) $k=\Theta(p^{\theta})$ for some $\theta\in(0,1)$, $\bmin^2 = \Theta\big( \frac{\log k}{k} \big)$, and $m_{\beta} = 1$.
Conversely, without any additional conditions, we have $\pe \to 1$ as $p\to\infty$ whenever
\begin{equation}
n \le \max_{\sdif\ne\emptyset} \frac{ \log {p-k+|\sdif| \choose |\sdif|} }{ \frac{1}{2}\log\Big(1 + \frac{1}{\sigma^2} \sum_{i \in \sdif}b_i^2 \Big)} (1 - \eta) \label{eq:linear2_conv}
\end{equation}
for some $\eta > 0 $.
\end{cor}
\begin{proof}
The converse part follows from \eqref{eq:final_conv} with $\delta_1\to0$ sufficiently slowly. To check the condition $n = \omega\big( \frac{1}{\bmin^2} \big)$ stated following \eqref{eq:linear_n_cond}, we may assume without loss of generality that \eqref{eq:linear2_conv} holds with equality, since the decoder can always choose to ignore additional measurements. When equality holds, we observe that for the worst-case $\sdif$ with $\ell=1$, the denominator therein behaves as $O(\bmin^2)$ (since $\bmin^2 = O(1)$) and the numerator behaves as $\Theta(\log p)$, and hence, the condition $n = \omega\big( \frac{1}{\bmin^2} \big)$ is satisfied.
For the achievability part, we first use \eqref{eq:final_ach} to obtain
\begin{equation}
n \ge \max_{\sdif\ne\emptyset} \frac{ \log {{p-k} \choose |\sdif|} + 2\log\big(k{k\choose|\sdif|}\big) + k\log m_{\beta} }{ \frac{1}{2}\log\Big(1 + \frac{1}{\sigma^2} \sum_{i \in \sdif}b_i^2 \Big)} (1 + \eta), \label{eq:linear2_ach_0}
\end{equation}
where the final term in the numerator arises from \eqref{eq:discrete_gamma} since $P_{\beta_s}(b_s)$ is the same for all permutations of $(b_1,\dotsc,b_k)$, and is lower bounded by $m_{\beta}^{-k}$. Observe that the first term in the numerator behaves as $\Theta(|\sdif| \log p)$ for each of the cases in the corollary statement, and the second term behaves as $\Theta\big(\log k + |\sdif| \log \frac{k}{|\sdif|}\big)$.
In cases (i)--(iii), we have $\log k = o(\log p)$, and it immediately follows that the numerator in \eqref{eq:linear2_ach_0} is dominated by the first term, and hence, the others can be factored into $\eta$ in \eqref{eq:linear2_ach}. Moreover, in case (i), both conditions in \eqref{eq:linear_n_cond} are dominated by the objective in \eqref{eq:linear2_ach_0} with $\ell:=|\sdif|=1$, which behaves as $\Theta\big(\frac{\log p}{\bmin^2}\big)$. In cases (ii)--(iii), the first condition in \eqref{eq:linear_n_cond} is again dominated by the term in \eqref{eq:linear2_ach_0} with $\ell=1$. The second condition is dominated by the term with $\ell = k$, which behaves as $\Theta\big( \frac{k\log p}{ \log(1+k\bmin^2) } \big) = \Omega\big(k \frac{\log p}{\log k}\big)$.
In case (iv), the first term in the numerator of \eqref{eq:linear2_ach_0} may not be dominant for small $\ell := |\sdif|$, since $\log k = \Theta(\log p)$. However, by observing that the objective scales as $\Theta\big( \frac{\ell \log p}{ \log(1+\ell\bmin^2) } \big)$ and using the assumed scaling of $\bmin^2$, it is readily verified that the maximum can only be achieved with $\ell = \Theta(k)$. For any such maximizer, we have $\log{ p-k \choose \ell } = \Theta(k \log p)$, and hence, the second term in the numerator of \eqref{eq:linear2_ach_0} can be factored into $\eta$, as it behaves as $O(k)$. The two conditions in \eqref{eq:linear_n_cond} are identical under the given scaling of $\bmin^2$, and are dominated by the objective in \eqref{eq:linear2_ach} with $\ell=k$, which behaves as $\Theta\big( \frac{k \log k}{\log\log k} \big)$.
\end{proof}
In the case that $\bmin = \Theta(1)$, the thresholds given in Corollary \ref{cor:linear_genK} coincide with those given in the main results of \cite{Jin11}. Our framework has the advantage of handling the case that $\bmin = o(1)$, as well as providing the strong converse ($\pe \to 1$) instead of the weak converse ($\pe \not\to 0$). However, it should be noted that the achievability parts of \cite{Jin11} have the notable advantage of using a decoder that does not depend on the distribution of $\beta_s$.
On first glance, the bounds in \eqref{eq:linear2_ach}--\eqref{eq:linear2_conv} may appear to be difficult to evaluate, since the maximizations are over $2^{k}-1$ non-empty subsets $\sdif$. However, it is in fact only $k$ of them that need to be computed, since for any given $\ell = |\sdif|$ the maximizing $\sdif$ is the one with the smallest corresponding value of $\sum_{i \in \sdif}b_i^2$.
\subsubsection*{Comparison to the LASSO}
Conditions for the support recovery of the computationally tractable LASSO algorithm were given by Wainwright \cite{Wai09a}. Several comparisons to the information-theoretic limits were given in \cite{Wai09,Wai09a} in terms of scaling laws; here we complement these comparisons by briefly discussing the corresponding constant factors. For simplicity, we focus on the case that the non-zero entries are all equal to a common value $b_0 = \frac{c_{\beta}}{k}$ (for some constant $c_{\beta}$ representing the per-sample SNR) and $k$ is poly-logarithmic in $p$, corresponding to case (iii) of Corollary \ref{cor:linear_genK}.
The results of \cite{Wai09a} state that LASSO requires at least $(2k\log p)(1+o(1))$ measurements regardless of $c_{\beta}$, and that this bound is also achievable in the limit as $c_{\beta} \to \infty$. On the other hand, Corollary \ref{cor:linear_genK} reveals that for the optimal decoder, the coefficient to $k \log p$ can be arbitrarily small provided that $c_{\beta}$ is large enough. More precisely, applying some simple manipulations to \eqref{eq:linear2_ach}, we find that the coefficient to $k \log p$ is $\sup_{\alpha\in(0,1]} \frac{\alpha}{\frac{1}{2}\log(1+c_{\beta}\alpha)}$, where $\alpha$ represents the ratio $\frac{|\sdif|}{k}$. It is easy to verify that the maximum is achieved at $\alpha = 1$, yielding the constant $\frac{2}{\log(1+c_{\beta})}$. We conclude that the LASSO provably yields a suboptimal constant when $c_{\beta} > 1$, and fails to achieve the optimal logarithmic decay. However, it should be noted that our decoder requires knowledge of $k$ and $c_{\beta}$, whereas the LASSO does not (except possibly via their role in determining the regularization parameter).
\subsection{Linear Model with Gaussian $\beta_S$ and Partial Recovery} \label{sec:GAUSSIAN_CONT}
In this subsection, we consider the setup of Section \ref{sec:GAUSSIAN_DISC} with two changes: We let the distribution of $\beta_s$ be continuous rather than discrete, and we consider partial recovery instead of exact recovery. More specifically, we let $\beta_s$ be i.i.d.~on $N(0,\sigbeta^2)$ for some variance $\sigbeta^2$, and we consider the recovery condition in \eqref{eq:pe_partial} with
\begin{equation}
\dmax = \lfloor \alpha^* k \rfloor
\end{equation}
for some $\alpha^* \in (0,1)$ not varying with $p$. We again choose $P_X \sim N(0,1)$. We assume $\sigbeta^2 = \frac{c_{\beta}}{k}$ for some $c_{\beta} > 0$ not depending on $p$, corresponding to a fixed per-sample SNR.
We begin with the following auxiliary result.
\begin{prop} \label{prop:boundMI}
Under the preceding setup for the linear model, the quantities $I_0$ and $V_0$ defined in \eqref{eq:I_0}--\eqref{eq:V_0} satisfy
\begin{align}
I_0 &\le \frac{k}{2} \log\Big( 1 + \frac{n\sigbeta^2}{\sigma^2} \Big) \label{eq:bmi_I0} \\
V_0 &\le 2n. \label{eq:bmi_V0}
\end{align}
\end{prop}
\begin{proof}
See Appendix \ref{sec:LINEAR_PROOFS}.
\end{proof}
We now proceed with the steps of Procedure \ref{pr:procedure} (with the suitable changes from exact recovery to partial recovery, \emph{cf.}~Section \ref{sec:PARTIAL}).
\subsubsection*{Step 1}
Our choice of the typical set $\Tc_{\beta}$ is based on the following proposition characterizing the behavior of the $\lfloor \alpha k \rfloor$ entries of $\beta_s$ having the smallest magnitude for fixed $\alpha$. We define the random variable $\beta'_s$ to be the permutation of $\beta_s$ whose entries are listed in increasing order of magnitude.
\begin{prop} \label{prop:empiricalB}
For any $\alpha\in(0,1]$, we have
\begin{equation}
\lim_{k\to\infty} \frac{1}{k\sigbeta^2}\sum_{i=1}^{\lfloor \alpha k \rfloor} (\beta'_s)_i^2 = g(\alpha) \label{eq:empiricalB}
\end{equation}
with probability one, where
\begin{equation}
g(\alpha) := \int_{0}^{\infty} \big[ \alpha - F_{\chi^2}(u)\big]^+ \,du, \label{eq:g_alpha}
\end{equation}
and $F_{\chi^2}$ is the cumulative distribution function of a $\chi^2$ random variable with one degree of freedom.
\end{prop}
\begin{proof}
Letting $\hat{F}_k$ be the empirical distribution of the values $\big\{\frac{1}{\sigbeta^2}\beta_i^2\big\}_{i=1}^k$, we have from the Glivenko-Cantelli theorem \cite[Thm.~19.1]{Van00} that $\sup_{u} |\hat{F}_k(u) - F_{\chi^2}(u)| \to 0$ almost surely. This immediately implies that the sum of the $\lfloor \alpha k \rfloor$ smallest values in $\big\{\frac{1}{\sigbeta^2}\beta_i^2\big\}$, normalized by the number of values $k$, converges almost surely to the integral of $F_{\chi^2}^{-1}(u)$ from $0$ to $\alpha$. It is easily verified graphically that this integral can equivalently be written as \eqref{eq:g_alpha}.
\end{proof}
Based on this result and its proof, we set $\Tc_{\beta}$ to be the set of vectors $b_s$ such that $\sup_{u} |\hat{F}_k(u) - F_{\chi^2}(u)| \le \epsilon$, where $\epsilon$ is chosen to decay sufficiently slowly so that $\PP[\beta_s \in \Tc_{\beta}] \to 1$. Thus, within the typical set, the empirical distribution of the non-zero entries closely follows a $\chi^2$ random variable.
An important consequence of this choice of typical set regards the behavior of the mutual information in \eqref{eq:I_linear}. For a fixed set size $|\sdif|$, the partition $(\sdif,\seq)$ minimizing this mutual information is the one with the smallest value of $\sum_{i\in\sdif}b_i^2$. Within the typical set, we immediately obtain from Proposition \ref{prop:empiricalB} that the corresponding mutual information behaves as follows when $|\sdif| = \lfloor \alpha k \rfloor$:
\begin{equation}
I_{\sdif,\seq}(b_s) \to \frac{1}{2} \log\Big(1 + \frac{c_{\beta}}{\sigma^2} g(\alpha) \Big), \label{eq:linear_I_asymp}
\end{equation}
where we recall that $c_{\beta} = k\sigbeta^2$ is a constant.
\subsubsection*{Step 2}
We again make use of Proposition \ref{prop:conc_linear2} and its subsequent expression for $\psi_{\ell}$ and $\psi'_{\ell}$ in \eqref{eq:linear_psi}.
\subsubsection*{Step 3}
We choose $\gamma = I_0 + \sqrt{\frac{V_0}{\delta_0}}$ as in \eqref{eq:gamma_cheby} for some $\delta_0 > 0$, thus ensuring that $P_0(\gamma) \le \delta_0$.
For the terms in Theorems \ref{thm:simplified_ach_p}--\ref{thm:simplified_conv_p} containing $\psi_{\ell}$ and $\psi'_{\ell}$, we first note that since we are considering partial recovery, we may focus on values of $\ell=|\sdif|$ greater than $\alpha^{*} k$. By our choice of $\Tc_{\beta}$, we may also focus on realizations $b_s$ of $\beta_s$ satisfying \eqref{eq:empiricalB}. For such realizations, we have for all $\sdif$ with $|\sdif|=\ell=\Theta(k)$ that $\sum_{i\in\sdif} b_i^2 = \Omega(1)$, which implies that $\alpha_{\sdif}^2 = \Theta(1)$ in \eqref{eq:alpha_linear} and $I_{\sdif,\seq}(b_s) = \Omega(1)$ in \eqref{eq:I_linear}. The analogous condition to \eqref{eq:lin_n_cond1} thus simplifies to $n I' \gg k$ for some $I' = \Omega(1)$, giving the following condition under which the second term in \eqref{eq:nonasymp_ach} vanishes:
\begin{equation}
n = \Omega(k), \label{eq:lin_cond_mod}
\end{equation}
with a sufficiently large implied constant. For the converse part, it suffices to have the weaker condition $n=\omega(1)$.
\subsubsection*{Step 4}
Combining the above steps, we get the following.
\begin{cor} \label{cor:linear_partial}
Under the preceding setup for the linear model with $k\to\infty$, $k=o(p)$, $\sigbeta^2 = \frac{c_{\beta}}{k}$ for some $c_{\beta} > 0$, and $\dmax = \lfloor \alpha^* k \rfloor$ for some $\alpha^*\in(0,1)$, we have $\pe(\dmax) \to 0$ as $p\to\infty$ provided that
\begin{equation}
n \ge \max_{\alpha\in[\alpha^*,1]} \frac{ \alpha k \log \frac{p}{k} }{ \frac{1}{2}\log\Big(1 + \frac{c_{\beta}}{\sigma^2} g(\alpha) \Big)} (1 + \eta) \label{eq:linear3_ach}
\end{equation}
for some $\eta > 0$, where $g(\cdot)$ is defined in \eqref{eq:g_alpha}. Conversely, $\pe(\dmax) \to 1$ as $p\to\infty$ whenever
\begin{equation}
n \le \max_{\alpha\in[\alpha^*,1]} \frac{ (\alpha - \alpha^*) k \log \frac{p}{k} }{ \frac{1}{2}\log\Big(1 + \frac{c_{\beta}}{\sigma^2} g(\alpha) \Big)} (1 - \eta) \label{eq:linear3_conv}
\end{equation}
for some $\eta > 0 $.
\end{cor}
\begin{proof}
The condition in \eqref{eq:linear3_ach} is obtained using \eqref{eq:final_ach} and \eqref{eq:linear_I_asymp}. By the assumption $k=o(p)$, the numerator in \eqref{eq:linear3_ach} coincides with $\log {{p-k} \choose \lfloor \alpha k \rfloor}$ up to remainder terms in Stirling's approximation that can be factored into $\eta$. The factor $\log\big(\frac{k^2}{\delta_1^2}{k \choose |\sdif|}^2\big)$ in \eqref{eq:final_ach} has been factored into $\eta$; this is valid when $\delta_1 \to 0$ sufficiently slowly due to the fact that $\log\big(k {k \choose |\sdif|}\big) = O(k)$, whereas (again using the assumption $k=o(p)$) the numerator in \eqref{eq:linear3_ach} behaves as $\omega(k)$. We claim that the factor $\gamma = I_0 + \sqrt{\frac{V_0}{\delta_0}}$ resulting from \eqref{eq:gamma_cheby} can also be factored into $\eta$ for some vanishing sequence of parameters $\delta_0$ indexed by $p$. To see this, we consider without loss of generality the ``worst-case'' setting in which \eqref{eq:linear3_ach} holds with equality. We readily obtain $n = \Theta( k \log \frac{p}{k} )$, which in turn implies from Proposition \ref{prop:boundMI} that $I_0 = O\big(k \log(1 + k\sigbeta^2 \log\frac{p}{k} )\big) = O(k\log\log\frac{p}{k}) $ and $\sqrt{V_0} = O(\sqrt{k \log \frac{p}{k}})$. Thus, $I_0 + \sqrt{\frac{V_0}{\delta_0}}$ is dominated by the numerator of \eqref{eq:linear3_ach} if $\delta_0$ is chosen to decay as (for example) $\Theta\big( \frac{1}{\log k} \big)$. The fact that $n = \Theta( k \log \frac{p}{k} )$ also implies \eqref{eq:lin_cond_mod}.
The converse bound in \eqref{eq:linear3_ach} is obtained similarly using \eqref{eq:final_ach}, except for the term $\alpha^*$ in the numerator. To see how this arises, we consider an arbitrary value of $\alpha\in(\alpha^*,1]$ and set $\ell = \lfloor \alpha k\rfloor$; the case $\alpha=\alpha^*$ follows by continuity. The term $\log {{p-k+\ell} \choose \ell}$ is handled in the same way as the term $\log {{p-k} \choose \ell}$ above, so we focus on the term $\log\sum_{d=0}^{\dmax}{{p-k} \choose d}{\ell \choose d}$. This is upper bounded by $\max_{d=0,\dotsc,\dmax}\log\big((1+\dmax){{p-k} \choose d}{k \choose d}\big)$. Similarly to the achievability part, we can factor $\log\big((1+\dmax)\log {k \choose d}\big)$ into $\eta$, so we are left with $\log {{p-k} \choose \dmax}$. Approximating this using Stirling's approximation as before, and recalling that $\dmax = \lfloor \alpha^* k \rfloor$, we obtain the desired term $\alpha^{*}k\log\frac{p}{k}$.
\end{proof}
While the achievability and converse bounds in Corollary \ref{cor:linear_partial} do not have the same constants, the two are similar, and always have the same scaling laws. In the limit as $c_{\beta} \to \infty$, we have $\frac{1}{2}\log\big(1 + \frac{c_{\beta}}{\sigma^2} g(\alpha) \big) = \frac{1}{2}(\log c_{\beta})(1+o(1))$; in this case, the maxima in \eqref{eq:linear3_ach}--\eqref{eq:linear3_conv} are both achieved with $\alpha\to1$, and hence, the two bounds coincide to within a multiplicative factor of $\frac{1}{1-\alpha^*}$.
Corollary \ref{cor:linear_partial} is related to the setting studied by Reeves and Gastpar \cite{Ree12,Ree13}, but considers $k = o(p)$ instead of $k = \Theta(p)$. Despite this difference, it is instructive to compare the bounds upon letting the implied constant in the $\Theta(p)$ scaling tend to zero. A careful comparison reveals that the converse bounds coincide in this limit, whereas our achievability bound is slightly better, in that the analogous bound in \cite{Ree12} multiplies $\frac{c_{\beta}}{\sigma^2} g(\alpha)$ by $(\sqrt{2}-1)^2 \approx 0.17$; see \cite[Eq.~(21)]{Ree12} and \cite[Eq.~(25)]{Ree13}.
In Section \ref{sec:NUMERICAL_PARTIAL}, we present some numerical results for this setting.
\subsection{1-bit Model with Discrete $\beta_S$} \label{sec:1BIT_DISC}
We now turn to the quantized counterpart of \eqref{eq:linear_model}:
\begin{equation}
Y = \sign\big(\langle X_S, \beta_S \rangle + Z\big).
\end{equation}
As in Section \ref{sec:GAUSSIAN_DISC}, we fix $s=\{1,\dotsc,k\}$ and let $\beta_s$ be a uniformly random permutation of a fixed vector $(b_1,\dotsc,b_k)$, and we set $P_{X} \sim N(0,1)$. We again write the minimum and maximum absolute values of $\{b_i\}_{i=1}^k$ as $\bmin$ and $\bmax$.
The following proposition gives the required characterizations on the mutual information terms and the corresponding variance terms. Recall the binary entropy function $H_2(\cdot)$ and the Q-function $Q(\cdot)$ defined in Section \ref{sec:NOTATION}.
\begin{prop} \label{prop:1bit_moments}
Under the preceding setup for the 1-bit model, we have the following:
(i) The mutual information $I_{\sdif,\seq}(b_s)$ is given by
\begin{multline}
I_{\sdif,\seq}(b_s) = \EE\bigg[ H_2\bigg( Q\bigg( W \sqrt{ \frac{\sum_{i\in\seq}b_i^2}{\sigma^2 + \sum_{i\in\sdif}b_i^2} } \bigg) \bigg) \\ - H_2\bigg( Q\bigg( W \sqrt{\frac{1}{\sigma^2}\sum_{i \in s} b_i^2} \bigg) \bigg) \bigg], \label{eq:1bit_Ilv}
\end{multline}
where $W \sim N(0,1)$.
(ii) If $k=\Theta(1)$, $\sigma^2=\Theta(1)$, $\bmin = \Theta(\bmax)$, and $\bmin^2 = o(1)$, then
\begin{equation}
I_{\sdif,\seq}(b_s) = \bigg(\frac{1}{\pi\sigma^2} \sum_{i\in\sdif}b_i^2\bigg)\big(1+o(1)\big). \label{eq:1bit_Ilv_asymp}
\end{equation}
(iii) If $k=\Theta(p)$, $\sigma^2 = \Theta(1)$, and the entries of $b_s$ all equal a common value $b_0$ such that $b_0^2 = \Theta\big(\frac{\log p}{p}\big)$, then the mutual information quantities $I_{\sdif,\seq}(b_s)$ with $|\sdif|=1$ all equal a common value $I_1$ satisfying
\begin{align}
I_1 &= \frac{1}{2} \frac{\frac{b_0^2}{\sigma^2}}{\sqrt{2\pi k\frac{b_0^2}{\sigma^2}}} \EE\Big[ W \log\frac{1-Q(W)}{Q(W)} \Big] (1+o(1)) \label{eq:1bit_I1_exact} \\
&= \Theta\bigg( \frac{\sqrt{\log p}}{p} \bigg), \label{eq:1bit_I1}
\end{align}
where $W \sim N(0,1)$.
(iv) The variance $V_{\sdif,\seq}(b_s)$ defined in \eqref{eq:Vb} satisfies
\begin{multline}
V_{\sdif,\seq}(b_s) \le c_0 \Bigg( \frac{1}{\sigma^2}\sum_{i\in\sdif}b_i^2 +
\bigg(\frac{1}{\sigma^2}\sum_{i\in\sdif}b_i^2\bigg)^2 \\ +
\min\bigg\{1,\bigg(\frac{1}{\sigma^2}\sum_{i\in\sdif}b_i^2\bigg)^2\bigg\}\frac{1}{\sigma^2}\sum_{i\in\seq}b_i^2
\Bigg) \label{eq:1bit_Vlv}
\end{multline}
for some universal constant $c_0$.
\end{prop}
\begin{proof}
See Appendix \ref{sec:1BIT_PROOFS}.
\end{proof}
Below we present two corollaries corresponding to different scalings of $k$ and the SNR, namely, those given in parts (ii) and (iii) of Proposition \ref{prop:1bit_moments}. We proceed by simultaneously presenting the steps of Procedure \ref{pr:procedure} for both settings.
\subsubsection*{Step 1}
As in Section \ref{sec:GAUSSIAN_DISC}, we choose the trivial typical set $\Tc_{\beta}$ containing all vectors on the support of $P_{\beta_S}$.
\subsubsection*{Step 2}
We make use of Chebyshev's inequality in Proposition \ref{prop:chebyshev} in Appendix \ref{sec:CONCENTRATION}. Choosing $\delta = \delta_2 I_{\sdif,\seq}(b_s)$ in \eqref{eq:conc_linear1}, it follows that we may set
\begin{multline}
\psi_{\ell}(n,\delta_2) = \psi'_{\ell}(n,\delta_2) \\ = \max_{(\sdif,\seq,b_s)\,:\,|\sdif|=\ell} \frac{V_{\sdif,\seq}(b_s)}{n\delta_2^2 I_{\sdif,\seq}(b_s)^2}. \label{eq:1bit_cheby}
\end{multline}
\subsubsection*{Step 3}
We again choose $\gamma$ as in \eqref{eq:discrete_gamma} so that $P_0(\gamma) = 0$. Consider the setting described in part (ii) of Proposition \ref{prop:1bit_moments}. Under the scalings therein, \eqref{eq:1bit_Ilv_asymp} and \eqref{eq:1bit_Vlv} both behave as $\Theta( \bmin^2 )$. Hence, and using \eqref{eq:1bit_cheby} and the fact that $k=\Theta(1)$, the second term in \eqref{eq:nonasymp_ach} vanishes provided that
\begin{equation}
n=\omega\Big( \frac{1}{\bmin^2} \Big). \label{eq:1bit_n_cond_a}
\end{equation}
The setting described in part (iii) of Proposition \ref{prop:1bit_moments} is handled similarly. We set $\Lc=\{1\}$ in Theorem \ref{thm:simplified_conv}, thus focusing only on $\ell := |\sdif| = 1$. Denoting the corresponding variance $V_{\sdif,\seq}(b_s)$ by $V_1$, it follows by substituting the scalings of $k$, $\sigma^2$ and $b_0^2$ into \eqref{eq:1bit_Vlv} that $V_1 = O\big( \frac{\log p}{p} \big)$. It thus follows from \eqref{eq:1bit_I1} and \eqref{eq:1bit_cheby} that $\psi'_{1}(n,\delta_2)$ vanishes provided that
\begin{equation}
n=\omega(p). \label{eq:1bit_n_cond_b}
\end{equation}
\subsubsection*{Step 4}
Combining the above steps and applying asymptotic simplifications, we obtain the following corollaries.
\begin{cor} \label{cor:1bit_fixedK}
Under the preceding setup for the 1-bit model with $k=\Theta(1)$, $\sigma^2 = \Theta(1)$, $\bmin = \Theta(\bmax)$, and $\bmin = o(1)$, we have $\pe \to 0$ as $p\to\infty$ provided that
\begin{equation}
n \ge \max_{\sdif\ne\emptyset} \frac{ |\sdif| \log p }{ \frac{1}{\pi\sigma^2}\sum_{i \in \sdif}b_i^2} (1 + \eta) \label{eq:1bit_n_cond1}
\end{equation}
for some $\eta > 0$. Conversely, $\pe \to 1$ as $p\to\infty$ whenever
\begin{equation}
n \le \max_{\sdif\ne\emptyset} \frac{ |\sdif| \log p }{ \frac{1}{\pi\sigma^2}\sum_{i \in \sdif}b_i^2} (1 - \eta) \label{eq:1bit_n_cond2}
\end{equation}
for some $\eta > 0$.
\end{cor}
\begin{proof}
We obtain \eqref{eq:1bit_n_cond1} and \eqref{eq:1bit_n_cond2} from \eqref{eq:final_ach} and \eqref{eq:final_conv} respectively. The denominators are obtained directly from part (ii) of Proposition \ref{prop:1bit_moments}, and the numerators follow from the identity $\log{p \choose |\sdif|} = (|\sdif| \log p)(1+o(1))$, which holds whenever $k=\Theta(1)$ and hence $|\sdif|=\Theta(1)$. By the assumption $k=\Theta(1)$, the remaining terms in \eqref{eq:final_ach} (including the choice of $\gamma$ in \eqref{eq:discrete_gamma}) can be factored into $\eta$. The condition in \eqref{eq:1bit_n_cond_a} is implied by \eqref{eq:1bit_n_cond1} (or by \eqref{eq:1bit_n_cond2} when equality holds) by the same argument as Corollary \ref{cor:linear_genK}.
\end{proof}
\begin{cor} \label{cor:1bit_genK}
Under the preceding setup for the 1-bit model with $k=\Theta(p)$, $\sigma^2 = \Theta(1)$, and the entries of $\beta_s$ deterministically equaling a common value $b_0$ such that $b_0^2 = \Theta\big(\frac{\log p}{p}\big)$, we have $\pe \to 1$ provided that
\begin{align}
n &\le \frac{\log p}{ \frac{1}{2} \frac{\frac{b_0^2}{\sigma^2}}{\sqrt{2\pi k\frac{b_0^2}{\sigma^2}}} \EE\Big[ W \log\frac{1-Q(W)}{Q(W)} \Big] } (1-\eta) \label{eq:1bit_n_cond3} \\
&= \Theta\big( p \sqrt{\log p} \big)
\end{align}
for some $\eta\in(0,1)$, where $W \sim N(0,1)$.
\end{cor}
\begin{proof}
The condition in \eqref{eq:1bit_n_cond3} follows using \eqref{eq:final_conv} with $|\sdif| = 1$; the numerator behaves as $(\log p) (1+o(1))$, and the denominator behaves according to \eqref{eq:1bit_I1_exact}. The additional condition in \eqref{eq:1bit_n_cond_b} is satisfied when \eqref{eq:1bit_n_cond3} holds with equality.
\end{proof}
In the same way as \eqref{eq:linear2_ach}--\eqref{eq:linear2_conv}, one can compute \eqref{eq:1bit_n_cond1}--\eqref{eq:1bit_n_cond2} without evaluating all $2^k - 1$ objective values; for a given value of $|\sdif|$, the maximum is achieved by the set $\sdif$ with the smallest value of $\sum_{i \in \sdif}b_i^2$.
The asymptotic identities used in the proof of Corollary \ref{cor:1bit_fixedK} can directly be applied to \eqref{eq:linear2_ach}--\eqref{eq:linear2_conv} with $k=\Theta(1)$ and $\bmin = o(1)$, and the resulting expressions are precisely those in \eqref{eq:1bit_n_cond1}--\eqref{eq:1bit_n_cond2} with $\frac{1}{\pi}$ replaced by $\frac{1}{2}$. Thus, this is a case where there is only a minor loss in the performance due to the quantization; the corresponding asymptotic number of measurements only increases by a factor of $\frac{\pi}{2} \approx 1.57$.
In contrast, Corollary \ref{thm:simplified_conv} describes a setting where the linear model and its 1-bit counterpart lead to significantly different requirements on the number of measurements. Under the scaling described therein, the necessary and sufficient number of measurements for the linear model behaves as $\Theta(p)$ \cite[Table I]{Rad11}. Thus, the 1-bit quantization increases the required number of measurements from linear to super-linear in the ambient dimension.
\subsection{1-bit Model with Gaussian $\beta_S$ and Partial Recovery} \label{sec:1BIT_CONT}
We now consider the 1-bit counterpart of the setting studied in Section \ref{sec:GAUSSIAN_CONT}, where $\beta_s$ is i.i.d.~on $N(0,\sigbeta^2)$ for some $\sigbeta^2 = \frac{c_{\beta}}{k}$, and we seek partial recovery as in \eqref{eq:pe_partial} with $\dmax = \lfloor \alpha^* k \rfloor$. We make use of the following.
\begin{prop} \label{prop:1bit_boundMI}
Under the preceding setup for the 1-bit model, the quantity $I_{0,+}$ in \eqref{eq:I_0+} satisfies
\begin{equation}
I_{0,+} \le \frac{k}{2} \log\Big( 1 + \frac{n\sigbeta^2}{\sigma^2} \Big) + \sqrt{ k\log\Big( 1 + \frac{n\sigbeta^2}{\sigma^2} \Big) }. \label{eq:1bit_bmi_I0+}
\end{equation}
for some universal constant $c'_0$.
\end{prop}
\begin{proof}
By the data processing inequality, $I_0$ must satisfy \eqref{eq:bmi_I0} even in the 1-bit setting. We immediately obtain \eqref{eq:1bit_bmi_I0+} from the identity $I_{0,+} \le I_0 + \sqrt{2I_0}$ given in \cite{Bar00}.
\end{proof}
We now turn to the steps for providing a counterpart to Corollary \ref{cor:linear_partial}. We define the function
\begin{multline}
\Psi(\alpha,c_{\beta},\sigma) := \EE\bigg[ H_2\bigg( Q\bigg( W \sqrt{\frac{c_{\beta} (1-g(\alpha)) }{ \sigma^2 + c_{\beta} g(\alpha)}} \bigg) \bigg) \\ - H_2\bigg( Q\bigg( W \sqrt{\frac{c_{\beta}}{\sigma^2}} \bigg) \bigg) \bigg], \label{eq:1bit_Psi}
\end{multline}
where $W \sim N(0,1)$, and $g(\alpha)$ is defined in \eqref{eq:g_alpha}.
\subsubsection*{Step 1}
We choose the same typical set $\Tc_{\beta}$ as that in Section \ref{sec:GAUSSIAN_CONT}, thus ensuring that \eqref{eq:empiricalB} holds for all sequences of typical vectors. It follows that $\sum_{i\in\sdif}b_i^2 \to c_{\beta}g(\alpha)$ and $\sum_{i\in\seq}b_i^2 \to c_{\beta}(1-g(\alpha))$ for the pair $(\sdif,\seq)$ with corresponding sizes $(\ell,k-\ell)$ ($\ell = \lfloor \alpha k \rfloor$) such that $\sum_{i\in\sdif}b_i^2$ is minimized. We observe from \eqref{eq:1bit_Ilv} that minimizing $\sum_{i\in\sdif}b_i^2$ also amounts to minimizing $I_{\sdif,\seq}(b_s)$ for a fixed value of $|\sdif|$, as was the case for the linear model. If $\frac{|\sdif|}{k}$ converges to a given constant $\alpha$, then the corresponding mutual information converges as follows, in accordance with \eqref{eq:1bit_Ilv} and \eqref{eq:1bit_Psi}:
\begin{equation}
I_{\sdif,\seq}(b_s) \to \Psi(\alpha,c_{\beta},\sigma). \label{eq:1bit_I_asymp}
\end{equation}
\subsubsection*{Step 2}
We make use of the general concentration inequality given in Proposition \ref{prop:gen_discrete} in Appendix \ref{sec:CONCENTRATION}; setting $\delta = \delta_2 I_{\sdif,\seq}(b_s)$ in \eqref{eq:conc_gen_disc} in Appendix \ref{sec:CONCENTRATION} gives
\begin{align}
& \psi_{\ell}(n,\delta_2) - \psi'_{\ell}(n,\delta_2) \nonumber \\
& \hspace*{-0.5ex}= \max_{(\sdif,\seq,b_s)\,:\,|\sdif|=\ell} 2\exp\bigg(- \frac{(\delta_2 I_{\sdif,\seq}(b_s))^2 n}{2(8|\Yc|+2\delta_2 I_{\sdif,\seq}(b_s))} \bigg). \label{eq:1bit_gen_disc}
\end{align}
\subsubsection*{Step 3}
We choose $\gamma = \frac{I_{0,+}}{\delta_0}$ as in \eqref{eq:gamma_markov}, ensuring that $P_0(\gamma) \le \delta_0$. The other remainder terms are controlled in the same way as Section \ref{sec:GAUSSIAN_CONT}. We again use the fact that the typical realizations $b_s$ of $\beta_s$ satisfy \eqref{eq:empiricalB}, and yield $\sum_{i\in\sdif} b_i^2 = \Omega(1)$, and hence $\alpha_{\sdif}^2 = \Theta(1)$ in \eqref{eq:alpha_linear}. We also have $I_{\sdif,\seq}(b_s) = \Theta(1)$ in \eqref{eq:1bit_Ilv}; this is seen by noting that the smallest mutual information for a fixed $|\sdif| = \lfloor \alpha k \rfloor$ satisfies \eqref{eq:1bit_I_asymp}, and the mutual information upper bounded by $\log2$ since the observations are binary. It follows that the exponent in \eqref{eq:1bit_gen_disc} behaves as $\Theta(n)$; hence, following the arguments in Section \ref{sec:GAUSSIAN_CONT}, we conclude that the second term in \eqref{eq:nonasymp_ach} vanishes provided that
\begin{equation}
n = \Omega(k) \label{eq:1bit_n_cond_c}
\end{equation}
with a sufficiently large implied constant. Once again, for the converse part, one can analogously show that the weaker condition $n=\omega(1)$ suffices.
\subsubsection*{Step 4}
Combining the above steps, we get the following.
\begin{cor} \label{cor:1bit_partial}
Under the preceding setup for the 1-bit model with $k\to\infty$ and $k= o(p)$, $\sigma^2 = \Theta(1)$, $\sigbeta^2 = \frac{c_{\beta}}{k}$ for some $c_{\beta} > 0$, and $\dmax = \lfloor \alpha^* k \rfloor$ for some $\alpha^*\in(0,1)$, we have $\pe(\dmax) \to 0$ as $p\to\infty$ provided that
\begin{equation}
n \ge \max_{\alpha\in[\alpha^*,1]} \frac{ \alpha k \log \frac{p}{k} }{ \Psi(\alpha,c_{\beta},\sigma) } (1 + \eta) \label{eq:1bit_ach2}
\end{equation}
for some $\eta > 0$, where $\Psi$ is defined in \eqref{eq:1bit_Psi}. Conversely, $\pe(\dmax) \to 1$ as $p\to\infty$ whenever
\begin{equation}
n \le \max_{\alpha\in[\alpha^*,1]} \frac{ (\alpha - \alpha^*) k \log \frac{p}{k} }{ \Psi(\alpha,c_{\beta},\sigma) } (1 - \eta) \label{eq:1bit_conv2}
\end{equation}
for some $\eta > 0 $.
\end{cor}
\begin{proof}
As usual, we begin with the conditions in \eqref{eq:final_ach} and \eqref{eq:final_conv}. The denominators in \eqref{eq:1bit_ach2}--\eqref{eq:1bit_conv2} follow directly by applying \eqref{eq:1bit_I_asymp}. Moreover, the terms $\alpha k\log\frac{p}{k}$ and $(\alpha - \alpha^*) k\log\frac{p}{k}$ in the numerators are obtained in an identical fashion to Corollary \ref{cor:linear_partial} once we show that there exists a vanishing sequence of constants $\delta_0$, indexed by $p$, such that the remainder term $\gamma = \frac{I_{0,+}}{\delta_0}$ resulting from \eqref{eq:gamma_markov} can be factored into $\eta$. To see this, we note that the right-hand side of \eqref{eq:1bit_ach2} behaves as $\Theta(k \log p)$, whereas from Proposition \ref{prop:1bit_boundMI} (with the scalings $n=\Theta(k\log p)$ and $\sigbeta^2 = \Theta\big(\frac{1}{k}\big)$), $I_{0,+}$ behaves as $O\big(k\log\log p\big)$. We may thus set $\delta_0$ to be (for example) $\frac{\log\log p}{\sqrt{\log p}}$. Finally, we observe that \eqref{eq:1bit_n_cond_c} holds whenever \eqref{eq:1bit_ach2} holds, and similarly for the converse part.
\end{proof}
The main difference in \eqref{eq:1bit_ach2}--\eqref{eq:1bit_conv2} compared to the linear counterparts in \eqref{eq:linear3_ach}--\eqref{eq:linear3_conv} is the behavior in the limit as $c_{\beta} := k\sigbeta^2 \to \infty$. As stated following Corollary \ref{cor:linear_partial}, the denominator in the linear setting behaves as $(\log c_{\beta}) (1+o(1))$, thus tending towards infinity. In contrast, for the 1-bit setting, we have $\Psi \le \log 2$ due to the fact that $H_2(\cdot) \in [0,\log 2]$, and thus the denominator cannot grow unbounded. These observations are consistent with Corollary \ref{cor:1bit_genK}, which shows that 1-bit CS can require significantly more measurements compared to the linear setting when the signal-to-noise ratio (SNR) is sufficiently high.
\subsection{Numerical Evaluations for Partial Recovery} \label{sec:NUMERICAL_PARTIAL}
In this subsection, we present numerical calculations for the settings considered in Sections \ref{sec:GAUSSIAN_CONT} and \ref{sec:1BIT_CONT}. We set $\alpha^* = 0.1$, $\sigma^2 = 1$, and $k = o(p)$. We consider values of $\sigbeta^2$ of the form $\sigbeta^2 = \frac{c_{\beta}}{k}$ for fixed $c_{\beta}$. Similarly to \cite{Ree12,Ree13}, we present our results in terms of
\begin{equation}
\SNRdB := 10\log\frac{k\sigbeta^2}{\sigma^2} = 10\log c_{\beta}, \label{eq:SNRdB}
\end{equation}
which represents the per-sample SNR in dB.
\begin{figure}
\begin{centering}
\includegraphics[width=0.95\columnwidth]{IS_partial.pdf}
\par
\end{centering}
\caption{ Asymptotic thresholds on the number of measurements required for partial support recovery for the linear and 1-bit models, with $\alpha^* = 0.1$. The number of measurements is normalized by $k\log\frac{p}{k}$, and $\SNRdB$ is defined in \eqref{eq:SNRdB}. \label{fig:partial_recovery}}
\end{figure}
Figure \ref{fig:partial_recovery} plots the asymptotic thresholds on the number of measurements from Corollaries \ref{cor:linear_partial} and \ref{cor:1bit_partial}. For both the linear and 1-bit settings, there is a close correspondence between the necessary and sufficient number of measurements. The bounds for the two models nearly coincide at low SNR, which is consistent with Corollary \ref{cor:1bit_fixedK}.
The behavior of the bounds at high SNRs is also consistent with our previous discussions. In the linear setting, the ratio between the bounds narrows to approximately $1.11$ as the SNR grows large, which coincides with the value $\frac{1}{1-\alpha^*}$ given in the discussion following Corollary \ref{cor:linear_partial}. Moreover, as discussed following Corollary \ref{cor:1bit_partial}, the number of measurements steadily decreases for increasing SNRs for the linear model, while saturating at an asymptotic limit for the 1-bit model.
\subsection{Group Testing} \label{sec:GROUP_TESTING}
\subsubsection{Noiseless Case with Exact Recovery}
Here we consider the noiseless group testing problem, where each observation is deterministically generated according to
\begin{equation}
Y = \openone\Big\{ \bigcup_{i \in S} \{X_i = 1\} \Big\}. \label{eq:gt_noiseless}
\end{equation}
We consider Bernoulli measurement matrices with $P_X(1) = 1 - P_X(0) = \frac{\nu}{k}$, where $\nu$ is a constant not depending on $p$. Here there is no latent variable $\beta_s$, which can equivalently be thought of as corresponding to $\beta_s$ equaling the vector of ones deterministically. This implies that $I_{\sdif,\seq}(b_s)$ depends only on $\ell = |\sdif|$, and we emphasize this by writing it as $I_{\ell}$. Our setting readily handles both fixed and growing $k$; since the former is already well-understood \cite{Mal78,Ati12,Laa14}, we focus our attention on the case that $k\to\infty$, and in particular on the case that $k = \Theta(p^\theta)$ for some $\theta\in(0,1)$.
\begin{prop} \label{prop:gt_boundI}
Under the noiseless group testing setup, consider arbitrary sequences of sparsity levels $k\to\infty$ and $\ell \in \{1,\dotsc,k\}$, indexed by $p$. If $\frac{\ell}{k} = o(1)$, then
\begin{equation}
I_{\ell} = \bigg(e^{-\nu}\nu \frac{\ell}{k} \log\frac{k}{\ell}\bigg)(1+o(1)). \label{eq:gt_I_ord}
\end{equation}
Moreover, if $\frac{\ell}{k}\to\alpha \in(0,1]$, then
\begin{equation}
I_{\ell} = e^{-(1-\alpha)\nu} H_2\big(e^{-\alpha \nu}\big) (1+o(1)). \label{eq:gt_I_const}
\end{equation}
\end{prop}
\begin{proof}
See Appendix \ref{sec:PROOFS_GT}.
\end{proof}
We proceed with the steps of Procedure \ref{pr:procedure}.
\subsubsection*{Step 1}
The first step is trivial; $\beta_s$ is deterministic, and thus the typical set $\Tc_{\beta}$ is a singleton.
\subsubsection*{Step 2}
In contrast to the previous examples, we use different concentration inequalities to handle different values of $\ell$. Moreover, in accordance with Remark \ref{rem:delta2}, we let $\delta_2$ depend on $\ell$, writing it as $\delta_{2,\ell}$. For ``large'' values of $\ell$ (to be made precise below), we will apply the general bound in Proposition \ref{prop:gen_discrete} in Appendix \ref{sec:CONCENTRATION}. For ``small'' values of $\ell$, we use the following to obtain an improved bound.
\begin{prop} \label{prop:conc_gt}
For the noiseless group testing problem, consider sequences $k\to\infty$ and $\ell$, indexed by $p$, such that $\frac{\ell}{k} \to 0$. For any $\epsilon > 0$ and $\delta_{2,\ell} \in (0,1)$ bounded away from zero and one, the following holds for sufficiently large $p$:
\begin{multline}
\PP\Big[ \imath^n(\Xv_{\sdif}; \Yv | \Xv_{\seq}, b_s) \le nI_{\ell}(1-\delta_{2,\ell}) \Big] \\ \le \exp\bigg(-n\frac{\ell}{k} e^{-\nu}\nu\Big( (1-\delta_{2,\ell})\log(1-\delta_{2,\ell}) + \delta_{2,\ell} \Big)(1-\epsilon)\bigg) \label{eq:conc_gt}
\end{multline}
for all $(\sdif,\seq)$ with $|\sdif|=\ell$.
\end{prop}
\begin{proof}
See Appendix \ref{sec:PROOFS_GT}.
\end{proof}
From the bounds in \eqref{eq:conc_gt} and \eqref{eq:conc_gen_disc} in Appendix \ref{sec:CONCENTRATION}, we may fix $\epsilon > 0$ and choose the following when $p$ is sufficiently large:
\begin{itemize}
\item For $\ell \le \ell \le \lfloor\frac{k}{\log k}\rfloor$:
\begin{multline}
\hspace*{-6ex}\psi_\ell(n,\delta_{2,\ell}) = \\ \hspace*{-3ex} \exp\bigg(-n\frac{\ell}{k} e^{-\nu}\nu\Big( (1-\delta_{2,\ell})\log(1-\delta_{2,\ell}) + \delta_{2,\ell} \Big)(1-\epsilon)\bigg). \label{eq:gt_psi1}
\end{multline}
\item For $\ell > \lfloor\frac{k}{\log k}\rfloor$:
\end{itemize}
\begin{equation}
\psi_\ell(n,\delta_{2,\ell}) = 2\exp\bigg(- \frac{(\delta_{2,\ell} I_{\ell})^2 n}{2(16+2\delta_{2,\ell} I_{\ell})} \bigg). \label{eq:gt_psi2}
\end{equation}
For the converse, we only use the latter of these two cases, setting $\psi'_{\ell}(n,\delta_{2,\ell}) = 2\exp\big(- \frac{(\delta_{2,\ell} I_{\ell})^2 n}{2(16+2\delta_{2,\ell} I_{\ell})} \big)$.
\subsubsection*{Step 3}
Since $\beta_s$ is deterministic, we may trivially set $\gamma=0$ to obtain $P_0(\gamma) = 0$ in \eqref{eq:P_0}.
For the converse, we set $\Lc = \{k\}$ in Theorem \ref{thm:simplified_conv}. From the above choice of $\psi'_{\ell}$ and the growth of $I_k$ in \eqref{eq:gt_I_const}, we immediately obtain that $\psi'_{k} \to 0$ whenever $n=\omega(1)$. The achievability part requires more effort; we summarize the findings in the following proposition.
\begin{prop} \label{prop:gt_psi}
Let $k = \Theta( p^{\theta} )$ for some $\theta \in (0,1)$.
(i) For any $\eta > 0$, there exists $\delta_2^{(1)} \in (0,1)$ and a choice of $\epsilon > 0$ in \eqref{eq:gt_psi1} such that $\sum_{\ell=1}^{\lfloor \frac{k}{\log k} \rfloor} {k \choose \ell} \psi_{\ell}(n,\delta_{2}^{(1)}) \to 0$ provided that
\begin{equation}
n \ge \frac{ \frac{\theta}{1-\theta} k\log\frac{p}{k} }{ e^{-\nu}\nu } (1+\eta). \label{eq:gt_n_cond_psi}
\end{equation}
(ii) For any $\delta_2^{(2)} \in (0,1)$, we have $\sum_{\lfloor \frac{k}{\log k} \rfloor +1}^k {k \choose \ell} \psi_{\ell}(n,\delta_{2}^{(2)}) \to 0$ provided $n = \Omega\big(k\log\frac{p}{k}\big)$.
\end{prop}
\begin{proof}
See Appendix \ref{sec:PROOFS_GT}.
\end{proof}
The idea here is that for the smaller values of $\ell$, it is the concentration inequality that dominates the final bound, so we let $\delta_{2,\ell} = \delta_2^{(1)}$ be closer to one to provide better concentration behavior. For large values of $\ell$, the opposite is true, so we let $\delta_{2,\ell} = \delta_2^{(2)}$ be close to zero.
\subsubsection*{Step 4}
We obtain the following corollary by combining the previous steps and applying asymptotic simplifications.
\begin{cor} \label{cor:gt}
For the noiseless group testing problem with $k=\Theta(p^{\theta})$ ($\theta\in(0,1)$) and an optimized parameter $\nu$, we have $\pe \to 0$ as $p\to\infty$ provided that
\begin{equation}
n \ge \inf_{\nu>0}\max\bigg\{ \frac{ \theta }{ e^{-\nu}\nu(1-\theta) }, \frac{ 1 }{ H_2(e^{-\nu}) }\bigg\} \Big(k \log{\frac{p}{k}}\Big) (1 + \eta) \label{eq:gt_ach}
\end{equation}
for some $\eta > 0$. Conversely, we have $\pe \to 1$ as $p\to\infty$ whenever
\begin{equation}
n \le \frac{ k \log{\frac{p}{k}} }{ \log 2 } (1 - \eta) \label{eq:gt_conv}
\end{equation}
for some $\eta > 0 $.
\end{cor}
\begin{proof}
We first consider the achievability part. We immediately obtain the first term in the maximum in \eqref{eq:gt_ach} from \eqref{eq:gt_n_cond_psi}, so it remains to derive the second term. We start with \eqref{eq:final_ach}; by substituting $\gamma = 0$ and taking $\delta_1 \to 0$ sufficiently slowly, we obtain
\begin{equation}
n \ge \max_{\ell=1,\dotsc,k} \frac{ \log{ {{p-k} \choose \ell} } + 2\log\big(k {k \choose \ell}\big) }{I_{\ell}(1 - \delta_{2,\ell})} \big(1+o(1)\big). \label{eq:gt_n_cond3}
\end{equation}
Using \eqref{eq:gt_I_ord}--\eqref{eq:gt_I_const} and the asymptotic identity $\log {{p-k} \choose \ell} = \Theta\big( \ell \log \frac{p}{\ell} \big)$ we see that the objective in \eqref{eq:gt_n_cond3} behaves as
\begin{equation}
\Theta\bigg( \frac{k \log \frac{p}{\ell} }{ 1 + \log\frac{k}{\ell}} \bigg) \label{eq:gt_growth}
\end{equation}
whenever the constants $\{\delta_{2,\ell}\}$ are bounded away from one. This behaves as $\Theta\big(k \log \frac{p}{k}\big)$ when $\frac{\ell}{k} = \Theta(1)$, and as $\Theta\big( \frac{ k \log \frac{p}{k} }{ \log\frac{k}{\ell} } + k \big)$ when $\frac{\ell}{k} = o(1)$ (the latter of these is seen by writing $\log\frac{p}{\ell} = \log\frac{p}{k} + \log\frac{k}{\ell}$). Thus, the maximum in \eqref{eq:gt_n_cond3} can only be achieved by a sequence such that $\frac{\ell}{k} = \Theta(1)$. Moreover, with $\frac{\ell}{k} = \Theta(1)$, we see from the assumption $k = o(p)$ that the term $2\log\big(k {k \choose \ell}\big) = O(k)$ is dominated by $\log { {p-k} \choose \ell } = \Theta\big( k \log \frac{p}{k} \big)$, and can thus be factored into the $o(1)$ remainder term in \eqref{eq:gt_n_cond3}. This yields the condition
\begin{equation}
n \ge \max_{\ell=1,\dotsc,k} \frac{\ell \log \frac{p}{\ell}}{ I_{\ell}(1 - \delta_{2,\ell}) }\big(1+o(1)\big). \label{gt:growth2}
\end{equation}
Since the maximum can only be achieved asymptotically if $\frac{\ell}{k} = \Theta(1)$, we proceed by considering $\frac{\ell}{k} \to \alpha$ for some arbitrary $\alpha\in(0,1]$. Under this scaling, $\ell\log\frac{p}{\ell}$ behaves as $\big(\alpha k\log\frac{p}{k}\big) (1+o(1))$. Moreover, according to Proposition \ref{prop:gt_psi}, we can choose $\delta_{2,\ell}$ to be arbitrarily small for all $\ell$ values except those below $\lfloor \frac{k}{\log k} \rfloor$. Such values behave as $o(k)$, and thus do achieve the maximum in \eqref{gt:growth2}. Combining these observations with \eqref{eq:gt_I_const}, the right-hand side of \eqref{gt:growth2} yields the condition
\begin{equation}
n \ge \max_{\alpha\in(0,1]}\frac{\alpha k\log\frac{p}{k}}{ e^{-(1-\alpha)\nu} H_2\big(e^{-\alpha \nu}\big) } \big(1+\eta\big), \label{gt:growth3}
\end{equation}
where $\eta$ may be arbitrarily small. By a change of variable $\lambda = e^{-\alpha\nu}$, the coefficient to $k\log\frac{p}{k}$ can be written as $\frac{1}{\nu}e^{\nu} \frac{\lambda\log\frac{1}{\lambda}}{H_2(\lambda)}$. This is easily verified to be decreasing in $\lambda\in[0,1]$, which implies that the maximizing value of $\alpha$ is one, and yields the second term in \eqref{eq:gt_ach}.
The converse part is similar but considerably simpler; by setting $\Lc=\{k\}$ in Theorem \ref{thm:simplified_ach}, we obtain $\alpha=1$ immediately. The denominator $\log2$ in \eqref{eq:gt_conv} is obtained by maximizing $H_2(e^{-\nu})$ over $\nu$, and the condition $n\to\infty$ stated before Proposition \ref{prop:gt_psi} is clearly satisfied when \eqref{eq:gt_conv} holds with equality.
\end{proof}
By setting $\nu=\log2$ in \eqref{eq:gt_ach}, it is readily verified that the necessary and sufficient conditions coincide for $\theta \le \frac{1}{3}$, and in fact yield the same threshold as \emph{adaptive} group testing \cite{Bal13}. To our knowledge, this was only known previously in the limit as $\theta\to0$ \cite{Ald14}. Further comparisons to previous works are provided at the end of this subsection.
\subsubsection{Noisy Case with Exact Recovery} \label{sec:GT_NOISY}
We now turn to the noisy counterpart of \eqref{eq:gt_noiseless}:
\begin{equation}
Y = \openone\bigg\{ \bigcup_{i \in S} \{X_i = 1\} \bigg\} \oplus Z, \label{eq:gtn_model}
\end{equation}
where $Z\in\{0,1\}$ is additive noise, and $\oplus$ denotes modulo-2 addition. For concreteness, we focus on the case that $Z \sim \mathrm{Bernoulli}(\rho)$ for some $\rho\in(0,\frac{1}{2})$ not varying with $p$, though other noise models also fall into our framework (e.g., see \cite{Ati12}). As discussed below, we do not attempt to provide results with constants that are optimized to the same extent as the noiseless case, and we thus set $\nu = \log 2$, i.e., $P_X \sim \mathrm{Bernoulli}\big( \frac{\log 2}{k} \big)$.
We follow Procedure \ref{pr:procedure} in a similar fashion to the noiseless case, altering the statements of Proposition \ref{prop:gt_boundI}--\ref{prop:gt_psi} accordingly. To avoid repetition, we give the modified propositions and their proofs in Appendix \ref{sec:PROOFS_GT_NOISY}, and state the resulting corollary here. The main difference is that in the analog of Proposition \ref{prop:gt_psi}, we let $\delta_2^{(1)}$ remain arbitrary, thus leading to the optimization parameter $\delta_2$ in the following.
\begin{cor} \label{cor:gt_noisy}
Under the preceding setup for the noisy group testing problem with $\rho \in (0,0.5)$, $\nu = \log 2$, and $k=\Theta(p^{\theta})$ ($\theta\in(0,1)$), we have $\pe \to 0$ as $p\to\infty$ provided that
\begin{multline}
n \ge \inf_{\delta_2\in(0,1)}\max\bigg\{ \zeta(\rho,\delta_2,\theta) , \frac{ 1 }{\log 2 - H_2(\rho) }\bigg\} \Big( k \log\frac{p}{k} \Big) \\ \times (1 + \eta) \label{eq:gtn_ach}
\end{multline}
for some $\eta > 0$, where
\begin{multline}
\zeta(\rho,\delta_2,\theta) := \frac{2}{\log 2} \max\bigg\{\frac{ 2(1+\frac{1}{3}\delta_2 (1-2\rho)) \frac{\theta}{1-\theta}}{\delta_2^2 (1-2\rho)^2 } , \\ \frac{\frac{1+4\theta}{1-\theta} }{(1-2\rho)\log\frac{1-\rho}{\rho} (1-\delta_2)} \bigg\}. \label{eq:gt_zeta}
\end{multline}
Conversely, we have $\pe \to 1$ as $p\to\infty$ whenever
\begin{equation}
n \le \frac{ k \log{\frac{p}{k}} }{ \log 2 - H_2(\rho) } (1 - \eta). \label{eq:gtn_conv}
\end{equation}
for some $\eta > 0 $.
\end{cor}
\begin{proof}
See Appendix \ref{sec:PROOFS_GT_NOISY}.
\end{proof}
As we will see in the numerical examples below, Corollary \ref{cor:gt_noisy} provides an exact asymptotic threshold for a narrower range of $\theta$ values compared to the noiseless case. This is due to the difficulty in precisely characterizing the concentration behavior of the information density tail probabilities. Nevertheless, the second term in the maximum in \eqref{eq:gtn_ach} is always dominant for sufficiently small $\theta$, thus matching the converse. To see this, we first note that the first term in the maximum in \eqref{eq:gt_zeta} tends to zero as $\theta \to 0$, and cannot be dominant in this limit. This implies that $\delta_2$ may be arbitrarily close to zero provided that $\theta$ is sufficiently small. Assuming then that $\delta_2 $ and $\theta$ are small and the maximum in \eqref{eq:gt_zeta} is achieved by the second term, we can write $\zeta(\rho,\delta_2,\theta) \approx \frac{2}{\log 2}\frac{1}{ (1-2\rho) \log\frac{1-\rho}{\rho} }$. This is strictly smaller than $\frac{1}{\log 2 - H_2(\rho)}$; see Proposition \ref{prop:gt_cmp} in Appendix \ref{sec:PROOFS_GT_NOISY}.
\subsubsection{Partial Recovery}
The consideration of partial recovery (\emph{cf.}~\eqref{eq:pe_partial}) in fact leads to \emph{simpler} expressions and proofs, as seen in the following.
\begin{cor} \label{cor:gt_partial}
Under the preceding setup for the group testing problem with $\rho\in[0,0.5)$ (i.e., possibly noiseless), $\nu=\log 2$, $k\to\infty$, $k=o(p)$, and $\dmax = \lfloor \alpha^* k \rfloor$ for some $\alpha^*\in(0,1)$, we have $\pe(\dmax) \to 0$ as $p\to\infty$ provided
\begin{equation}
n \ge \frac{ k \log{\frac{p}{k}} }{ \log 2 - H_2(\rho) } (1 + \eta) \label{eq:gt_ach3}
\end{equation}
for some $\eta > 0$. Conversely, $\pe(\dmax) \to 1$ as $p\to\infty$ whenever
\begin{equation}
n \le \frac{(1-\alpha^{*}) \big(k \log{\frac{p}{k}}\big)}{ \log 2 - H_2(\rho) } (1 - \eta) \label{eq:gt_conv3}
\end{equation}
for some $\eta > 0 $.
\end{cor}
\begin{proof}
The achievability part follows the proofs of Corollaries \ref{cor:gt} and \ref{cor:gt_noisy}, except that the ``small'' values of $\ell$ need not be handled. That is, we only make use of the general concentration inequality in \eqref{eq:conc_gen_disc} in Appendix \ref{sec:CONCENTRATION}, and we end up with the single condition in \eqref{eq:gt_ach3}. For the converse part, we again choose $\Lc=\{k\}$ in Theorem \ref{thm:simplified_conv_p}, and the steps are again similar, with the multiplicative factor $1-\alpha^{*}$ arising via identical reasoning to Corollary \ref{cor:linear_partial}.
\end{proof}
Corollary \ref{cor:gt_partial} shows that at least for sufficiently small $\theta$ (e.g., $k = O\big(p^\frac{1}{3}\big)$ in the noiseless case), there is not much to be saved by moving from exact recovery to partial recovery: Allowing for a fraction $\alpha^{*}$ of errors leads to at most a reduction in the number of measurements of a multiplicative factor $1-\alpha^{*}$.
\subsubsection{Numerical Evaluations}
In Figure \ref{fig:gt_noiseless}, we compare the bounds in Corollary \ref{cor:gt} with existing asymptotic bounds in the literature. For convenience, we switch to base-2 logarithms and plot the asymptotic limit of the ratio $\frac{k\log_2\frac{p}{k}}{n}$, so that a higher value corresponds to fewer measurements. We see that our achievability bound improves on all of the existing bounds; however, we note that the Combinatorial Optimal Matching Pursuit (COMP) \cite{Cha11} and Definite Defective (DD) \cite{Ald14a} algorithms are computationally tractable and do not require knowledge of $k$.
\begin{figure}
\begin{centering}
\includegraphics[width=0.95\columnwidth]{GT_Asymp.pdf}
\par
\end{centering}
\caption{ Asymptotic thresholds on the number of measurements required for noiseless group testing. \label{fig:gt_noiseless}}
\end{figure}
\begin{figure}
\begin{centering}
\includegraphics[width=0.95\columnwidth]{GT_Asymp_Noisy2.pdf}
\par
\end{centering}
\caption{ Asymptotic thresholds on the number of measurements required for noisy group testing. \label{fig:gt_noisy}}
\end{figure}
The converse bound shown is known to hold even for adaptive measurement matrices \cite{Bal13}. Thus, a key implication of our results is that adaptivity provides no asymptotic gain over non-adaptive Bernoulli measurements when $k = O(p^{\frac{1}{3}})$. It remains an important open problem to derive \emph{practical} decoding schemes for achieving the bound in the non-adaptive setting.
Figure \ref{fig:gt_noisy} provides an analogous plot for the noisy case, with three different noise levels (i.e., values of $\rho$). In each case, we obtain an exact threshold for sufficiently small $\theta$, albeit over a narrower range than the noiseless case. Once again, the converse is known to hold even in the adaptive setting \cite{Cha11}, and we have thus provided cases where non-adaptive Bernoulli measurements yield the same asymptotics as optimal adaptive measurements. To our knowledge, this has not been shown previously even in the limit as $\theta \to 0$.
\subsection{General Strong Impossibility result for Discrete Observation Models} \label{sec:STRONG_IMP_DISC}
Equation \eqref{eq:uniformV} in Appendix \ref{sec:CONCENTRATION} bounds the variance of the information density uniformly in terms of the output alphabet size for models with discrete observations. Notable examples include group testing, the 1-bit model (or more generally, quantizations with more than two levels), and logistic regression. We obtain the following general strong impossibility result (i.e., conditions under which $\pe \to 1$) by combining Proposition \ref{prop:chebyshev} in Appendix \ref{sec:CONCENTRATION} with a variant of Theorem \ref{thm:simplified_conv}.
\begin{cor} \label{cor:gen_finite}
If the observations lie in a finite set $\Yc \subset \RR$ with probability one, then $\pe \to 1$ whenever there exist vanishing sequences $\delta_{1,p} \to 0$ and $\epsilon_{p} \to 0$ such that
\begin{equation}
n \ge \max_{(\sdif,\seq)\,:\,\sdif\ne\emptyset} \frac{\log {{p-k+|\sdif|} \choose |\sdif|} - \log\delta_{1,p} }{ I_{\sdif,\seq}(b_s) + \sqrt{\frac{|\Yc|}{n\epsilon_{p}}} } \label{eq:gen_finite}
\end{equation}
for all $b_s \in \RR^k$ within a set whose probability under $P_{\beta_s}$ approaches one.
\end{cor}
\begin{proof}
In this application, we do not use Theorem \ref{thm:simplified_conv} directly, but instead follow the arguments leading up to it with \eqref{eq:B_cond_conv}--\eqref{eq:psi_conv} replaced by
\begin{equation}
\log{{p-k+|\sdif|} \choose |\sdif|} - \log \delta_1 \ge n(I_{\sdif,\seq}(b_s) + \delta); \label{eq:B_cond_conv_2}
\end{equation}
and
\begin{multline}
\PP\Big[ \imath^n(\Xv_{\sdif}; \Yv | \Xv_{\seq}, b_s) \le n(I_{\sdif,\seq}(b_s) + \delta) \,\big|\, \beta_s = b_s \Big] \\ \ge 1 - \frac{|\Yc|(\frac{4}{e})^2}{\delta^2 n}. \label{eq:psi_conv_2}
\end{multline}
By Proposition \ref{prop:chebyshev} in Appendix \ref{sec:CONCENTRATION}, we have for all $(\sdif,\seq,b_s)$ that \eqref{eq:psi_conv_2} holds, so the analogous probability to that on the right-hand side of \eqref{eq:nonasymp_conv} is dictated only by \eqref{eq:B_cond_conv_2}. Moreover, the right-hand side of \eqref{eq:psi_conv_2} tends to one upon setting $\delta = \sqrt{\frac{|\Yc|}{n\epsilon_p}}$ for some $\epsilon_p \to 0$. By also setting $\delta_1 = \delta_{1,p} \to 0$ (so that the analogous additive term to that of $\delta_1$ in \eqref{eq:nonasymp_conv} vanishes), we see that \eqref{eq:B_cond_conv_2} coincides with \eqref{eq:gen_finite}.
\end{proof}
When $\beta_s$ deterministic, this theorem recovers a recent result by Tan and Atia \cite{Tan14}, which was proved using combinatorial techniques. Our result is in fact slightly stronger in the sense that the additive term in the denominator only behaves as $\omega\big(\frac{1}{\sqrt{n}}\big)$, whereas the corresponding term in \cite{Tan14} behaves as $\omega\big(\frac{1}{n^{1/4}}\big)$. Thus, in our result, the mutual information term remains the dominant one in a wider range of settings.
\section{Proofs of General Bounds} \label{sec:PROOFS}
Here we provide the proofs of Theorems \ref{thm:ach1} and \ref{thm:conv}, and then give the changes required to obtain the results for partial recovery in Section \ref{sec:PARTIAL}. As mentioned previously, the proofs bear some resemblance to those of mixed channels in channel coding \cite[Sec.~3.3]{Han03}. However, the analysis here is more involved, primarily due to the fact that the ``codewords'' $\Xv_{s}$ are not independent for different values of $s\in\Sc$, but instead share common columns corresponding to the overlapping parts of the support set. See \cite{Wai09,Wan10,Jin11} for further discussions on the differences between support recovery and channel coding.
\subsection{Proof of Theorem \ref{thm:ach1}} \label{sec:PROOF_GEN_ACH}
\subsubsection{Initial Non-Asymptotic Bound}
Recall the definitions of the random variables in \eqref{eq:joint_dist}--\eqref{eq:joint_dist_n}, and the information densities in \eqref{eq:idens_multi}--\eqref{eq:idens_b}. We fix the constants $\gamma_{1},\dotsc,\gamma_{k}$ arbitrarily, and consider a decoder that searches for the unique set $s\in\Sc$ such that
\begin{equation}
\imath(\xv_{\sdif}; \yv | \xv_{\seq} ) > \gamma_{|\sdif|} \label{eq:decoder_cond}
\end{equation}
for all $2^{k} - 1$ partitions $(\sdif,\seq)$ of $s$ with $\sdif\ne\emptyset$. An error occurs if no such $s$ exists, if multiple exist, or if such a set differs from the true value.
Since the joint distribution of $(\beta_s,\Xv_s,\Yv_s \,|\, S=s)$ is the same for all $s$ in our setup (\emph{cf.}~Section \ref{sec:SETUP}), and the decoder that we have chosen exhibits a similar symmetry, we can condition on a fixed and arbitrary value of $S$, say $s=\{1,\dotsc,k\}$. By the union bound, the error probability is upper bounded by
\begin{multline}
\pe \le \PP\bigg[ \bigcup_{(\sdif,\seq)} \Big\{ \imath(\Xv_{\sdif}; \Yv | \Xv_{\seq} ) \le \gamma_{|\sdif|} \Big\} \bigg] \\ + \sum_{\sbar \in \Sc \backslash \{s\}} \PP\Big[\imath(\Xvdif; \Yv | \Xveq) > \gamma_{|\sdif|} \Big], \label{eq:thresh2}
\end{multline}
where here and subsequently we let the condition $\sdif\ne\emptyset$ remain implicit. The first term in \eqref{eq:thresh2} corresponds to the true set failing the threshold test, and the second term corresponds to some incorrect set $\sbar$ passing the threshold test. In the summand of the second term, we have upper bounded the probability of an intersection of $2^k - 1$ events by just one such event, namely, the one corresponding to $\sdif = \sbar \backslash s$ and $\seq = s \cap \sbar$.
Using the shorthand $\ell := |\sbar \backslash s|$, we can weaken the second probability in \eqref{eq:thresh2} as follows:
\begin{align}
&\PP\Big[\imath(\Xvdif; \Yv | \Xveq) > \gamma_{\ell} \Big] \nonumber \\
&= \sum_{\xveq,\xvdif,\yv} P_X^{n \times (k-\ell)}(\xveq) P_X^{n \times \ell}(\xvdif) P_{\Yv|\Xv_{\seq}}(\yv|\xveq) \nonumber \\
&\qquad \times\openone\bigg\{ \log\frac{P_{\Yv|\Xv_{\sdif}\Xv_{\seq}}(\yv|\xvdif,\xveq)}{P_{\Yv|\Xv_{\seq}}(\yv|\xveq)} > \gamma_{\ell} \bigg\} \label{eq:ach_na_3} \\
&\le \sum_{\xveq,\xvdif,\yv} P_X^{n \times (k-\ell)}(\xveq) P_X^{n \times \ell}(\xvdif) \nonumber \\
&\qquad \times P_{\Yv|\Xv_{\sdif}\Xv_{\seq}}(\yv|\xvdif,\xveq) e^{-\gamma_{\ell}} \label{eq:ach_na_4} \\
&= e^{-\gamma_{\ell}}, \label{eq:ach_na_5}
\end{align}
where in \eqref{eq:ach_na_3} we used the fact that the output vector depends only on the columns of $\xv_{\sbar}$ corresponding to entries of $\sbar$ that are also in $s$, and \eqref{eq:ach_na_4} follows by bounding $P_{\Yv|\Xv_{\seq}}$ using the event within the indicator function, and then upper bounding the indicator function by one. Substituting \eqref{eq:ach_na_5} into \eqref{eq:thresh2} gives
\begin{multline}
\pe \le \PP\bigg[ \bigcup_{(\sdif,\seq)} \Big\{ \imath(\Xv_{\sdif}; \Yv | \Xv_{\seq} ) \le \gamma_{\ell} \Big\} \bigg] \\ + \sum_{\ell=1}^k {{p-k} \choose \ell}{k \choose \ell}e^{-\gamma_{\ell}}, \label{eq:thresh3}
\end{multline}
where the combinatorial terms arise from a standard counting argument \cite{Wai09}.
Note that while the bound in \eqref{eq:thresh3} appears to be simpler than that in the theorem statement, it is difficult to directly apply it to specific problems, since $\imath(\Xv_{\sdif}; \Yv | \Xv_{\seq} )$ is not an i.i.d.~summation in general.
\subsubsection{Completion of the Proof}
We fix the constants $\gamma'_1,\dotsc,\gamma'_{\ell}$ arbitrarily, and apply the following elementary steps with $\ell = |\sdif|$:
\begin{align}
&\PP\bigg[ \bigcup_{(\sdif,\seq)} \Big\{ \imath(\Xv_{\sdif}; \Yv | \Xv_{\seq} ) \le \gamma_{\ell} \Big\} \bigg] \nonumber \\
&= \PP\bigg[ \bigcup_{(\sdif,\seq)} \bigg\{ \log\frac{ P_{\Yv|\Xv_{\sdif}\Xv_{\seq}}(\Yv|\Xv_{\sdif},\Xv_{\seq}) }{ P_{\Yv|\Xv_{\seq}}(\Yv|\Xv_{\seq}) } \le \gamma_{\ell} \bigg\} \bigg] \label{eq:ach_disc1} \\
&\le \PP\bigg[ \bigcup_{(\sdif,\seq)} \bigg\{ \log\frac{ P_{\Yv|\Xv_{\sdif}\Xv_{\seq}}(\Yv|\Xv_{\sdif},\Xv_{\seq}) }{ P_{\Yv|\Xv_{\seq}}(\Yv|\Xv_{\seq}) } \le \gamma_{\ell} \nonumber \\
& \qquad\qquad\qquad \,\cap\,\log\frac{ P_{\Yv|\Xv_{\seq}}(\Yv|\Xv_{\seq}) }{ P_{\Yv|\Xv_{\seq}\beta_s}(\Yv|\Xv_{\seq},\beta_s) } \le \gamma'_{\ell} \bigg\} \bigg] \nonumber \\
& \quad + \PP\bigg[ \bigcup_{(\sdif,\seq)} \bigg\{ \log\frac{ P_{\Yv|\Xv_{\seq}}(\Yv|\Xv_{\seq}) }{ P_{\Yv|\Xv_{\seq}\beta_s}(\Yv|\Xv_{\seq},\beta_s) } > \gamma'_{\ell} \bigg\} \bigg] \label{eq:ach_disc2}
\end{align}
\begin{align}
&\le \PP\bigg[ \bigcup_{(\sdif,\seq)} \bigg\{ \log\frac{ P_{\Yv|\Xv_{\sdif}\Xv_{\seq}}(\Yv|\Xv_{\sdif},\Xv_{\seq}) }{ P_{\Yv|\Xv_{\seq}\beta_s}(\Yv|\Xv_{\seq},\beta_s) } \le \tilde{\gamma}_{\ell} \bigg\} \bigg] \nonumber \\
& \quad + \PP\bigg[ \bigcup_{(\sdif,\seq)} \bigg\{ \log\frac{ P_{\Yv|\Xv_{\seq}}(\Yv|\Xv_{\seq}) }{ P_{\Yv|\Xv_{\seq}\beta_s}(\Yv|\Xv_{\seq},\beta_s) } > \gamma'_{\ell} \bigg\} \bigg],\label{eq:fein1_split}
\end{align}
where $\tilde{\gamma}_{\ell} = \gamma_{\ell} + \gamma'_{\ell}$. The second term in \eqref{eq:fein1_split} is upper bounded as
\begin{align}
& \PP\bigg[ \bigcup_{(\sdif,\seq)} \bigg\{ \log\frac{ P_{\Yv|\Xv_{\seq}}(\Yv|\Xv_{\seq}) }{ P_{\Yv|\Xv_{\seq}\beta_s}(\Yv|\Xv_{\seq},\beta_s) } > \gamma'_{\ell} \bigg\} \bigg] \nonumber \\
&\le \sum_{(\sdif,\seq)} \PP\bigg[ \log\frac{ P_{\Yv|\Xv_{\seq}}(\Yv|\Xv_{\seq}) }{ P_{\Yv|\Xv_{\seq}\beta_s}(\Yv|\Xv_{\seq},\beta_s) } > \gamma'_{\ell} \bigg] \label{eq:ach_denom_ub0} \\
& = \sum_{(\sdif,\seq)}\sum_{b_s,\xv_{\seq},\yv}P_{\beta_s}(b_s)P_{X}^{n\times(k-\ell)}(\xv_{\seq}) \nonumber \\
& \quad\qquad\times P_{\Yv|\Xv_{\seq}\beta_s}(\yv|\xv_{\seq},b_s) \nonumber \\
& \quad\qquad \times\openone\bigg\{ \log\frac{ P_{\Yv|\Xv_{\seq}}(\yv|\xv_{\seq}) }{ P_{\Yv|\Xv_{\seq}\beta_s}(\yv|\xv_{\seq},b_s) } > \gamma'_{\ell} \bigg\} \\
&\le \sum_{(\sdif,\seq)}\sum_{b_s,\xv_{\seq},\yv}P_{\beta_s}(b_s)P_{X}^{n\times(k-\ell)}(\xv_{\seq}) \nonumber \\
& \qquad\qquad\qquad\qquad\qquad \times P_{\Yv|\Xv_{\seq}}(\yv|\xv_{\seq}) e^{-\gamma'_{\ell}} \\
& = \sum_{\ell=1}^k {k \choose \ell}e^{-\gamma'_{\ell}}, \label{eq:ach_denom_ub}
\end{align}
where \eqref{eq:ach_denom_ub0} follows from the union bound, and the remaining steps follow the arguments used in \eqref{eq:ach_na_3}--\eqref{eq:ach_na_5}.
We now upper bound the first term in \eqref{eq:fein1_split}. The numerator in \eqref{eq:fein1_split} equals $P_{\Yv|\Xv_{s}}(\Yv|\Xv_{s})$ for all $(\sdif,\seq)$ (\emph{cf.}, \eqref{eq:p_split1}), and we can thus write the overall term as
\begin{multline}
\PP\bigg[ \log P_{\Yv|\Xv_{s}}(\Yv|\Xv_{s}) \\ \le \max_{(\sdif,\seq)} \big\{ \log P_{\Yv|\Xv_{\seq}\beta_s}(\Yv|\Xv_{\seq},\beta_s) + \gamma_{\ell} + \gamma'_{\ell} \big\} \bigg]. \label{eq:simp_gen_1}
\end{multline}
Using the same steps as those used in \eqref{eq:ach_disc1}--\eqref{eq:fein1_split}, we can upper bound this by
\begin{multline}
\hspace*{-2ex}\PP\bigg[ \log P_{\Yv|\Xv_s\beta_s}(\Yv|\Xv_{s},\beta_s) \\ \le \max_{(\sdif,\seq)} \big\{ \log P_{\Yv|\Xv_{\seq}\beta_s}(\Yv|\Xv_{\seq},\beta_s) + \gamma_{\ell} + \gamma'_{\ell} + \gamma \big\} \bigg] \\ + \PP\bigg[ \log \frac{P_{\Yv|\Xv_s,\beta_s}(\Yv|\Xv_{s},\beta_s)}{P_{\Yv|\Xv_{s}}(\Yv|\Xv_{s})} > \gamma \bigg] \label{eq:simp_gen_2}
\end{multline}
for any constant $\gamma$. Reversing the step in \eqref{eq:simp_gen_1}, this can equivalently be written as
\begin{multline}
\hspace*{-2ex}\PP\bigg[ \bigcup_{(\sdif,\seq)} \bigg\{ \log\frac{ P_{\Yv|\Xv_{\sdif}\Xv_{\seq}\beta_s}(\Yv|\Xv_{\sdif},\Xv_{\seq},\beta_s) }{ P_{\Yv|\Xv_{\seq}\beta_s}(\Yv|\Xv_{\seq},\beta_s) } \\ \le \gamma_{\ell} + \gamma'_{\ell} + \gamma \bigg\} \bigg] + \PP\bigg[ \log \frac{P_{\Yv|\Xv_s,\beta_s}(\Yv|\Xv_{s},\beta_s)}{P_{\Yv|\Xv_{s}}(\Yv|\Xv_{s})} > \gamma \bigg]. \label{eq:simp_gen_3}
\end{multline}
Observe that the first logarithm appearing here is precisely the information density in \eqref{eq:idens_bn}. Moreover, the choices
\begin{align}
\gamma_{\ell} &= \log\bigg(\frac{k}{\delta_1}{{p-k} \choose \ell}{k \choose \ell}\bigg) \label{eq:gamma_l} \\
\gamma'_{\ell} &= \log\bigg(\frac{k}{\delta_1}{k \choose \ell}\bigg) \label{eq:gamma'_l}
\end{align}
make \eqref{eq:ach_denom_ub} and the second term in \eqref{eq:thresh3} be upper bounded by $\delta_1$ each. Hence, and combining \eqref{eq:fein1_split} with \eqref{eq:ach_denom_ub} and \eqref{eq:simp_gen_3}, and recalling that $\ell = |\sdif|$, we obtain \eqref{eq:thresh_sl}.
\subsection{Proof of Theorem \ref{thm:conv}} \label{sec:PROOF_GEN_CONV}
As has been done in several previous proofs of information-theoretic converse bounds for sparsity pattern recovery \cite{Wan10,Ree13,Ati12}, we consider an argument based on a genie. As explained formally below, the genie reveals some of elements of the support set to the decoder, which is left to estimate the remaining entries. An important novelty in our arguments is that we also let the revealed indices depend on the random non-zero entries of $\beta$; this leads to the improvement stated following Theorem \ref{thm:simplified_conv}.
It will prove convenient to present the proof under the following assumption of symmetry.
\begin{assump} \label{ass:genie}
The pair $(\sdif(b_s),\seq(b_s))$ in Theorem \ref{thm:conv} satisfies the following property: If $b'_s$ is a permutation of $b_s$, then the entries of $b'_s$ indexed by $\sdif(b'_s)$ (respectively, $\seq(b'_s)$) are a permutation of the entries of $b_s$ indexed by $\sdif(b_s)$ (respectively, $\seq(b_s)$).
\end{assump}
We claim that the theorem statement under this assumption also implies the more general case. To see this, we use the symmetry of $P_{Y|X_S\beta_S}$ with respect to $S$ from in Section \ref{sec:SETUP}, and the fact that $\Xv$ has an i.i.d.~distribution. Among all the possible choices of functions $(\sdif(\cdot),\seq(\cdot))$, there always exists a pair that maximizes the lower bound in \eqref{eq:conv_sl} and satisfies Assumption \ref{ass:genie}. More precisely, for any realization of $\beta_s$, the probability in \eqref{eq:conv_sl} is determined by entries appearing in the partition $(\beta_{\sdif},\beta_{\seq})$ but not by their order, so one can always maximize \eqref{eq:conv_sl} by forming this partition in a manner which is symmetric with respect to permutations of $\beta_s$.
We now formally define the genie-aided setup as follows:
\begin{enumerate}
\item Generate a random $k$-dimensional vector $\betatil' \sim P_{\beta_S}$.
\item Given $\betatil'$, let $\betatildif'$ and $\betatileq'$ be the subvectors indexed by $\sdif(\betatil')$ and $\seq(\betatil')$ respectively.
\item Let $\betatildif$ (respectively, $\betatileq$) be a uniformly random permutation of $\betatildif'$ (respectively, $\betatileq'$).
\item Generate $\Seq$ uniformly on $\Sceq(\ell)$, defined to contain the $p \choose k - \ell$ subsets of $\{1,\dotsc,p\}$ having cardinality $k - \ell$, where $\ell = |\betatildif|$. Set $\beta_{\Seq} = \betatileq$.
\item Generate $\Sdif$ uniformly on $\Scdif(\Seq)$, defined to contain the ${p - k + \ell} \choose \ell$ subsets of $\{1,\dotsc,p\} \backslash \Seq$ having cardinality $\ell$. Set $\beta_{\Sdif} = \betatildif$.
\item Set $S = \Sdif \cup \Seq$ and $\beta_{S^c} = 0$. The measurement matrix $\Xv$ is i.i.d.~on $P_{X}$, and the observation vector $\Yv$ is generated from $S$, $\Xv$, and $\beta$ according to \eqref{eq:single_output}, as in the original problem setup.
\item Reveal the indices $\Seq$ and the vectors $\betatildif$ and $\betatileq$ to the decoder. The decoder forms an estimate $\Sdifhat$ of $\Sdif$, and an error occurs if $\Sdifhat \ne \Sdif$.
\end{enumerate}
The joint distribution of $S$ and $\beta$ is the same here as in the original setup: The support set is uniform on the $p \choose k$ elements of $\Sc$, and the distribution of the non-zero entries $\beta_S$ is that of a uniformly random permutation of $\betatil' \sim P_{\beta_S}$. Since $P_{\beta_S}$ is permutation-invariant by assumption, this yields $\beta_S \sim P_{\beta_S}$ as required. Thus, the only difference in this modified setup is that the decoder has further information, and it follows that any converse for this setup implies the same converse for the original setup.
Throughout the proof, we make use of the random variables defined in the preceding steps, departing from the notation implicitly conditioned on $S$ equaling a fixed value $s$ (see \eqref{eq:joint_dist_n}) until the final step in obtaining \eqref{eq:conv_sl}.
We first study the error probability for the genie-aided setting conditioned on $(\Seq,\betatildif,\betatileq) = (\seq,\btildif,\btileq)$, denoted by $\pe(\seq,\btildif,\btileq)$. By the identity $\PP[\Ac] = \PP[\Ac \cap \Ec] + \PP[\Ac \cap \Ec^c]$, we have for any event $\Ac(\seq,\btildif,\btileq)$ that
\begin{multline}
\pe(\seq,\btildif,\btileq) \ge \PP[\Ac(\seq,\btildif,\btileq)] \\ - \PP[\Ac(\seq,\btildif,\btileq) \cap \text{no error}]. \label{eq:conv_init}
\end{multline}
We fix the constant $\gamma_{\ell}$ and choose
\begin{equation}
\Ac(\seq,\btildif,\btileq) = \big\{ \imath^n(\Xv_{\Sdif};\Yv|\Xv_{\seq},\btil) \le \gamma_{\ell} \big\},
\end{equation}
where $\ell = k - |\seq|$, and $\btil := \btil(\btildif,\btileq,\sdif,\seq)$ equals $\btildif$ (respectively, $\btileq$) on the entries indexed by $\sdif$ (respectively, $\seq$). Using the definitions in \eqref{eq:idens_bn}--\eqref{eq:idens_b}, and defining $\Dc(\sdif | \seq,\btildif,\btileq)$ to be the set of pairs $(\xv,\yv)$ such that the decoder outputs $\sdif$ given $(\seq,\btildif,\btileq,\xv,\yv)$, we obtain
\begin{align}
&\PP[\Ac(\seq,\btildif,\btileq) \cap \text{no error}] \nonumber \\
&=\sum_{\sdif\in\Scdif(\seq)} \frac{1}{{{p-k+\ell} \choose \ell}} \sum_{(\xv,\yv)\in\Dc(\sdif | \seq,\btildif,\btileq)} P_{X}^{n \times p}(\xv) \nonumber \\
&\qquad \times P^n_{Y|X_{\sdif}X_{\seq}\beta_s}(\yv|\xv_{\sdif},\xv_{\seq},\btil) \nonumber \\
&\qquad \times \openone\bigg\{ \log\frac{P^n_{Y|X_{\sdif}X_{\seq}\beta_s}(\yv|\xv_{\sdif},\xv_{\seq},\btil)}{P^n_{Y|X_{\seq}\beta_s}(\yv|\xv_{\seq},\btil)} \le \gamma_{\ell} \bigg\} \label{eq:vh_deriv1}
\end{align}
\begin{align}
&\le \frac{1}{{{p-k+\ell} \choose \ell}}\sum_{\sdif\in\Scdif(\seq)} \sum_{(\xv,\yv)\in\Dc(\sdif | \seq,\btildif,\btileq)} P_{X}^{n \times p}(\xv) \nonumber \\
&\qquad\qquad\qquad\qquad\qquad \times P^n_{Y|X_{\seq}\beta_s}(\yv|\xv_{\seq},\btil)e^{\gamma_{\ell}} \label{eq:vh_deriv2} \\
&= \frac{e^{\gamma_{\ell}}}{{ {p-k+\ell} \choose \ell }}, \label{eq:vh_deriv3}
\end{align}
where \eqref{eq:vh_deriv1} follows since an error occurs if and only if $(\xv,\yv)\notin\Dc(\sdif | \seq,\btildif,\btileq)$, \eqref{eq:vh_deriv2} follows by upper bounding $P^n_{Y|X_{\seq}\beta_s}$ using the event in the indicator function, and \eqref{eq:vh_deriv3} follows since the sets $\Dc(\sdif | \seq,\btildif,\btileq)$ are disjoint, and their union over $\sdif$ is the entire space of $(\xv,\yv)$ pairs.
Averaging \eqref{eq:conv_init} over $(\Seq,\betatil',\betatildif,\betatileq)$ and applying \eqref{eq:vh_deriv3}, we obtain
\begin{align}
\pe &\ge \sum_{\btil'} P_{\beta_S}(\btil') \sum_{\btildif,\btileq} \PP\big[ (\betatildif,\betatileq) = (\btildif,\btileq) \,|\, \btil' \big] \nonumber \\
&\times \sum_{\seq\in\Seq(\ell)}\sum_{\sdif\in\Sdif(\seq)} \frac{1}{{p \choose {k -\ell}}} \frac{1}{{{p-k+\ell} \choose \ell}} \nonumber \\
&\times \bigg(\PP\Big[ \imath^n(\Xv_{\sdif};\Yv|\Xv_{\seq},\btil) \le \gamma_{\ell} \,\big|\, \sdif,\seq,\btildif,\btileq \Big] \nonumber \\
&\qquad\qquad\qquad\qquad\qquad\qquad\qquad- \frac{ e^{\gamma_{\ell}}}{{ {p-k+\ell} \choose \ell }}\bigg),\label{eq:vh_deriv4}
\end{align}
where $\ell = |\btildif|$, and the conditioning on $\btil'$ is a shorthand for $\betatil' = \btil'$, and similarly for the second probability. Finally, we claim that this recovers \eqref{eq:conv_sl} upon setting
\begin{align}
\gamma_{\ell} &= \log{{p-k+\ell} \choose \ell} + \log \delta_1.
\end{align}
To see this, we first note that all of the terms in the summations over $\sdif$ and $\seq$ in \eqref{eq:vh_deriv4} are equal, since in the probability appearing in the summand, the entries $\btil_{\sdif}$ and $\btil_{\seq}$ are the same for any such pair, namely, $\btil_{\sdif} = \btildif$ and $\btil_{\seq} = \btileq$ (recall also the symmetry of $P_{Y|X_S\beta_S}$ with respect to $S$ assumed in Section \ref{sec:SETUP}). Due to Assumption \ref{ass:genie}, this probability also coincides with that in \eqref{eq:conv_sl} with $b_s := \btil'$, regardless of the realization of $(\betatildif,\betatileq)$ given $\betatil'$; the only randomness in the corresponding distribution is that of the two random permutations of the subvectors.
\subsection{Extensions to Partial Recovery} \label{sec:PROOF_PARTIAL}
The achievability analysis in Section \ref{sec:PROOF_GEN_ACH} extends immediately to handle the partial recovery criterion in \eqref{eq:pe_partial}, since we have already split the error events according to the amount of overlap between the true support and the incorrect support. The only difference is that the decoder searches for a set $s$ such that \eqref{eq:decoder_cond} holds whenever $|\sdif| > \dmax$ (as opposed to $\sdif\ne\emptyset$), and chooses one arbitrarily if multiple such $s$ exist. It follows that Theorem \ref{thm:ach1} remains true when the union in \eqref{eq:thresh_sl} is restricted to $|\sdif|\in\{\dmax+1,\dotsc,k\}$.
The extension of the converse analysis in Section \ref{sec:PROOF_GEN_CONV} is less immediate, but still straightforward. We first recall the observation from \cite{Ree13} that the performance metric in \eqref{eq:pe_partial} allows us to focus without loss of generality on decoders such that the estimated support $\hat{S}$ (or $\Sdifhat \cup \Seq$ in the genie-aided setting) has cardinality $k$ almost surely. For any such decoder, the definition in \eqref{eq:pe_partial} is unchanged when the second term in the union is removed.
We restrict the partitions $(\sdif(b_s),\seq(b_s))$ of $s$ to satisfy $|\sdif(b_s)| > \dmax$. In \eqref{eq:vh_deriv1}--\eqref{eq:vh_deriv2}, we change the definition of $\Dc(\sdif | \seq,\btildif,\btileq)$ to be the set of pairs $(\xv,\yv)$ such that the decoder outputs a sequence $\sdifhat$ such that $|\sdif \backslash \sdifhat| \le \dmax$. This means that the sets $\Dc(\cdot | \seq,\btildif,\btileq)$ are no longer disjoint. However, we can easily count the number of such sets that each $(\xv,\yv)$ pair falls into. For fixed $(\seq,\sdif)$ and $d\in\{0,\dotsc,\dmax\}$, the number of sets $\sdifhat \subseteq \{1,\dotsc,p\} \backslash \seq$ such that $|\sdif \backslash \sdifhat| = d$ is ${{p-k} \choose d}{|\sdif| \choose {|\sdif| - d}} = {{p-k} \choose d}{|\sdif| \choose d}$. Thus, each $(\xv,\yv)$ pair is included in $\sum_{d=0}^{\dmax}{{p-k} \choose d}{|\sdif| \choose d}$ of the sets $\Dc(\cdot | \seq,\btildif,\btileq)$, and \eqref{eq:vh_deriv3} is replaced by
\begin{align}
\PP[\Ac(\seq,\btildif,\btileq) \cap \text{no error}] \le \frac{ \sum_{d=0}^{\dmax}{{p-k} \choose d}{\ell \choose d} }{{ {p-k+\ell} \choose \ell }} e^{-\gamma_{\ell}}.
\end{align}
Thus, Theorem \ref{thm:conv} remains true when the pair $(\sdif(\cdot),\seq(\cdot))$ is constrained to satisfy $|\sdif|\in\{\dmax+1,\dotsc,k\}$, and ${ {p-k+|\sdif|} \choose |\sdif| }$ is replaced by ${ {p-k+|\sdif|} \choose |\sdif| } - \sum_{d=0}^{\dmax}{{p-k} \choose d}{|\sdif| \choose d}$.
\section{Conclusion} \label{sec:CONCLUSION}
Taking an approach motivated by thresholding techniques in channel coding, we have presented a framework for developing necessary and sufficient conditions on the number of measurements for exact and partial support recovery with probabilistic models. We have provided several new results for the linear, 1-bit, and group testing models, as well as general discrete observation models. In several cases, we have provided exact asymptotic thresholds on the number of measurements with strong converse results.
There are several possible directions for future research. While we have focused on i.i.d.~measurement matrices, it would be of significant interest to consider other types of random matrices, and to present converse results that hold for arbitrary measurement matrices, subject to suitable constraints such as power constraints. We provided some work in these directions for specific models in \cite{Sca16b,Sca16c}.
One could also attempt to move from standard sparsity models to structured sparsity models \cite{Bar10}, and from probabilistic guarantees with random $\beta$ to minimax guarantees. There are several additional non-linear models that our general results could be applied to, such as the Poisson and gamma models. Finally, it may be interesting to apply similar analysis techniques to other statistical problems beyond support recovery.
\appendices
\section{Concentration Inequalities} \label{sec:CONCENTRATION}
In order to apply our general bounds to specific models, we use concentration inequalities to obtain expressions for $\psi_{\ell}$ and $\psi'_{\ell}$ in \eqref{eq:psi_ach} and \eqref{eq:psi_conv}, seeking to make the corresponding terms in \eqref{eq:nonasymp_ach} and \eqref{eq:nonasymp_conv} vanish. Here we present two general inequalities that will be used throughout Section \ref{sec:EXAMPLES}.
\begin{prop} \label{prop:chebyshev}
For general observation models, we have for all $(\sdif,\seq,b_s)$ and $\delta>0$ that
\begin{multline}
\PP\Big[ \big|\imath^n(\Xv_{\sdif}; \Yv | \Xv_{\seq}, b_s) - nI_{\sdif,\seq}(b_s) \big| \ge n\delta \,\Big|\, \beta_s = b_s \Big] \\ \le \frac{V_{\sdif,\seq}(b_s)}{\delta^2 n}, \label{eq:conc_linear1}
\end{multline}
where $V_{\sdif,\seq}(b_s)$ is defined in \eqref{eq:Vb}. Moreover, if the observations lie in a finite set $\Yc \subset \RR$ with probability one, then the following holds for all $(\sdif,\seq,b_s)$ and $\delta>0$:
\begin{equation}
V_{\sdif,\seq}(b_s) \le |\Yc|\Big( \frac{4}{e} \Big) ^2. \label{eq:uniformV}
\end{equation}
\end{prop}
Before providing the proof, we state the following generalization of \eqref{eq:uniformV} to higher-order moments.
\begin{prop} \label{prop:gen_discrete}
If the observations lie in a finite set $\Yc \subset \RR$ with probability one, then the following holds for all $(\sdif,\seq,b_s)$ and $\delta > 0$:
\begin{multline}
\PP\bigg[ \big|\imath^n(\Xv_{\sdif}; \Yv | \Xv_{\seq}, b_s) - nI_{\sdif,\seq}(b_s) \big| \ge n\delta \,\Big|\, \beta_s = b_s \bigg] \\ \le 2\exp\bigg(- \frac{\delta^2 n}{2(8|\Yc|+2\delta)} \bigg). \label{eq:conc_gen_disc}
\end{multline}
\end{prop}
In the remainder of this appendix, we prove these propositions. Equation \eqref{eq:conc_linear1} follows from Chebyshev's inequality, so we focus our attention on \eqref{eq:uniformV}--\eqref{eq:conc_gen_disc}. We make use of the following form of Bernstein's inequality \cite[Sec.~2.8]{Bou13}.
\begin{lem} \label{lem:bernstein}
Let $W_1,\dotsc,W_n$ be independent real-valued random variables such that
\begin{align}
\sum_{i=1}^n \EE[W_i^2] &\le \tau \\
\sum_{i=1}^n \EE[ |W_i|^q ] &\le \frac{q!}{2}\tau c^{q-2} \qquad (q \ge 3)
\end{align}
for some $\tau,c>0$. Then
\begin{equation}
\PP\bigg[ \sum_{i=1}^n \big( W_i - \EE[W_i] \big) \ge t \bigg] \le \exp\bigg( \frac{t^2}{2(\tau+ct)} \bigg) \label{eq:bernstein}
\end{equation}
for all $t>0$.
\end{lem}
To bound the moments of $\imath$, we follow the arguments of \cite[Rmk.~3.1.1]{Han03} and \cite[App.~D]{Tan14a}. Recall the definition of the information density in \eqref{eq:idens_b}. For any $q \ge 2$, we have from Minkowski's inequality that
\begin{multline}
\EE\big[ |\imath(X_{\sdif}; Y | X_{\seq}, b_s)|^q\big]^{1/q} \\ \le \EE\bigg[ \Big(\log\frac{1}{ P_{Y|X_{\sdif}X_{\seq}\beta_s}(Y|X_{\sdif},X_{\seq},b_s) }\Big)^q\bigg]^{1/q} \\ + \EE\bigg[ \Big(\log\frac{ 1 }{ P_{Y|X_{\seq}\beta_s}(Y|X_{\seq},b_s) }\Big)^q\bigg]^{1/q}, \label{eq:disc_unif1}
\end{multline}
where here and subsequently we implicitly condition on $\beta_s = b_s$. For any given $(x_{\sdif},x_{\seq})$, the remaining averaging over $Y$ in the first term has the form
\begin{multline}
\sum_{y} P_{Y|X_{\sdif}X_{\seq}\beta_s}(y|x_{\sdif},x_{\seq},b_s) \\ \times\Big(\log\frac{1}{ P_{Y|X_{\sdif}X_{\seq}\beta_s}(y|x_{\sdif},x_{\seq},b_s) }\Big)^q,
\end{multline}
and is thus upper bounded by $|\Yc| \big( \frac{q}{e} \big)^{1/q}$, since the function $f(z) = z \log^{q}\frac{1}{z}$ has a maximum value of $\big( \frac{q}{e} \big)^{1/q}$ for $z\in[0,1]$. Handling the second term in \eqref{eq:disc_unif1} similarly, we obtain
\begin{equation}
\EE\big[ \big|\imath(X_{\sdif}; Y | X_{\seq}, b_s)\big|^q\big]^{1/q} \le 2\Big( |\Yc| \Big( \frac{q}{e} \Big)^{q} \Big)^{1/q},
\end{equation}
or equivalently
\begin{align}
\EE\big[ \big|\imath(X_{\sdif}; Y | X_{\seq}, b_s)\big|^q\big]
& \le \Big( \frac{q}{e} \Big)^{q} 4|\Yc| 2^{q-2} \label{eq:disc_unif4} \\
& \le \frac{q!}{2} 8|\Yc| 2^{q-2}, \label{eq:disc_unif5}
\end{align}
where \eqref{eq:disc_unif5} follows since $\big( \frac{q}{e} \big)^{q} \le q!$.
We obtain \eqref{eq:uniformV} by setting $q=2$ in \eqref{eq:disc_unif4}. Furthermore, we obtain Proposition \ref{prop:gen_discrete} using Lemma \ref{lem:bernstein} with $c = 2$, $\tau = n\cdot8|\Yc|$, and $t=\delta n$.
\section{Proofs of Auxiliary Results for the Linear Model} \label{sec:LINEAR_PROOFS}
\subsection{Proof of Proposition \ref{prop:conc_linear2}}
We again use Lemma \ref{lem:bernstein}, and we thus seek suitable values for $\tau$ and $c$. Throughout the proof, we consider the random variables $(X_{\sdif},X_{\seq},Y)$ distributed according to \eqref{eq:distr_sl}, implicitly conditioning on $\beta_s=b_s$. From \eqref{eq:linear_model}, we have $Z = Y - \sum_{i \in s}X_i b_i$, and a direct calculation gives
\begin{gather}
P_{Y|X_{\sdif}X_{\seq}\beta_s}(Y|X_{\sdif},X_{\seq},b_s) = \phi(Z; 0, \sigma^2)
\end{gather}
\vspace*{-3ex}
\begin{multline}
P_{Y|X_{\seq}\beta_s}(Y|X_{\seq},b_s) \\ = \phi\Big( \sum_{i\in\sdif} X_i b_i + Z; 0, \sigma^2 + \sum_{i\in\sdif}b_i^2 \Big),
\end{multline}
where $\phi(\cdot;\mu,\sigma^2)$ is the $N(\mu,\sigma^2)$ density function. Substituting these into \eqref{eq:idens_b} gives
\begin{multline}
\imath(X_{\sdif}; Y | X_{\seq}, b_s) = I_{\sdif,\seq}(b_s) - \frac{Z^2}{2\sigma^2} \\ + \frac{1}{2\big(\sigma^2 + \sum_{i\in\sdif}b_i^2 \big)} \bigg( \sum_{i\in\sdif} X_i b_i + Z \bigg)^2, \label{eq:n_i_dens}
\end{multline}
where $I_{\sdif,\seq}(b_s)$ is given in \eqref{eq:I_linear}.
The mean of \eqref{eq:n_i_dens} is $I_{\sdif,\seq}(b_s)$, and we will apply Lemma \ref{lem:bernstein} with $W_i$ corresponding to the sum of the second and third terms on the right-hand side. We can write these in terms of independent $N(0,1)$ random variables (denoted by $\Zhat_1$ and $\Zhat_2$) as follows:
\begin{align}
W &= - \frac{\Zhat_1^2}{2} + \frac{1}{2(\sigma^2 + \siglv^2)} \big( \sigma \Zhat_1 + \siglv \Zhat_2 \big)^2 \label{eq:linear4} \\
&= \frac{\siglv^2}{2(\sigma^2 + \siglv^2)}\big( \Zhat_2^2 - \Zhat_1^2 \big) + \frac{\sigma\siglv}{\sigma^2 + \siglv^2} \Zhat_1\Zhat_2, \label{eq:linear4a}
\end{align}
where we have used the definitions in the proposition statement, and \eqref{eq:linear4a} follows from simple manipulations. Defining $\Zhatmax = \max\{ |\Zhat_1|,|\Zhat_2| \}$, we have the following with probability one:
\begin{align}
|W| &\le \frac{\siglv^2}{2(\sigma^2 + \siglv^2)} 2\Zhatmax^2 + \frac{\sigma\siglv}{\sigma^2 + \siglv^2} \Zhatmax^2 \label{eq:linear5} \\
&= \frac{\siglv(\sigma + \siglv)}{\sigma^2 + \siglv^2} \Zhatmax^2. \label{eq:linear6}
\end{align}
Since $\EE[\Zhatmax^4] \le \EE[\Zhat_1^4 + \Zhat_2^4] = 6$, we obtain
\begin{align}
\EE[W^2] &\le 6 \bigg(\frac{\siglv(\sigma + \siglv)}{\sigma^2 + \siglv^2}\bigg)^2. \label{eq:linear7}
\end{align}
Similarly, we can bound the higher moments as follows:
\begin{align}
\EE[|W|^q] &\le \bigg(\frac{\siglv(\sigma + \siglv)}{\sigma^2 + \siglv^2}\bigg)^{q} \EE[\Zhat_1^{2q} + \Zhat_2^{2q}] \label{eq:linear9} \\
&\le \bigg(\frac{2\siglv(\sigma + \siglv)}{\sigma^2 + \siglv^2}\bigg)^{q} \frac{2}{\sqrt{\pi}} \Gamma\Big(q + \frac{1}{2}\Big) \label{eq:linear10} \\
&\le 2 \cdot \bigg(\frac{2\siglv(\sigma + \siglv)}{\sigma^2 + \siglv^2}\bigg)^{q} \cdot q! \label{eq:linear11},
\end{align}
where \eqref{eq:linear10} follows by the same argument as \eqref{eq:linear6} and the fact that the $2q$-th moment of an $N(0,1)$ random variable is $\frac{2^q}{\sqrt{\pi}} \Gamma\big(q + \frac{1}{2}\big)$, and \eqref{eq:linear11} follows since $\Gamma\big(q + \frac{1}{2}\big) \le \sqrt{\pi} q!$.
Combining \eqref{eq:linear7} and \eqref{eq:linear11}, we see that the random variables $W_i = \imath(X_{\sdif}^{(i)}; Y^{(i)} | X^{(i)}_{\seq}, b_s) - I_{\sdif,\seq}(b_s)$ satisfy the conditions of Lemma \ref{lem:bernstein} with $\tau = n\cdot4\alpha_{\sdif}^2$ and $c=\alpha_{\sdif}$ (see \eqref{eq:alpha_linear}). We thus obtain the desired result from \eqref{eq:bernstein} by identifying $t = \delta n$.
\subsection{Proof of Proposition \ref{prop:boundMI}}
Since $\Yv = \Xv_s\beta_s + \Zv$, we have
\begin{align}
I_0 = I(\beta_s;\Yv|\Xv_s) &= H(\Yv|\Xv_s) - H(\Yv|\Xv_s,\beta_s) \\
&= H(\Xv_s\beta_s + \Zv|\Xv_s) - H(\Zv). \label{eq:pf_bmi2}
\end{align}
From \cite[Ch.~9]{Cov01}, we have $H(\Zv) = \frac{n}{2}\log(2\pi e \sigma^2)$ and $H(\Xv_s\beta_s + \Zv|\Xv_s = \xv_s) = \frac{1}{2}\log\big( (2\pi e)^n \det( \sigma^2 \Iv_n + \sigbeta^2 \xv_s \xv_s^T ) \big) $, where $\Iv_n$ is the $n \times n$ identity matrix. Averaging the latter over $\Xv_s$ and substituting these into \eqref{eq:pf_bmi2} gives
\begin{align}
I_0 &= \frac{1}{2} \EE\Big[ \log \det \Big( \Iv_n + \frac{\sigbeta^2}{\sigma^2} \Xv_s \Xv_s^T \Big) \Big] \\
&= \frac{1}{2} \EE\Big[ \log \det \Big( \Iv_k + \frac{\sigbeta^2}{\sigma^2} \Xv_s^T \Xv_s \Big) \Big] \label{eq:pf_bmi4} \\
&= \frac{1}{2} \sum_{i=1}^k \EE\Big[ \log \Big( 1+ \frac{\sigbeta^2}{\sigma^2} \lambda_i (\Xv_s^T \Xv_s) \Big) \Big] \label{eq:pf_bmi5} \\
&\le \frac{k}{2} \log \Big( 1 + \frac{n\sigbeta^2}{\sigma^2} \Big), \label{eq:pf_bmi6}
\end{align}
where \eqref{eq:pf_bmi4} follows from the identity $\det(\Iv + \Av\Bv) = \det(\Iv + \Bv\Av)$, \eqref{eq:pf_bmi5} follows by writing the determinant as a product of eigenvalues (denoted by $\lambda_i(\cdot)$), and \eqref{eq:pf_bmi6} follows from Jensen's inequality and the following calculation:
\begin{equation}
\frac{1}{k}\EE\Big[ \sum_{i=1}^k \lambda_i (\Xv_s^T \Xv_s) \Big] = \frac{1}{k} \EE[ \Tr( \Xv_s^T \Xv_s )] = \EE[ \Xv_1^T \Xv_1 ] = n.
\end{equation}
This concludes the proof of \eqref{eq:bmi_I0}.
We now turn to the bounding of the variance. Again using the fact that $\Yv = \Xv_s\beta_s + \Zv$, we have
\begin{align}
&\log \frac{P_{\Yv|\Xv_s,\beta_s}(\Yv|\Xv_{s},\beta_s)}{P_{\Yv|\Xv_{s}}(\Yv|\Xv_{s})} \nonumber \\
&= \log\frac{P_{\Zv}(\Zv)}{P_{\Yv|\Xv_{s}}(\Xv_s\beta_s + \Zv|\Xv_s)} \\
&= I_0 - \frac{1}{2\sigma^2}\Zv^T\Zv \nonumber \\
&~~ + \frac{1}{2}(\Xv_s\beta_s + \Zv)^T \big( \sigma^2 \Iv + \sigbeta^2 \Xv_s\Xv_s^T \big)^{-1} (\Xv_s\beta_s + \Zv), \label{eq:pf_bmi8}
\end{align}
where $P_{\Zv}$ is the density of $\Zv$, and \eqref{eq:pf_bmi8} follows by a direct substitution of the densities $P_{\Zv} \sim N(\bzero,\sigma^2\Iv)$ and $P_{\Yv|\Xv_S}(\cdot|\xv_s) \sim N(\bzero, \sigma^2\Iv + \sigbeta^2\xv_s\xv_s^T)$, where $\bzero$ is the zero vector. Observe now that $\frac{1}{\sigma^2} \Zv^T \Zv$ is a sum of $n$ independent $\chi^2$ random variables with one degree of freedom (each having a variance of $2$), and hence, the second term in \eqref{eq:pf_bmi8} has a variance of $\frac{n}{2}$. Moreover, by writing $\Mv^{-1} = (\Mv^{-\frac{1}{2}})^T \Mv^{-\frac{1}{2}}$ for the symmetric positive definite matrix $\Mv = \sigma^2 \Iv + \sigbeta^2 \Xv_s\Xv_s^T$, where $(\cdot)^{-\frac{1}{2}}$ denotes the positive definite matrix square root of the inverse, we find that the final term in \eqref{eq:pf_bmi8} is distributed as a sum of $\chi^2$ variables when conditioned on any value of $\Xv_s$, and hence, the same is true unconditionally. We therefore again obtain a variance of $\frac{n}{2}$, and \eqref{eq:bmi_V0} follows using the identity $\var[A+B] \le \var[A] + \var[B] + 2\max\{\var[A], \var[B]\}$.
\section{Proofs of Auxiliary Results for the 1-bit Model} \label{sec:1BIT_PROOFS}
We first write down the relevant probability distributions and information densities conditioned on a fixed value $b_s$ of $\beta_s$. Under the model $Y = \sign\big(\sum_{i\in s}X_i b_i + Z\big)$ with $X_i \sim N(0,1)$ and $Z \sim N(0,\sigma^2)$, we have
\begin{align}
P_{Y|X_s\beta_s}(1|x_s,b_s)
&= \PP\bigg[ Z \ge - \sum_{i\in s}x_i b_i\bigg] \\
&= Q\bigg( - \frac{1}{\sigma} \sum_{i\in s}x_i b_i \bigg). \label{eq:1bit_p1}
\end{align}
Similarly, for any partition of $s$ into $(\sdif,\seq)$, we can write $Y = \sign\big(\sum_{i\in \seq}X_i b_i + \sum_{i\in \sdif}X_i b_i + Z\big)$ and use the same steps to conclude that
\begin{align}
P_{Y|X_{\seq}\beta_s}(1|x_{\seq},b_s) = Q\bigg( \frac{ -\sum_{i\in \seq}x_i b_i }{\sqrt{\sigma^2 + \sum_{i\in \sdif}b_i^2}} \bigg). \label{eq:1bit_p2}
\end{align}
The corresponding probabilities for $y=0$ are one minus these expressions, which amounts to multiplying the argument to the Q-function by $-1$. Substitution into \eqref{eq:idens_b} gives
\begin{equation}
\imath(x_{\sdif}; y | x_{\seq}, b_s) = \log \frac{ Q\Big( -y \frac{1}{\sigma} \sum_{i\in s}x_i b_i\Big) }{ Q\Big( \frac{ -y \sum_{i\in \seq}x_i b_i }{\sqrt{\sigma^2 + \sum_{i\in \sdif}b_i^2}} \Big) } \label{eq:1bit_i_dens}
\end{equation}
for $y\in\{-1,1\}$.
Throughout this appendix, we will use the fact that the first two derivatives of the function
\begin{equation}
f(x) := H_2(Q(x)) \label{eq:1bit_f}
\end{equation}
are given by
\begin{align}
f'(x) &= \log\frac{1-Q(x)}{Q(x)} \frac{-1}{\sqrt{2\pi}} e^{-\frac{x^2}{2}} \label{eq:1bit_f'} \\
f''(x) &= -\frac{1}{2\pi} e^{-x^2}\frac{1}{Q(x)(1-Q(x))} \nonumber \\
& \hspace*{13ex} + \log\frac{1-Q(x)}{Q(x)} \frac{x}{\sqrt{2\pi}} e^{\frac{-x^2}{2}}. \label{eq:1bit_f''}
\end{align}
\subsection{Proof of Proposition \ref{prop:1bit_moments} Part (i)}
Recalling that the coefficients $X_i$ ($i\in s$) are i.i.d. on $N(0,1)$, we directly obtain from \eqref{eq:1bit_p1} that
\begin{align}
H(Y|X_s,\beta_s=b_s)
&= \EE\bigg[ H_2\bigg( Q\bigg( \frac{1}{\sigma} \sum_{i\in s}X_i b_i \bigg) \bigg)\bigg] \\
&= \EE\bigg[ H_2\bigg( Q\bigg( W \sqrt{\frac{1}{\sigma^2}\sum_{i \in s} b_i^2} \bigg) \bigg) \bigg],
\end{align}
where $W \sim N(0,1)$. By evaluating $H(Y|X_{\seq},\beta_s=b_s)$ similarly using \eqref{eq:1bit_p2} and taking the difference between the two, we obtain \eqref{eq:1bit_Ilv}.
\subsection{Proof of Proposition \ref{prop:1bit_moments} Part (ii)}
We obtain from \eqref{eq:1bit_f'}--\eqref{eq:1bit_f''} that $f'(0) = 0$ and $f''(0) = -\frac{2}{\pi}$. By performing further differentiations, one can also verify that $f^{(3)}(0) = 0$, and that $|f^{(4)}(x)|$ is uniformly upper bounded by $f^{(4)}(0) = \frac{8(\pi - 1)}{\pi^2}$. We thus obtain via a fourth-order Taylor expansion that
\begin{multline}
\log 2 - \frac{1}{\pi}x^2 - \frac{4(\pi - 1)}{3\pi^2}x^4 \le H_2(Q(x)) \\ \le \log 2 - \frac{1}{\pi}x^2 + \frac{4(\pi - 1)}{3\pi^2}x^4 \label{eq:1bit_pfii_1}
\end{multline}
for all $x \in \RR$. Substituting \eqref{eq:1bit_pfii_1} into \eqref{eq:1bit_Ilv} and noting that the fourth moments of the arguments to $H_2(Q(\cdot))$ therein decay to zero strictly faster than the second moments (by the assumptions on $k$, $\bmin$ and $\bmax$), we obtain
\begin{equation}
I_{\sdif,\seq}(b_s) = \frac{1}{\pi}\bigg( \frac{1}{\sigma^2}\sum_{i \in s} b_i^2 - \frac{\sum_{i\in\seq}b_i^2}{\sigma^2 + \sum_{i\in\sdif}b_i^2} \bigg)(1+o(1)).
\end{equation}
Again using the assumptions on $k$, $\bmin$ and $\bmax$, we observe that the denominator is dominated by the term $\sigma^2$, thus yielding \eqref{eq:1bit_Ilv_asymp}.
\subsection{Proof of Proposition \ref{prop:1bit_moments} Part (iii)}
In this part, we have assumed that the values $\{b_i\}$ take a common value $b_0$. Since $\sigma^2=\Theta(1)$, we may set $\sigma^2=1$ without loss of generality; the implied constant can be factored into $b_0$. In this case, \eqref{eq:1bit_Ilv} with $\ell=1$ simplifies to
\begin{multline}
I_1 = \EE\bigg[ H_2\bigg( Q\bigg( W \sqrt{ \frac{(k-1)b_0^2}{1 + b_0^2} } \bigg) \bigg) \\ - H_2\bigg( Q\Big( W \sqrt{kb_0^2} \Big) \bigg) \bigg]. \label{eq:1bit_Ilv_simp}
\end{multline}
By the assumptions $k=\Theta(p)$ and $b_0^2 = \Theta\big( \frac{\log p}{p} \big)$, it is easily verified by a Taylor expansion of the function $f(z) = \frac{1}{\sqrt{1+z}}$ as $z\to0$ that $\sqrt{\frac{(k-1)b_0^2}{1 + b_0^2}} = \sqrt{kb_0^2}\big(1-\frac{b_0^2}{2} + o(b_0^2)\big)$. For convenience, we write this identity as
\begin{equation}
\sqrt{\frac{(k-1)b_0^2}{1 + b_0^2}} = \sqrt{kb_0^2}\big(1 - \zeta b_0^2\big), \label{eq:sqrt_Taylor}
\end{equation}
where $\zeta$ is a constant depending on $p$ such that $\zeta \to \frac{1}{2}$. Substituting \eqref{eq:sqrt_Taylor} into \eqref{eq:1bit_Ilv_simp}, we obtain
\begin{multline}
I_1 = \EE\bigg[ H_2\bigg( Q\Big( W \sqrt{kb_0^2}\big(1 - \zeta b_0^2\big) \Big) \bigg) \\ - H_2\bigg( Q\Big( W \sqrt{kb_0^2} \Big) \bigg) \bigg]. \label{eq:1bit_Ilv_simp2}
\end{multline}
The next step is to Taylor expand the function $f(x) = H_2(Q(x))$. For any $x$ and $\delta>0$, we have
\begin{equation}
f(x - \delta) = f(x) + \frac{\delta}{\sqrt{2\pi}} \log\frac{1-Q(x)}{Q(x)}e^{-\frac{x^2}{2}} + \frac{\delta^2}{2}f''(x - \delta_0) \label{eq:1bit_taylor}
\end{equation}
for some $\delta_0 \in [0,\delta]$, where the middle term follows from \eqref{eq:1bit_f'}. Next, we claim that $f''$ in \eqref{eq:1bit_f''} is bounded as follows:
\begin{equation}
|f''(x)| \le \frac{2}{\sqrt{2\pi}} (1+|x|) e^{\frac{-x^2}{2}} + \frac{|x|^3}{\sqrt{2\pi}} e^{\frac{-x^2}{2}}. \label{eq:1bit_f''_bound}
\end{equation}
In the case that $x \ge 0$, this is seen by applying $Q(x) \ge \frac{1}{\sqrt{2\pi}(1+x)}e^{-\frac{x^2}{2}}$ and $1-Q(x) \ge \frac{1}{2}$ to obtain the first term, and applying $Q(x) \le e^{-x^2}$ (and hence $\log\frac{1-Q(x)}{Q(x)} = \log\big(\frac{1}{Q(x)} - 1\big) \le x^2$) to obtain the second term (e.g., see \cite{Fan12} for bounds on the Q-function). The case $x < 0$ follows since \eqref{eq:1bit_f''} is symmetric about zero.
Substituting \eqref{eq:1bit_taylor} into \eqref{eq:1bit_Ilv_simp2} with the identifications $x=W\sqrt{kb_0^2}$ and $\delta = W\sqrt{kb_0^2}\zeta b_0^2$, we can write
\begin{equation}
T_1 - T_2 - T_3 \le I_1 \le T_1 + T_2 + T_3,
\end{equation}
where
\begin{align}
T_1 &:= \zeta b_0^2 \EE\bigg[ \frac{W \sqrt{kb_0^2}}{\sqrt{2\pi}} \log\frac{1-Q(W \sqrt{kb_0})}{Q(W \sqrt{kb_0})}e^{-\frac{W^2 kb_0^2}{2}} \bigg] \\
T_2 &:= (\zeta b_0^2)^2 \EE\bigg[ \frac{W^2 kb_0^2}{\sqrt{2\pi}} \big(1 + |W|\sqrt{kb_0^2}\big) e^{-\frac{W^2 kb_0^2}{2}\big(1 - \zeta b_0^2\big)^2} \bigg] \\
T_3 &:= (\zeta b_0^2)^2 \EE\bigg[ \frac{W^2 kb_0^2}{2\sqrt{2\pi}} |W|^3 (kb_0^2)^{3/2} e^{-\frac{W^2 kb_0^2}{2}\big(1 - \zeta b_0^2\big)^2} \bigg],
\end{align}
and where for $T_2$ and $T_3$ we used the fact that $\delta_0 \in [0,\delta]$ in \eqref{eq:1bit_taylor} to upper bound the corresponding terms by the value at $\delta_0=0$ or $\delta_0 = \delta$.
We will complete the proof by showing that $T_1$ behaves as \eqref{eq:1bit_I1_exact} (with $\sigma^2 = 1$), and that $T_2$ and $T_3$ behave as $o\big( \frac{\sqrt{\log p}}{p} \big)$. Letting $\phi(\cdot)$ denote the standard normal PDF, we have
\begin{align}
T_1 &= \frac{\zeta b_0^2}{\sqrt{2\pi}} \int_{-\infty}^{\infty} \phi(w) w \sqrt{kb_0^2} \log\frac{1-Q(w\sqrt{kb_0^2})}{Q(w\sqrt{kb_0^2}} \nonumber \\
& \hspace*{30ex}\times e^{-\frac{w^2 kb_0^2}{2}} dw \label{eq:1bit_T1_1} \\
&= \frac{\zeta b_0^2}{\sqrt{2\pi}} \int_{-\infty}^{\infty} \phi\Big( \frac{t}{\sqrt{kb_0^2}} \Big) t \log\frac{1-Q(t)}{Q(t)} e^{-\frac{t^2}{2}} \frac{1}{\sqrt{kb_0^2}} dt \label{eq:1bit_T1_2} \\
&= \frac{\zeta b_0^2}{\sqrt{2\pi kb_0^2}} \int_{-\infty}^{\infty} \frac{1}{\sqrt{2\pi}} e^{-\frac{t^2}{2}(1 + \frac{1}{kb_0^2})} t \log\frac{1-Q(t)}{Q(t)} dt \label{eq:1bit_T1_3} \\
&= \frac{\zeta b_0^2}{\sqrt{2\pi kb_0^2}} \frac{1}{\sqrt{1 + \frac{1}{kb_0^2}}} \int_{-\infty}^{\infty} \frac{1}{\sqrt{2\pi(1 + \frac{1}{kb_0^2})^{-1}}} \nonumber \\
& \hspace*{15ex} \times e^{-\frac{t^2}{2}(1 + \frac{1}{kb_0^2})} t \log\frac{1-Q(t)}{Q(t)} dt \label{eq:1bit_T1_4} \\
&= \frac{1}{2} \frac{b_0^2}{\sqrt{2\pi kb_0^2}} \EE\Big[ W \log\frac{1-Q(W)}{Q(W)} \Big] (1+o(1)), \label{eq:1bit_T1_5}
\end{align}
where \eqref{eq:1bit_T1_2} follows by a change of variable of the form $t = w\sqrt{k b_0^2}$, \eqref{eq:1bit_T1_3} follows from the definition of $\phi$, and \eqref{eq:1bit_T1_5} follows since $\zeta \to \frac{1}{2}$, and since the integral in \eqref{eq:1bit_T1_4} is the average of $t\log\frac{1-Q(t)}{Q(t)} \openone\{t \ge 0\} $ over an $N(0,(1 + \frac{1}{kb_0^2})^{-1})$ random variable; since $kb_0^2 \to \infty$, this converges to the corresponding average over $W \sim N(0,1)$, which is easily verified to be finite.
The terms $T_2$ and $T_3$ are handled similarly to $T_1$, so we only briefly comment on the analysis of $T_3$. By the same arguments as those leading to \eqref{eq:1bit_T1_3}, we obtain
\begin{equation}
T_3 = \frac{(\zeta b_0^2)^2}{2\sqrt{2\pi kb_0^2}} \int_{-\infty}^{\infty} \frac{1}{\sqrt{2\pi}} e^{-\frac{t^2}{2}\big( (1 - \zeta b_0^2)^2 + \frac{1}{kb_0^2} \big)} |t|^5 dt.
\end{equation}
The integral is once again $\Theta(1)$, and thus $T_3 = \Theta\Big( \frac{b_0^4}{\sqrt{kb_0^2}} \Big)$, which decays to zero strictly faster than \eqref{eq:1bit_T1_5}.
\subsection{Proof of Proposition \ref{prop:1bit_moments} Part (iv)}
We again assume without loss of generality that $\sigma^2 = 1$. Defining $\Weq := \sum_{i\in\seq} X_i b_i$ and $\Wdif := \sum_{i\in\sdif} X_i b_i$, it follows from \eqref{eq:1bit_i_dens} that
\begin{equation}
\imath(X_{\sdif}; Y | X_{\seq}, b_s) = \log \frac{ Q\big( -Y(\Wdif+\Weq)\big) }{ Q\big( -Y \tau \Weq \big)}, \label{eq:1bit_iii_0}
\end{equation}
where $\tau := \frac{1}{1+\sum_{i\in\sdif}b_i^2}$, and we implicitly condition on $\beta_s = b_s$. Using \eqref{eq:1bit_p1} and the fact that the variance is upper bounded by the second moment, we have
\begin{align}
& V_{\sdif,\seq}(b_s) \nonumber \\
&\le \EE\bigg[ Q\big( -(\Wdif+\Weq) \big) \bigg(\log \frac{ Q\big( -(\Wdif+\Weq)\big) }{ Q\big( -\tau \Weq \big)}\bigg)^2 \nonumber \\
& \qquad + Q\big( \Wdif+\Weq \big) \bigg(\log \frac{ Q\big( \Wdif+\Weq\big) }{ Q\big( \tau \Weq \big)}\bigg)^2 \bigg] \label{eq:1bit_iii_1} \\
&= 2 \EE\bigg[ Q\big( \Wdif+\Weq \big) \bigg(\log \frac{ Q\big( \Wdif+\Weq\big) }{ Q\big( \tau \Weq \big)}\bigg)^2 \bigg] \label{eq:1bit_iii_2},
\end{align}
where \eqref{eq:1bit_iii_2} follows since the distributions of $\Wdif$ and $\Weq$ are symmetric about zero, and the two are independent.
The function $g(x) := - \log Q(x)$ is convex, and hence it lies above any given tangent vector. This implies that
\begin{align}
|g(x_1) - g(x_2)| &\le \max\big\{ |g'(x_1)|, |g'(x_2)| \big\} |x_1 - x_2| \\
&\le \big( |g'(x_1)| + |g'(x_2)| \big) |x_1 - x_2|, \label{eq:1bit_iii_4}
\end{align}
where $g'(x) = -\frac{\phi(x)}{Q(x)}$ is the derivative of $g$. Writing the logarithm of the ratio in \eqref{eq:1bit_iii_2} as a difference of logarithms and applying \eqref{eq:1bit_iii_4}, we obtain
\begin{equation}
V_{\sdif,\seq}(b_s) \le 2(T_1 + T_2),
\end{equation}
where, overloading the notation from part (iii), we define
\begin{align}
& T_1 := \iint \fdif(\wdif)\feq(\weq) Q(\wdif+\weq) \nonumber \\
& \times \bigg( \frac{\phi(\wdif + \weq)}{Q(\wdif + \weq)} \bigg)^2 \big( |\wdif| + (1-\tau)|\weq| \big)^2 d\wdif d\weq \label{eq:1bit_iii_T1} \\
& T_2 := \iint \fdif(\wdif)\feq(\weq) Q(\wdif+\weq) \nonumber \\
& \times\bigg( \frac{\phi(\tau\weq)}{Q(\tau\weq)} \bigg)^2 \big( |\wdif| + (1-\tau)|\weq| \big)^2 d\wdif d\weq \label{eq:1bit_iii_T2}
\end{align}
with $\fdif$ and $\feq$ denoting the densities of $\Wdif$ and $\Weq$. The function $Q(x)\big(\frac{\phi(x)}{Q(x)}\big)^2$ lies between $0$ and $\frac{1}{2}$, and hence $T_1 \le \frac{1}{2}\EE\big[ \big( |\Wdif| + (1-\tau)|\Weq| \big)^2 \big]$, yielding
\begin{equation}
T_1 = O\big( \EE[\Wdif]^2 + (1-\tau)^2 \EE[\Weq]^2 \big).
\label{eq:1bit_T1bound}
\end{equation}
We will further simplify this expression below, but we first bound $T_2$, which requires more effort.
We split the integral over $\RR^2$ in \eqref{eq:1bit_iii_T2} according to whether $|\wdif| \le \frac{1}{2}|\weq|$ or $|\wdif| > \frac{1}{2}|\weq|$; the resulting expressions are denoted by $T_{1,1}$ and $T_{1,2}$ respectively. In each case, we use the following standard bounds on the Q-function (e.g., see \cite{Fan12}):
\begin{align}
\frac{\phi(\tau\weq)}{Q(\tau\weq)}
&\le \begin{cases} 1+\tau\weq & \weq \ge 0 \\ 1 & \weq < 0 \end{cases} \label{eq:1bit_cases1} \\
Q(\wdif+\weq)
&\le \begin{cases} \frac{1}{2}e^{-\frac{(\wdif + \weq)^2}{2}} & \wdif + \weq \ge 0 \\ 1 & \wdif + \weq < 0. \end{cases} \label{eq:1bit_cases2}
\end{align}
To bound $T_{1,1}$, we note that the condition $|\wdif| \le \frac{1}{2}|\weq|$ implies that $\sign(\wdif + \weq) = \sign(\weq)$, and hence only two of the four combinations of the cases in \eqref{eq:1bit_cases1}--\eqref{eq:1bit_cases2} can occur. When $\weq < 0$, we can use the second of each of these cases to upper bound the integrand in \eqref{eq:1bit_iii_T2} by $\fdif(\wdif)\feq(\weq) \big( |\wdif| + (1-\tau)|\weq| \big)^2$. On the other hand, when $\weq \ge 0$ we can use the first of each of the cases to upper bound the integrand by
\begin{multline}
\fdif(\wdif)\feq(\weq) \frac{1}{2}e^{-\frac{(\wdif + \weq)^2}{2}} \\ \times \big(1+\tau|\weq|\big)^2 \big( |\wdif| + (1-\tau)|\weq| \big)^2.
\end{multline}
Again using the condition $|\wdif| \le \frac{1}{2}|\weq|$, we find that $e^{-\frac{(\wdif + \weq)^2}{2}} \le e^{-\frac{1}{8}\weq^2}$. Since $\tau \le 1$ by its definition following \eqref{eq:1bit_iii_0}, it follows that $e^{-\frac{(\wdif + \weq)^2}{2}} \big(1+\tau\weq\big)^2$ is upper bounded by a universal constant, and we are again left only with $\fdif(\wdif)\feq(\weq) \big( |\wdif| + (1-\tau)|\weq| \big)^2$. Combining the two cases, we conclude that
\begin{equation}
T_{2,1} = O\big( \EE[\Wdif^2] + (1-\tau)^2 \EE[\Weq^2] \big).
\label{eq:1bit_T21bound}
\end{equation}
To upper bound $T_{2,2}$, we upper bound the integrand in \eqref{eq:1bit_iii_T2} by
\begin{align}
&\fdif(\wdif)\feq(\weq) \big(1+\tau|\weq|\big)^2 \big( |\wdif| + (1-\tau)|\weq| \big)^2 \label{eq:1bit_iii_end1} \\
&\qquad \le \fdif(\wdif)\feq(\weq) \big(1+2|\wdif|\big)^2 \big( 3|\wdif| \big)^2, \label{eq:1bit_iii_end2}
\end{align}
where \eqref{eq:1bit_iii_end1} follows by taking the higher of the two cases in both \eqref{eq:1bit_cases1} and \eqref{eq:1bit_cases2}, and \eqref{eq:1bit_iii_end2} follows since $|\wdif| > \frac{1}{2}|\weq|$ and $\tau \in [0,1]$. It follows that
\begin{equation}
T_{2,2} = O\big( \EE[\Wdif^2] + \EE[\Wdif^4] \big).
\label{eq:1bit_T22bound}
\end{equation}
We now observe that the first two terms in \eqref{eq:1bit_Vlv} account for all of the terms in \eqref{eq:1bit_T1bound}, \eqref{eq:1bit_T21bound} and \eqref{eq:1bit_T22bound} except for $(1-\tau)^2 \EE[\Weq^2]$. Recalling that $\tau = \frac{1}{1+\sum_{i\in\sdif}b_i^2}$, we see that $(1-\tau)^2 = \Theta(1)$ whenever $\sum_{i\in\sdif}b_i^2 = \Omega(1)$, whereas a Taylor expansion yields $(1-\tau)^2 = \Theta\big( \big(\sum_{i\in\sdif}b_i^2\big)^2 \big)$ whenever $\sum_{i\in\sdif}b_i^2 = o(1)$. Combining these cases, we obtain the third term in \eqref{eq:1bit_Vlv}; recall that $\sigma^2=1$ throughout this proof.
\section{Proofs of Auxiliary Results for Noiseless Group Testing} \label{sec:PROOFS_GT}
\subsection{Proof of Proposition \ref{prop:gt_boundI}}
As stated in \cite[Eq.~(36)]{Ati12}, we have $I_{\ell} = \big(1-\frac{\nu}{k}\big)^{k-\ell} H_2\big( \big(1-\frac{\nu}{k}\big)^\ell \big)$, where $H_2(p)$ is the binary entropy function. For $k\to\infty$ and $\frac{\ell}{k}\to\alpha$, we immediately obtain \eqref{eq:gt_I_const} using the limits $\big(1-\frac{\nu}{k}\big)^{k-\ell} \to e^{-(1-\alpha)\nu}$ and $\big(1-\frac{\nu}{k}\big)^\ell \to e^{-\alpha\nu}$, along with the continuity of the binary entropy function. In the case that $\frac{\ell}{k} \to 0$, the analogous limits are $\big(1-\frac{\nu}{k}\big)^{k-\ell} \to e^{-\nu}$ and $\big(1-\frac{\nu}{k}\big)^\ell = 1 - \frac{\nu\ell}{k}(1+o(1))$, and we obtain \eqref{eq:gt_I_ord} using the fact that $H_2(1-\epsilon) = (-\epsilon\log\epsilon)(1+o(1))$ as $\epsilon\to0$. Note also that $\log\frac{k}{\nu\ell} = \big(\log\frac{k}{\ell}\big)(1+o(1))$ since $\frac{k}{\ell} \to \infty$.
\subsection{Proof of Proposition \ref{prop:conc_gt}}
We begin by evaluating the information density in \eqref{eq:idens_b}; for brevity, we write $\imath_{\ell} := \imath(X_{\sdif};Y|X_{\seq},b_s)$ and $\imath_{\ell}^n := \imath(\Xv_{\sdif};\Yv|\Xv_{\seq},b_s)$. Recalling that $P_X \sim \mathrm{Bernoulli}\big(\frac{\nu}{k}\big)$, $\ell = o(k)$, and we are considering the noiseless case, we obtain the following:
\begin{enumerate}
\item We have $X_{\seq} \ne \bzero$ with probability $1-\big(1-\frac{\nu}{k}\big)^{k-\ell} = (1 - e^{-\nu})(1+o(1))$, and in this case we have $\imath_{\ell} = 0$.
\item Given $X_{\seq} = \bzero$, we have $X_{\sdif} \ne \bzero$ with probability $1-\big(1-\frac{\nu}{k}\big)^{\ell} = \frac{\nu\ell}{k}(1+o(1))$, and in this case we have $\imath_{\ell} = \log\frac{1}{1-(1-\frac{\nu}{k})^{\ell}} = \big(\log\frac{k}{\ell}\big)(1+o(1))$.
\item Given $X_{\seq} = \bzero$, we have $X_{\sdif} = \bzero$ with probability $\big(1-\frac{\nu}{k}\big)^{\ell} = 1+o(1)$, and in this case we have $\imath_{\ell} = \log\frac{1}{(1-\frac{\nu}{k})^{\ell}} = \frac{\nu\ell}{k}(1+o(1))$.
\end{enumerate}
The asymptotic identities given here follow from the assumption $\ell = o(k)$, along with standard Taylor expansions.
Let $N_0$ (respectively, $N_1$) be the random number of measurements such that $X_{\seq} = \bzero$ and $X_{\sdif} = \bzero$ (respectively, $X_{\seq} = \bzero$ and $X_{\sdif} \ne \bzero$). For any $\epsilon_1 \in (0,1)$, the above observations imply the following with probability one when $p$ is sufficiently large:
\begin{align}
\imath_{\ell}^n &\ge N_1 \Big(\log\frac{k}{\ell}\Big) (1-\epsilon_1) + N_0 \nu\frac{\ell}{k} (1-\epsilon_1) \\
&\ge N_1 \Big(\log\frac{k}{\ell}\Big) (1-\epsilon_1).
\end{align}
We also have from \eqref{eq:gt_I_ord} that $I_{\ell} \le \big(e^{-\nu}\nu \frac{\ell}{k} \log\frac{k}{\ell}\big)(1+\epsilon_1)$ for sufficiently large $p$. Combining these, we conclude that
\begin{equation}
N_1 > n\frac{1+\epsilon_1}{1-\epsilon_1} e^{-\nu}\nu\frac{\ell}{k} (1-\delta_2) \implies \imath_{\ell}^n > nI_{\ell}(1-\delta_2).
\end{equation}
By considering the contrapositive statement, we have for any $\epsilon_2 > 0$ and sufficiently large $p$ that
\begin{multline}
\PP\Big[ \imath^n(\Xv_{\sdif}; \Yv | \Xv_{\seq}, b_s) \le nI_{\ell}(1-\delta_{2}) \Big] \\ \le \PP\Big[ N_1 \le ne^{-\nu}\nu\frac{\ell}{k}(1-\delta_2)(1+\epsilon_2) \Big]. \label{eq:gt_conc4}
\end{multline}
By the observations at the start of this subsection, we have $N_1 \sim \mathrm{Binomial}(n,q)$ with $q = e^{-\nu}\nu\frac{\ell}{k} (1+o(1))$. We can thus further upper bound the right-hand of \eqref{eq:gt_conc4} by
\begin{equation}
\PP\big[ N_1 \le nq(1-\delta_2(1-\epsilon_3)) \big]
\end{equation}
for any $\epsilon_3\in(0,1)$ and sufficiently large $p$; here we have used the fact that $(1-\delta_2)(1+o(1)) = (1-\delta_2(1+o(1))$, since $\delta_2$ is fixed. It follows from a standard Chernoff-based tail bound for Binomial random variables (e.g., see \cite[Sec.~4.1]{Mot10}) that
\begin{align}
&\PP\Big[ \imath^n(\Xv_{\sdif}; \Yv | \Xv_{\seq}, b_s) \le nI_{\ell}(1-\delta_{2}) \Big] \nonumber \\
&\le e^{-nq \big((1-\delta_2(1-\epsilon_3))\log(1-\delta_2(1-\epsilon_3)) + \delta_2(1-\epsilon_3) \big)}.
\end{align}
The proof is concluded by substituting $q = e^{-\nu}\nu\frac{\ell}{k} (1+o(1))$ and noting that $\epsilon_3$ may be arbitrarily small.
\subsection{Proof of Proposition \ref{prop:gt_psi}}
For the first part, we write $\sum_{\ell=1}^{\lfloor \frac{k}{\log k} \rfloor} {k \choose \ell} \psi_{\ell}(n,\delta_{2}^{(1)}) =: T_1 + T_2$, where $T_1$ sums the terms from $1$ to $\lfloor \log k \rfloor$, and $T_2$ sums the terms from $\lfloor \log k \rfloor +1$ to $\lfloor \frac{k}{\log k} \rfloor$. For each of these, we upper bound the summation by the number of terms times the maximum term.
For $T_1$, there are at most $\log k$ terms, and we apply \eqref{eq:gt_psi1}, with $\delta_{2,\ell} = \delta_2^{(1)}$. The term $(1-\delta_2^{(1)})\log(1-\delta_2^{(1)}) + \delta_2^{(1)}$ can be made arbitrarily close to one by choosing $\delta_2^{(1)}$ to be sufficiently close to one. Writing $\log{k \choose \ell} = \big(\ell\log\frac{k}{\ell}\big)(1+o(1))$ and performing some simple rearrangements, we obtain the following condition for $T_1 \to 0$:
\begin{equation}
n \ge \max_{\ell} \frac{k\log \frac{k}{\ell} + \frac{k}{\ell}\log\log k }{ e^{-\nu}\nu } (1+\eta_1), \label{eq:gt_psi_pf1}
\end{equation}
where $\eta_1$ may be arbitrarily small. Note that $\log\log k$ arises as the logarithm of the number of terms in the summation. We obtain \eqref{eq:gt_n_cond_psi} by noting that this bound is minimized at $\ell=1$ and writing $k\log k = \big(\frac{\theta}{1-\theta} k \log \frac{p}{k}\big) (1+o(1))$, which follows from $k=\Theta(p^\theta)$.
For $T_2$, a similar argument yields \eqref{eq:gt_psi_pf1} with $\frac{1}{\ell}\log k$ in place of $\frac{1}{\ell}\log\log k$; this follows by upper bounding the number of terms in the summation by $k$. Since $\ell \ge \log k$, we have $\frac{1}{\ell}\log k = O(1)$, and we conclude that $T_2 \to 0$ provided that \eqref{eq:gt_n_cond_psi} holds.
Finally, for the second part of the proposition, we substitute \eqref{eq:gt_psi2}. By an analogous argument to that leading to \eqref{eq:gt_psi_pf1}, along with the scaling laws of $I_{\ell}$ in \eqref{eq:gt_I_ord}--\eqref{eq:gt_I_const}, it is readily verified that it suffices that $n = \Omega\big( \max_{\ell} \frac{ \ell\log\frac{k}{\ell} }{ 1 + ( \frac{\ell}{k} \log \frac{k}{\ell} )^2 } \big)$ with a sufficiently large implied constant. Using the fact that $\ell > \frac{k}{\log k}$ for this part, this reduces to $\Omega\big( \frac{ k\log k }{\log\log k} \big)$. Thus, any $\Omega(k\log k)$ scaling suffices, and the proof is concluded by noting that $\log k = \Theta\big(\log\frac{p}{k}\big)$.
\section{Noisy Group Testing} \label{sec:PROOFS_GT_NOISY}
Here we provide the relevant details for noisy group testing, leading to Corollary \ref{cor:gt_noisy}. We focus our attention on the parts that differ from the noiseless case. Throughout the appendix, we use the notation $q_1 \star q_2 := q_1q_2 + (1-q_1)(1-q_2)$. We work with an arbitrary Bernoulli distribution $P_X \sim \mathrm{Bernoulli}\big(\frac{\nu}{k}\big)$ to begin, and later substitute $\nu = \log 2$.
Before proceeding, we analyze the values taken by the information density
$\imath_{\ell} := \imath(X_{\sdif};Y|X_{\seq},b_s)$ (with $\ell := |\sdif|$) given in \eqref{eq:idens_b}, under the model in \eqref{eq:gtn_model}:
\begin{enumerate}
\item We have $X_{\seq} \ne \bzero$ with probability $1-\big(1-\frac{\nu}{k}\big)^{k-\ell}$, and in this case we have $\imath_{\ell} = 0$.
\item Given $X_{\seq} = \bzero$, we have the following, where we define $\xi := \big(1-\frac{\nu}{k}\big)^{\ell}$:
\begin{itemize}
\item $X_{\sdif} = \bzero \cap Y = 0$ with probability $(1-\rho)\xi$, yielding $\imath_{\ell} = \log\frac{1-\rho}{(1-\rho)\xi + \rho(1-\xi)}$;
\item $X_{\sdif} = \bzero \cap Y = 1$ with probability $\rho\xi$, yielding $\imath_{\ell} = \log\frac{\rho}{\rho\xi + (1-\rho)(1-\xi)}$;
\item $X_{\sdif} \ne \bzero \cap Y = 0$ with probability $\rho(1-\xi)$, yielding $\imath_{\ell} = \log\frac{\rho}{(1-\rho)\xi + \rho(1-\xi)}$;
\item $X_{\sdif} \ne \bzero \cap Y = 1$ with probability $(1-\rho)(1-\xi)$, yielding $\imath_{\ell} = \log\frac{1-\rho}{\rho\xi + (1-\rho)(1-\xi)}$.
\end{itemize}
\end{enumerate}
In the case that $\ell = o(k)$, we can write $\xi = 1 - \frac{\nu\ell}{k} (1+o(1))$, yielding the following simplifications:
\begin{enumerate}
\item The preceding four probabilities behave as $(1-\rho)\big(1 - \frac{\nu\ell}{k}(1+o(1))\big)$, $\rho\big(1 - \frac{\nu\ell}{k}(1+o(1)) \big)$, $\rho\frac{\nu\ell}{k}(1+o(1))$, and $(1-\rho)\frac{\nu\ell}{k}(1+o(1))$.
\item The corresponding information densities behave as $\frac{1-2\rho}{1-\rho}\frac{\nu\ell}{k}(1+o(1))$, $-\frac{1-2\rho}{\rho}\frac{\nu\ell}{k}(1+o(1))$, $-\log\frac{1-\rho}{\rho}(1+o(1))$ and $\log\frac{1-\rho}{\rho}(1+o(1))$. For example, the first of these follows by writing $\log\frac{1-\rho}{(1-\rho)(1-\frac{\nu\ell}{k}) + \rho\frac{\nu\ell}{k} } = \log\frac{1 - \rho}{1 - \rho - (1-2\rho)\frac{\nu\ell}{k}}$, dividing the numerator and denominator by $1-\rho$, and Taylor expanding the logarithm.
\end{enumerate}
\subsection{Analogs of Propositions \ref{prop:gt_boundI}--\ref{prop:gt_psi}}
The analog of Proposition \ref{prop:gt_boundI} is as follows.
\begin{prop} \label{prop:gtn_boundI}
Under the noisy group testing setup in Section \ref{sec:GROUP_TESTING}, consider arbitrary sequences of sparsity levels $k\to\infty$ and $\ell \in \{1,\dotsc,k\}$ (both indexed by $p$). If $\frac{\ell}{k} = o(1)$, then
\begin{equation}
I_{\ell} = \bigg(e^{-\nu}\nu \frac{\ell}{k} (1-2\rho)\log\frac{1-\rho}{\rho}\bigg)(1+o(1)). \label{eq:gtn_I_ord}
\end{equation}
Moreover, if $\frac{\ell}{k}\to\alpha \in(0,1]$, then
\begin{equation}
I_{\ell} = e^{-(1-\alpha)\nu} \big(H_2\big(e^{-\alpha \nu} \star \rho\big) - H_2(\rho)\big) (1+o(1)). \label{eq:gtn_I_const}
\end{equation}
\end{prop}
\begin{proof}
We obtain \eqref{eq:gtn_I_ord} by recalling that the mutual information is the average of the information density, and applying the above-given asymptotic expansions, along with $1-\big(1-\frac{\nu}{k}\big)^{k-\ell} \to e^{-\nu}$.
To prove \eqref{eq:gtn_I_const}, we write $I(X_{\sdif};Y|X_{\seq}) = H(Y|X_{\seq}) - H(Y|X_{\seq},X_{\sdif})$. The system model \eqref{eq:gtn_model} immediately gives $H(Y|X_{\seq},X_{\sdif}) = H_2(\rho)$. Moreover, a direct calculation reveals that $H(Y|X_{\seq}=x_{\seq})$ equals $H_2(\rho)$ if $x_{\seq}$ has an entry equal to one, and $H_2\big(\xi \star \rho\big)$ otherwise, where we again write $\xi := \big(1-\frac{\nu}{k}\big)^{\ell}$. The proof is concluded by noting that $\xi \to e^{-\alpha\nu}$ when $\frac{\ell}{k} \to \alpha$, and by similarly noting that $\PP[X_{\seq} = \bzero] = \big(1 - \frac{\nu}{k}\big)^{k - \ell} \to e^{-(1-\alpha)\nu}$.
\end{proof}
As in the noiseless case, we use Proposition \ref{prop:gen_discrete} to characterize $\psi_{\ell}$ for $\ell > \lfloor \frac{k}{\log k} \rfloor$, and $\psi'_{\ell}$ for $\ell = k$. For $\ell \le \lfloor \frac{k}{\log k} \rfloor$, we instead use the following.
\begin{prop} \label{prop:conc_gtn}
Under the noisy group testing setup in Section \ref{sec:GROUP_TESTING}, consider sequences $k\to\infty$ and $\ell$, indexed by $p$, such that $\frac{\ell}{k} \to 0$. For any $\epsilon > 0$ and $\delta_{2} > 0$ not depending on $p$, the following holds for sufficiently large $p$:
\begin{multline}
\PP\Big[ \imath^n(\Xv_{\sdif}; \Yv | \Xv_{\seq}, b_s) \le nI_{\ell}(1-\delta_{2}) \Big] \\ \le \exp\bigg(-n\frac{\ell}{k} e^{-\nu}\nu\Big( \frac{\delta_2^2 (1-2\rho)^2}{2(1+\frac{1}{3}\delta_2(1-2\rho))} \Big)(1-\epsilon)\bigg). \label{eq:conc_gt_noisy}
\end{multline}
for all $(\sdif,\seq)$ with $|\sdif|=\ell$.
\end{prop}
\begin{proof}
We make use of the asymptotic identities for $\imath_{\ell}$ at the start of this appendix. We first note that by simple averaging analogous to that used to obtain \eqref{eq:gtn_I_ord}, we have $v := \EE[\imath_{\ell}^2] = e^{-\nu}\nu \frac{\ell}{k} \big(\log^2\frac{1-\rho}{\rho}\big) (1+o(1))$. Moreover, we have $\imath_{\ell} \le \big(\log\frac{1-\rho}{\rho}\big)(1+o(1))$ with probability one. Using the form of Bernstein's inequality based on Bennet's inequality \cite[Sec.~2.7]{Bou13}, we have $\PP[\imath^n \le n(I_{\ell} - \delta)] \exp\big(-n \frac{\delta^2}{2(v + \frac{1}{3}\delta M)} \big)$, where $M$ is any almost-sure upper bound on $\imath_{\ell}$. Setting $\delta = \delta_2 I_{\ell}$, substituting \eqref{eq:gtn_I_ord} and the preceding expressions for $v$ and $M$, and canceling the common terms in the numerator and denominator, we obtain \eqref{eq:conc_gt_noisy}.
\end{proof}
Letting $\psi_{\ell}$ equal the right-hand side of \eqref{eq:conc_gt_noisy} for $\ell \le \lfloor \frac{k}{\log k} \rfloor$, while being the same as in \eqref{eq:gt_psi2} for $\ell > \lfloor \frac{k}{\log k} \rfloor$, we obtain the following.
\begin{prop} \label{prop:gtn_psi}
Let $k = \Theta( p^{\theta} )$ for some $\theta \in (0,1)$.
(i) For any $\eta > 0$ and $\delta_2 \in (0,1)$, there exists a choice of $\epsilon > 0$ in \eqref{eq:gt_psi1} such that $\sum_{\ell=1}^{\lfloor \frac{k}{\log k} \rfloor} {k \choose \ell} \psi_{\ell}(n,\delta_{2}) \to 0$ provided
\begin{equation}
n \ge \frac{ 2(1+\frac{1}{3}\delta_2(1-2\rho)) \frac{\theta}{1-\theta}}{ e^{-\nu}\nu \delta_2^2 (1-2\rho)^2 } \, \Big(k\log\frac{p}{k}\Big) (1+\eta). \label{eq:gtn_n_cond_psi}
\end{equation}
(ii) For any $\delta_2 \in (0,1)$, we have $\sum_{\lfloor \frac{k}{\log k} \rfloor +1}^k {k \choose \ell} \psi_{\ell}(n,\delta_{2}) \to 0$ provided that $n = \Omega\big(k\log\frac{p}{k}\big)$.
\end{prop}
\begin{proof}
The proof is nearly identical to that of Proposition \ref{prop:gt_psi}, except that \eqref{eq:conc_gt_noisy} is used in place of \eqref{eq:conc_gt}, and $\delta_2$ is kept arbitrary in the first part.
\end{proof}
Note that the choices of $\delta_2$ in the two cases above need not coincide; see Remark \ref{rem:delta2}.
\subsection{Remaining Details in the Proof of Corollary \ref{cor:gt_noisy}}
Recall that we have set $\nu = \log 2$. This yields $e^{-\nu}\nu$ = $\frac{\log 2}{2}$, and thus the first term in \eqref{eq:gt_zeta} follows from \eqref{eq:gtn_n_cond_psi}.
Next, we consider the condition in \eqref{eq:final_ach} with $\ell = |\sdif| \le \lfloor \frac{k}{\log k} \rfloor$. Setting $\gamma = 0$, letting $\delta_1 \to 0$ sufficiently slowly, applying Stirling's approximation, and substituting \eqref{eq:gtn_I_ord}, we obtain the condition
\begin{equation}
n \ge \max_{\ell} \frac{ k \log \frac{p}{\ell} + 2k\log k + 2\frac{k}{\ell}\log k}{ e^{-\nu}\nu (1-2\rho)\log\frac{1-\rho}{\rho} (1-\delta_2) } (1+o(1)). \label{eq:gtn_main_cond0}
\end{equation}
This is maximized for $\ell = 1$, thus yielding the second term in \eqref{eq:gt_zeta} upon writing $k\log k = \frac{\theta}{1-\theta} \big(k\log\frac{p}{k}\big) (1+o(1))$ and $k \log p = \frac{1}{1-\theta} \big(k\log\frac{p}{k}\big) (1+o(1))$ (since $k=\Theta(p^{\theta})$).
Finally, we consider \eqref{eq:final_ach} with $\ell > \lfloor \frac{k}{\log k} \rfloor$. In this case, the numerator is dominated by the first term, and for the case that $\frac{\ell}{k} \to \alpha\in(0,1]$, we obtain the condition
\begin{equation}
n \ge \frac{ \alpha k \log \frac{p}{k}}{ e^{-(1-\alpha)\nu} \big(H_2(e^{-\alpha \nu} \star \rho) - H_2(\rho)\big) (1-\delta_2) } (1+o(1)), \label{eq:gtn_main_cond1}
\end{equation}
where we have used \eqref{eq:gtn_I_const}. For the case that $\frac{\ell}{k} \to 0$ with $\ell > \lfloor \frac{k}{\log k} \rfloor$, we obtain a condition of the form \eqref{eq:gtn_main_cond0} where only the first term of the numerator is kept. Such a condition is clearly dominated by \eqref{eq:gtn_main_cond0}.
Using the result in \cite[Thm.~3a]{Mal78} in the limiting case that the number of defective items grows large, we have for the worst-case choice of $\alpha \in [0,1]$ and an optimized choice of $\nu > 0$ that the minimax threshold resulting from \eqref{eq:gtn_main_cond1} is obtained with $\alpha = 1$ and $\nu = \log 2$. Substituting these values yields the second term in \eqref{eq:gtn_ach}.
\subsection{An Auxiliary Result for Comparing the Terms}
The following result allows us to compare the terms appearing in the achievability part of Corollary \ref{cor:gt_noisy}.
\begin{prop} \label{prop:gt_cmp}
For all $\rho \in (0,0.5)$, we have
\begin{equation}
(1-2\rho) \log\frac{1-\rho}{\rho} \ge 4\big(\log 2 - H_2(\rho)\big).
\end{equation}
\end{prop}
\begin{proof}
By some simple manipulations, the left-hand side can be written as $\log\frac{1}{\rho(1-\rho)} - 2H_2(\rho)$, and we may thus equivalently prove that $\log\frac{1}{\rho(1-\rho)} + 2H_2(\rho) \ge 4\log 2$. This, in turn, can be verified by showing that the minimum of the function $\log\frac{1}{\rho(1-\rho)} + 2H_2(\rho)$ occurs at $\rho=0.5$, i.e., the point about which it is symmetric.
\end{proof}
\section*{Acknowledgment}
We gratefully acknowledge Ya-Ping Hsieh for helpful comments and suggestions.
\bibliographystyle{IEEEtran}
|
\section[Introduction]{Introduction}
In this paper we establish a connection between two overdetermined boundary value problems,
Serrin's symmetry problem and what we call the {\it parallel surface problem}. As a consequence,
we obtain optimal stability for the former, thus significantly improving previous results (\cite{ABR}, \cite{CMV}).
\par
Serrin's symmetry problem concerns solutions of elliptic partial differential equations subject to both Dirichlet and Neumann boundary conditions. Let $\Omega\subset{\mathbb R^N}$ be a bounded domain with sufficiently smooth boundary $\Gamma$ (say $C^{2,\tau}$, $0<\tau<1$). As shown in one of his seminal papers, \cite{Se},
if the following problem
\begin{eqnarray}
&\Delta u + f(u) = 0 \ \mbox{ and } \
u>0 \ \mbox{ in } \ \Omega, \ u=0 \ \mbox{ on } \ \Gamma,
\label{serrin1}
\\
&u_\nu=\frak a \ \mbox{ on } \ \Gamma,
\label{serrin2}
\end{eqnarray}
admits a solution for a given positive constant $\frak a$, then $\Omega$ must be a ball. Here, $f:[0,+\infty)\to\mathbb R$ is a locally Lipschitz continuous function and $\nu$ is the {\it inward} unit vector field to $\Gamma$.
Generalizations of these result
are innumerable and we just mention \cite{BCN}, \cite{BNV}, \cite{GL}, \cite{GNN}, \cite{Re}.
\par
The parallel surface problem concerns solutions of \eqref{serrin1} with a level surface {\it parallel} to $\Gamma$,
that is to say the solution $u$ of \eqref{serrin1} is required to be constant at a fixed distance from $\Gamma$:
\begin{equation}
\label{parallelcondition}
u=\frak b \ \mbox{ on } \ \Gamma^\delta.
\end{equation}
Here,
\begin{equation} \label{Gamma delta}
\Gamma^\delta=\{ x\in\Omega: d_\Omega(x)=\delta\} \ \ \mbox{ with } \ \ d_\Omega(x)=\min_{y\notin\Omega} |x-y|, \ x\in{\mathbb R^N},
\end{equation}
is a surface parallel to $\Gamma$, and $\frak b$ and $\delta$ are positive (sufficiently small) constants.
Under sufficient conditions on $\Gamma^\delta$, also in this case, if a solution of \eqref{serrin1} and \eqref{parallelcondition} exists, then $\Omega$ must be a ball (see \cite{MSaihp, MSmmas, Sh, CMS1, GGS}). In the sequel, we will occasionally refer to problem \eqref{serrin1}, \eqref{parallelcondition} as the {\it parallel surface problem}.
\par
A condition like \eqref{parallelcondition} was considered in \cite{MSaihp, MSmmas} in connection with
{\it time-invariant} level surfaces of a solution $v$ of the
non-linear equation
$$
v_t=\Delta\phi(v) \ \mbox{ in } \ \Omega\times(0,\infty)
$$
subject to the initial and boundary conditions
$$
v=1 \ \mbox{ on } \ \Omega\times\{0\} \ \mbox{ and } \ v=0 \ \mbox{ on } \ \partial\Omega\times(0,\infty);
$$
here, $\phi$ is a $C^2$-smooth non-linearity with $\phi(0)=0$ and $\phi'$ bounded
from below and above by two positive constants (hence, we are dealing with a non-degenerate fast-diffusion equaton).
A (spatial) level surface $\Sigma\subset\Omega$ of $v$ is time-invariant
if $v$ is constant on $\Sigma$ for each fixed time $t>0$.
\par
It is proved in \cite{MSaihp} that for $x\in\overline{\Omega}$
$$
4t\,\int_0^{v(x,t)}\frac{\phi'(\eta)}{1-\eta}\,d\xi\longrightarrow d_\Omega(x)^2 \ \mbox{ as } \ t\to 0^+,
$$
uniformly on compact subsets of $\Omega$. Hence, if $\Sigma$ is time-invariant, then it has to be parallel
to $\Gamma$ at some distance $\delta$.
Also, it is not difficult to show that the function
$$
u(x)=\int_0^\infty \phi(v(x,t))\,dt, \ x\in\overline{\Omega},
$$
satisfies \eqref{serrin1}, with $f(u)=1$ and, being $\Sigma$ time-invariant, $u$ satisfies \eqref{parallelcondition}, with $\Gamma^\delta=\Sigma$, for some positive constant $\frak b$.
As a consequence, $\Omega$ is a ball and, in this situation, the time-invariant surface $\Sigma$ turns out to be
a sphere.
\par
Condition \eqref{parallelcondition} can also be re-interpreted to give a connection to transnormal and
isoparametric functions and surfaces. We recall that, in differential geometry, a function $u$ is {\it transnormal} in $\Omega$ if it is a solution of the equation
\begin{equation}
\label{transnormal}
|Du|^2=g(u) \ \mbox{ in } \ \Omega,
\end{equation}
for some suitably smooth function $g:\mathbb R\to (0,\infty)$; the level surfaces of $u$ are called
{\it transnormal surfaces}. A transnormal function that also satisfies the first equation in \eqref{serrin1}
is called an {\it isoparametric function} and its level surfaces are {\it isoparametric surfaces}. Isoparametric surfaces in the euclidean space can only be (portions of) spheres, spherical cylinders or hyperplanes (\cite{L-C}, \cite{Sg});
a list of essential references about transnormal and isoparametric functions and their properties
in other spaces includes \cite{Bo}, \cite{Ca}, \cite{Mi}, \cite{Wa}.
\par
Now, notice that the (viscosity) solution $u$ of \eqref{transnormal} such that $u=0$ on $\Gamma$
takes the form $u(x)=h(\mbox{\rm dist}(x,\partial\Omega))$, where $h$ is defined by
$$
\int_0^{h(t)}\frac{ds}{\sqrt{g(s)}}=t, \ \ t\ge 0.
$$
It is then clear that a solution of \eqref{serrin1} satisfies \eqref{parallelcondition} if and only if $u=\frak b$ on $\{x\in\Omega: w(x)=h(\delta)\}$. Thus, we can claim that {\it if a solution of \eqref{serrin1} has a level surface that is also a level surface of a transnormal function, then $\Omega$ must be a ball.}
\par
Both problems \eqref{serrin1}, \eqref{serrin2} and \eqref{serrin1}, \eqref{parallelcondition} have at least one feature in common: the proof of symmetry relies on the {\it method of moving planes}, a refinement, designed by J. Serrin, of a previous idea of V.I. Aleksandrov's \cite{Al}. The evident similarity between the two problems arouses a natural question:
{\it to obtain the symmetry of $\Omega$, is condition \eqref{serrin2} weaker or stronger than \eqref{parallelcondition}?}
\par
As noticed in \cite{CMS2} and \cite{CM}, condition
\eqref{parallelcondition} seems to be weaker than \eqref{serrin2}, in the sense clarified hereafter. As \eqref{parallelcondition} does not imply \eqref{serrin2}, the latter can be seen as the limit of a sequence of conditions of type \eqref{parallelcondition} with $\frak b=\frak b_n$ and $\delta=\delta_n$ and
$\frak b_n$ and $\delta_n$ vanishing as $n\to\infty$. As \eqref{serrin2} does not imply \eqref{parallelcondition} either, nonetheless the oscillation on a surface parallel to $\Gamma$ of a solution of \eqref{serrin1}, \eqref{serrin2} becomes smaller than usual, the closer the surface is to $\Gamma$.
A way to quantitavely express this fact is to consider the Lipschitz seminorm
\begin{equation*}
[u]_{\Gamma^\delta} = \sup_{{\substack{x,y \in\Gamma^\delta, \\ x\neq y}}} \frac{|u(x)-u(y)|}{|x-y|}\,,
\end{equation*}
that controls the oscillation of $u$ on $\Gamma^\delta$: it is not difficult to show by a Taylor-expansion argument (see the proof of Theorem \ref{thm Serrin by stability}) that,
if $u\in C^{2,\tau}(\overline{\Omega})$, $0<\tau<1$, satisfies \eqref{serrin1}, \eqref{serrin2}, then
\begin{equation}
\label{seminorm-tau}
[u]_{\Gamma^\delta}=O(\delta^{1+\tau}) \ \mbox{ as } \ \delta\to 0.
\end{equation}
\par
This remark suggests the possibility that Serrin's symmetry result may be obtained {\it by stability} by the following strategy: (i) for the solution $u$ of \eqref{serrin1} prove that,
for some constant $C_\delta$ depending on $\delta$,
an estimate of type
\begin{equation}
\label{estimate}
r_e-r_i\le C_\delta\, [u]_{\Gamma^\delta}
\end{equation}
holds for any sufficiently small $\delta>0$, where $r_e$ and $r_i$ are the radii of a spherical annulus centered at some point $\frak 0$, $\{x\in{\mathbb R^N}: r_i<|x-\frak 0|<r_e\}$, containing $\Gamma$ --- this means that $\Omega$ is nearly a ball, if $u$ does not oscillate too much on $\Gamma^\delta$;
(ii)
if in addition $u$ satisfies \eqref{serrin2}, that is $u_\nu$ is constant on $\Gamma$, show that
$$
\ [u]_{\Gamma^\delta}=o(C_\delta)\ \mbox{ as } \ \delta\to 0^+.
$$
The spherical symmetry of $\Gamma$ then will follow by choosing $\delta$ arbitrarily small.
\par
In \cite{CMS2}, the first two authors of this paper proved an estimate of type \eqref{estimate}. Unfortunately, that estimate is not sufficient for our aims, since the computed constant $C_\delta$ blows up exponentially as $\delta\to 0^+$. Nevertheless, in the case examined in
\cite{CM}, we showed that our stategy is successful for the very special class of ellipses.
\par
In this paper we shall extend the efficacy of our strategy to the class of $C^{2,\tau}$-smooth domains.
The crucial step in this direction is Theorem \ref{thm stability dependence on s}, where we considerably improve inequality \eqref{estimate} by
showing that it holds with a constant $C_\delta$ that is a $O(\delta^{-1})$ as $\delta\to 0^+$. Thus, \eqref{seminorm-tau} will imply that $r_i=r_e$, that is $\Omega$ is a ball (see Theorem \ref{thm Serrin by stability}). By a little more effort, in Theorem \ref{thm stability by stability} we will prove that
$$
r_e-r_i\le C\,[u_\nu]_\Gamma,
$$
for some positive constant $C$. This inequality enhances to optimality two previous results, both
also based on the method of moving planes. In fact, it
improves the logarithmic stability obtained in \cite{ABR} for $C^{2,\tau}$-regular domains and extends the linear stability obtained for convex domains in \cite{CMV}. We notice that, in our inequality the seminorm $[u_\nu]_\Gamma$ replaces the deviation in the $C^1(\Gamma)$-norm
of the function $u_\nu$ from a given constant, considered in \cite{ABR}. Moreover, the inequality also
improves \cite{BNST}[Theorem 1.2], where a H\"older-type estimate was obtained, for the case in which $f(u)=-N$, by means of integral identities.
\par
The outline of the proof of Theorem \ref{thm stability dependence on s} will be recalled in Section 2: it is the same as that of \cite[Theorem 4.1]{CMS2}, that relies on ideas introduced in \cite{ABR}, and the use of {\it Harnack's} and
{\it Carleson's (or the boundary Harnack's) inequalities}. In Section 3 --- the heart of this paper --- by the careful use of refined versions of those inequalities (see \cite{Ba}, \cite{BCN}), we prove the necessary lemmas that in Section 4 allow us to obtain
our optimal version of \eqref{estimate}.
Finally, in Section 5, we present our new linear stability estimate for the radial symmetry in Serrin's problem; it implies symmetry for \eqref{serrin1}, \eqref{serrin2}: thus, the new strategy is successful.
\setcounter{equation}{0}
\setcounter{theorem}{0}
\section{A path to stability: \\
the quantitative method of moving planes}
\label{sec:preliminaries}
In this Section we introduce some notation and we review the quantitative study of the method of moving planes as carried out in \cite{ABR} and \cite{CMS2} (see also \cite{CV}).
Let $\Omega$ be a bounded domain in $\mathbb R^N$ $(N\geq 2)$ and $\Gamma$ be its boundary; we shall denote the diameter of $\Omega$ by $\frak d_\Omega$. The distance $d_\Omega$ defined in \eqref{Gamma delta} is always Lipschitz continuous on $\overline{\Omega}$ and of class $C^{2,\tau}$
in a neighborhood of $\Gamma$, if this is of class $C^{2,\tau}$, $0<\tau<1$. In fact, under this assumption on $\Gamma$, for every $x\in\Gamma$ there are balls
$B\subset\Omega$ such that $x\in\partial B$; denote
by $r_x$ the supremum of the radii of such balls and set
$$
\frak r_\Omega=\min_{x\in\partial\Omega}r_x.
$$
We then denote by $\frak R_\Omega$ the number obtained by this procedure, where instead the interior ball $B$ is replaced by an exterior one.
\par
For $\delta>0$, let $\Omega^\delta$ be the {\it parallel set} (to $\Gamma$), i.e.
\begin{equation}\label{Om delta}
\Omega^\delta= \{x \in \Omega:\ d_\Omega(x)>\delta\} \,.
\end{equation}
We know that if $0\le\delta<\frak r_\Omega$, then each level surface $\Gamma^\delta$ of $d_\Omega$, as defined by \eqref{Gamma delta}, is of class $C^{2,\tau}$ and will be referred to as
a {\it parallel surface} (to $\Gamma$).
The following notations are useful to carry out the method of moving planes and its quantitative version;
for $\omega\in\SS^{N-1}$ and $\mu\in\mathbb R$, we set:
\begin{equation}
\label{definitions}
\begin{array}{lll}
&\pi_{\mu}=\{ x\in{\mathbb R^N}: x\cdot\omega=\mu\}\ &\mbox{a hyperplane orthogonal to $\omega,$}\\
&\mathcal{H}_\mu=\{ x\in{\mathbb R^N}: x\cdot\omega>\mu\}\ &\mbox{the half-space \emph{on the right} of $\pi_\mu$,}\\
&A_{\mu}=A\cap\mathcal{H}_\mu &\mbox{the right-hand cap of a set $A$},\\
&x'=x-2(x\cdot\omega-\mu)\,\omega\ &\mbox{the reflection of $x$ in $\pi_{\mu},$}\\
&(A_\mu)'=\{x\in{\mathbb R^N}:x'\in A_{\mu}\}\ &\mbox{the reflected cap in $\pi_{\mu}$.}
\end{array}
\end{equation}
In the sequel, we will generally use the simplified notation $A'=(A_\mu)'$ every time in which the dependence on $\mu$ is not important.
\par
Set $\Lambda=\sup\{x\cdot\omega: x\in \Omega\}$; if $\mu<\Lambda$ is close to $\Lambda$, the reflected cap $(\Omega_\mu)'$ is contained in $\Omega$ (see \cite{Fr}), and hence we can define the number
\begin{equation}\label{m def}
\lambda=\inf\{\mu: (\Omega_{\tilde\mu})'\subset \Omega \mbox{ for all } \tilde\mu\in(\mu,\Lambda)\}.
\end{equation}
Thus, at least one of the following two cases occurs (\cite{Se},\cite{Fr}):
\begin{enumerate}
\item[(S1)]
$\Omega'=(\Omega_{\lambda})'$ is internally tangent to $\partial \Omega$ at some
point $p'\in\partial \Omega'\setminus\pi_{\lambda}$, which is the reflection in $\pi_\lambda$ of a point $p\in \partial \Omega_\lambda\setminus\pi_{\lambda}$;
\item[(S2)]
$\pi_{\lambda}$ is orthogonal to $\partial \Omega$ at some point $q\in\partial \Omega\cap\pi_{\lambda}$.
\end{enumerate}
In the sequel, $\pi_\lambda$ and $\Omega_\lambda$ will be referred to as the {\it critical hyperplane} and the {\it critical cap} (in the direction $\omega$), respectively. Corresponding to the points $p$ and $q$, we will also consider
the points $p^\delta=p+\delta\,\nu(p)$ and $q^\delta=q+\delta\,\nu(q)$ for $0<\delta< \frak r_\Omega$; notice that
$\Gamma^\delta=\{p^\delta:p\in\Gamma\}$.
\par
Let $\frak r\in (0, \frak r_\Omega)$. From now on $G$ will denote the parallel set
$$
G=\{ x\in\Omega: d_\Omega(x)>\frak r\}.
$$
In Section 4, we shall choose $\frak r$ appropriately. Also, to simplify notations, by $P$ and $Q$ we shall denote $p^{\frak r}$ and $q^{\frak r}$, respectively --- two points on $\partial G$ that will be frequently used.
\par
Now, the function $w$ defined by
\begin{equation}\label{w def}
w(x)= u(x')-u(x),\quad x\in \Omega_\lambda,
\end{equation}
satisfies
\begin{equation*} \label{eq w semilinear}
\Delta w+c(x)\,w=0 \ \mbox{ in } \Omega_{\lambda},
\end{equation*}
where for $x\in\Omega_{\lambda}$
\begin{equation*}
\label{defc}
c(x)=\left\{
\begin{array}{lll}
\displaystyle\frac{f(u(x'))-f(u(x))}{u(x^\lambda)-u(x)} &\mbox{ if } u(x')\not= u(x),\\
\displaystyle 0 &\mbox{ if } u(x') = u(x).
\end{array}
\right.
\end{equation*}
Notice that $c(x)$ is bounded by the Lipschitz constant $\frak L$ of $f$ in the interval
$[0,\max\limits_{\overline{\Omega}}u].$
\par
By an argument introduced in \cite[Theorem 2]{Se} and improved in \cite{BNV} (see also \cite{Fr}),
we can assume that $w\ge 0$ in $\Omega_\lambda$ and hence, by the strong maximum principle
applied to the inequality $\Delta w-c^-(x)\, w\le 0$ with $c^-(x)=\max[-c(x),0]$, we can suppose that
$$
w>0 \ \mbox{ in } \ \Omega_\lambda.
$$
\par
One ingredient in our estimates of Section 3 will be {\it Harnack's inequality}: thanks to this result, for
fixed $a\in (0,1)$, $w$ satisfies the inequality
\begin{equation} \label{harnack}
\sup_{B_{ar}} w \leq \frak{H}_a \inf_{B_{ar}} w ,
\end{equation}
for any ball $B_r \subset \Omega_\lambda$ (see \cite[Theorem 8.20]{GT}); the Harnack constant $\frak{H}_a$ can be bounded by the power
$\sqrt{N} + \sqrt{r \frak L}$ of a constant
only depending on $N$ and $a$ (see \cite{GT}). For instance, if $c(x) \equiv 0$, by the explicit Poisson's representation formula for harmonic functions, we have that
\begin{equation*
\sup_{B_{ar}} w \leq \left( \frac{1+a}{1-a} \right)^N \inf_{B_{ar}} w ,
\end{equation*}
for any $B_r \subset \Omega_\lambda$ (see \cite{GT} and \cite{DBGV}).
Now, we review the quantitative study of the method of moving planes established in \cite{CMS2}, partly
based on the work in \cite{ABR}. As already mentioned in the Introduction, the stability of the radial configuration for problem \eqref{serrin1}, \eqref{parallelcondition} is obtained in \eqref{estimate} in terms of the Lipschitz seminorm on parallel surfaces to $\Gamma$.
For a fixed direction $\omega$ we consider the critical positions and the corresponding points $p$ and $q$, as detailed in (S1) and (S2). As shown in \cite{MSmmas}, the method of moving planes can be applied to $G$ instead of $\Omega$, and the tangency points of cases (S1) and (S2) are $P$ and $Q$,
respectively. It is clear that if an estimate like \eqref{estimate} holds for $G$, then the same holds for $\Omega$, since the difference of the radii does not change.
\par
The procedure to obtain \eqref{estimate} is quite delicate. For the reader's convenience, we give an outline of it, in which we identify $8$ salient steps.
\begin{enumerate}[(i)]
\item
Following the proof of \cite[Theorem 3.3]{CMS2}, we show that the values of $w(P)/\mbox{\rm dist}(P,\pi_\lambda)$, in case (S1), and of the partial derivative $w_\omega(Q)$, in case (S2), are bounded by some constant times $[u]_{\partial G}$.
\item By Harnack's inequality, the smallness obtained in (i) at the points $P$ and $Q$ propagates to any point in $G_\lambda$ sufficiently far from $\partial G_\lambda$ (see \cite[Lemma 3.1]{CMS2}).
\item By using Carleson's inequality, the estimation obtained in the previous step extends to any point in the cap $G_\lambda$ (see \cite[Lemma 3.1]{CMS2}), thus obtaining the inequality
\begin{equation*}
\|w\|_{L^\infty(G_\lambda)} \leq C [u]_{\partial G}.
\end{equation*}
Here, the key remark is that $C$ only depends on $N$, $\frak r$, the diameter and the $C^2$-regularity of $G$, but {\it does not} depend on the particular direction $\omega$ chosen.
\item The union of $G\cap\overline{\mathcal{H}_\lambda}$ with its reflection in $\pi_\lambda$ defines a set $X$, symmetric in the direction $\omega$, that approximates $G$, since the smallness of $w$ bounds that of $u$ on $\partial X$.
\item Since $u$ is the solution of \eqref{serrin1}, then $u(x)$ grows linearly with $d_\Omega(x)$, when $x$ moves inside $\Omega$ from $\Gamma$; this implies that $u$ can not be too small on $\partial G=\Gamma^{\frak r}$.
\item By using both steps (iv) and (v), we find that the distance of every point in $\partial X$ from $\partial G$ is not greater than some constant times $[u]_{\partial G}$ (\cite[Lemma 3.4]{CMS2}). This means that $X$ {\it fits well} $G$, in the sense that $X$ contains the parallel set $G^\sigma$ (related to $G$) for some positive (small) number $\sigma$ controlled by $[u]_{\partial G}$ (\cite[Theorem 3.5]{CMS2}). This fact is what we call a {\it quantitative approximate symmetry} of $G$ in the direction $\omega$.
\item An approximate center of symmetry $\frak 0$ is then determined as the intersection of $N$ mutually orthogonal critical hyperplanes. As shown in \cite[Proposition 6]{ABR}, the distance between
$\frak 0$ and any other critical hyperplane can be uniformly bounded in terms of the parameter $\sigma$ in item (vi) and hence of $[u]_{\partial G}$.
\item The point $\frak 0$ is finally chosen as the center of the spherical annulus $\{x\in{\mathbb R^N}: r_i<|x-\frak 0|<r_e\}$ and the estimate \eqref{estimate} follows from \cite[Proposition 7]{ABR}.
\end{enumerate}
Based on this plan, to improve \eqref{estimate}, it is sufficient to work on the estimates in step (ii). This will be done in Section 3, by refining our use of Harnack's and Carleson's inequalities. As a matter of fact, in \cite{CMS2} we merely
used a standard application of Harnack's inequality, by constructing a Harnack's chain of balls of suitably chosen fixed radius $\frak r$. This strategy only yields an exponential dependence on $\frak r^{-1}$
of the constant in \eqref{estimate}. In \cite{CMV}, we improved these estimates by chosing a chain of balls with radii that decay linearly when the balls approach $\Gamma$; however, this could be done only when $\Omega$ is convex (or little more) and leads to a H\"{o}lder type dependency on $\frak r^{-1}$ of the constant in \eqref{estimate}.
\par
In Section 3 instead, we use the following plan, that we sketch for the case $\omega=e_1$ and $\lambda=0$;
$p$ and $q$ are the points defined in (S1) and (S2).
\par
We fix $\frak r=\frak r_\Omega/4$, so that $G=\Omega^{\frak r_\Omega/4}$. For any $0<\delta<\frak r_\Omega/4$, the values of $w(x)/x_1$ at the points $p^\delta$ and $P=p^{\frak r}$ can be compared in the following way
\begin{equation*}
w(P)/P_1 \leq C\,\delta^{-1}\, w(p^{\delta})/p^{\delta}_1,
\end{equation*}
where $C$ is a constant not depending on $\delta$. Correspondingly, we prove that
\begin{equation*}
w_{x_1}(Q) \leq C_G\,\delta^{-1}\, w_{x_1}(q^{\delta}),
\end{equation*}
where $Q=q^\frak r$. Then, by exploiting steps (i) and (iii), we obtain that
\begin{equation} \label{w leq seminorm sect2}
w \leq C_G\,\delta^{-1}\, [u]_{\Gamma^\delta} \ \mbox{ on the maximal cap of $G_\lambda$,}
\end{equation}
where the constant $C_G$ is the one obtained in step (iii) by letting $\frak r=\frak r_\Omega/4$, and hence it does not depends on $\delta$.
\par
Once this work is done, steps (iv)--(viii) can be repeated to find the improved approximate symmetry for the parallel set $G$, which clearly implies that for $\Omega$. We underline the fact that the dependence on $\delta$ in \eqref{w leq seminorm sect2} is optimal, as \cite{CM} indicates.
\setcounter{equation}{0}
\setcounter{theorem}{0}
\section{Enhanced stability estimates}
\label{sec: stability estimates}
In this section, we line up the major changes needed to obtain \eqref{estimate}; they only concern step (ii). The following lemma will be useful in the sequel.
\begin{lm} \label{lemma ball ABR}
Let $p$ and $q$ be the points defined in (S1) and (S2), respectively.
\par
If $B$ is a ball of radius $\frak r_\Omega$, contained in $\Omega$ and such that $p$ or $q$ belong to $\partial B$,
then the center of $B$ must belong to $\overline{\Omega}_\lambda$.
\end{lm}
\begin{proof}
The assertion is trivial for case (S2). If case (S1) occurs, without loss of generality, we can assume that $\omega=e_1$ and $\lambda=0$. Since (S1) holds, the reflected point $p'$ lies on $\partial \Omega$ and cannot fall inside $B$, since $p\in\partial B$ and $B\subset \Omega$.
\par
Thus, if $c$ is the center of $B$, we have that $|c-p'|\geq \frak r_\Omega=|c-p|$ and hence $|c_1+p_1| \geq |c_1-p_1|$, which implies that $c_1 \geq 0$, being $p_1>0$.
\end{proof}
Our first estimate is a quantitative version of Hopf's lemma, that will be useful to treat both occurrences (S1) and (S2).
\begin{lm} \label{lemma 1 semilinear}
Let $B_R=\{ x\in{\mathbb R^N}: |x|<R\}$ and, for $p\in \partial B_R$ and
$s\in (0,R)$, set $p^s=p+s\,\nu(p)=(1-s/R)\, p.$
\par
Let $c\in L^\infty(B_R)$ and suppose that $w\in C^0(B_R \cup \{p\}) \cap C^2(B_R)$ satisfies the conditions:
\begin{equation*}
\label{delta w semilinear}
\Delta w + c(x) w = 0 \ \mbox{ and } \
w \geq 0 \ \mbox{ in } B_R.
\end{equation*}
\par
Then, there is a constant $A=A(N, R, \|c\|_{\infty})$\footnote{See the proof for its expression.} such that
\begin{equation}
\label{w 0 leq semilinear}
w(0) \leq A\,s^{-1} w(p^s) \ \mbox{ for any } \ s \in (0,R/2).
\end{equation}
Moreover, if $w(p)=0$, then
\begin{equation}
\label{w 0 leq semilinear II}
w(0) \leq A\, w_{\nu} (p).
\end{equation}
\end{lm}
\begin{proof}
We proceed as in the standard proof of Hopf's boundary point lemma.
\par
Notice that $w$ also satisfies
\begin{equation*}
\Delta w - c^-(x) w \leq 0 \ \mbox{ in } \ B_R,
\end{equation*}
where $c^-(x)=\max(-c(x),0)$. Thus, the strong maximum principle implies that $w >0$ in $B_R$ (unless $w\equiv 0$, in which case the conclusion is trivial).
\par
For a fixed $a\in (0,1)$ and some parameter $\alpha>0$, set
\begin{equation*}
v(x)=\frac{|x|^{-\alpha}-R^{-\alpha}}{(a^{-\alpha}-1) R^{-\alpha}} \ \mbox{ for } \ a R\le |x|\le R;
\end{equation*}
notice that $v>0$ in $B_R$, $v=0$ on $\partial B_R$ and $v=1$ on $\partial B_{aR}$.
For $a R< |x|< R$ we then compute that
\begin{multline*}
\Delta v-c^-(x)\,v\ge \frac{\alpha^2-(N-2)\alpha -|x|^2 c^-(x)}{(a^{-\alpha}-1) R^{-\alpha}}\,|x|^{-\alpha-2}\\
\ge
\frac{\alpha^2-(N-2)\alpha -R^2\| c^-\|}{(a^{-\alpha}-1) R^{-\alpha}}\,|x|^{-\alpha-2}.
\end{multline*}
Hence, we see that
$$
\Delta v-c^-(x)\,v\ge 0 \quad \textmd{ in } B_R \setminus \overline{B}_{aR},
$$
if we choose
\begin{equation*}
\alpha =\frac{ N-2+\sqrt{(N-2)^2+4 R^2 \| c^-\|}}{2}.
\end{equation*}
\par
With this choice of $\alpha$, the function
\begin{equation*}
z=w-\left[\min_{\partial B_{aR}} w\right] v
\end{equation*}
satisfies the inequalities:
\begin{equation*}
\Delta z - c^-(x)\, z \leq 0 \ \textmd{ in } \ B_R \setminus \overline{B}_{a R} \quad \mbox{ and } \quad
z \geq 0 \ \textmd{ on } \ \partial (B_R \setminus B_{a R}).
\end{equation*}
Thus, the maximum principle gives that $z\ge 0$ and hence that
\begin{equation*}
\min_{\partial B_{aR}} w \leq \frac{w(x)}{v(x)},
\end{equation*}
for $x\in B_R \setminus \overline{B}_{a R}$.
\par
Now, choose $x = p^s$ (since we want that $p^s\in B_R \setminus \overline{B}_{a R}$, the constraint $s/R<1-a$ is needed); we thus have that
\begin{equation} \label{1 semilinear}
\min_{\partial B_{aR}} w \leq \frac{a^{-\alpha}-1}{(1-s/R)^{-\alpha}-1}\,w(p^s) \le R\,\frac{a^{-\alpha}-1}{\alpha}\,
\frac{w(p^s)}{s},
\end{equation}
where the last inequality holds for the convexity of the function $t\mapsto t^{-\alpha}$.
\par
Harnack's inequality \eqref{harnack} then yields
\begin{equation*}
\label{3 semilinear}
w(0) \leq \sup_{B_{aR}} w \leq \frak{H}_a \inf_{B_{aR}} w \leq R \frak H_a\,\frac{a^{-\alpha}-1}{\alpha}\,
\frac{w(p^s)}{s}.
\end{equation*}
Consequently, by chosing $a=1/2$ and setting
$
A=R \frak{H}_{1/2} (2^{\alpha}-1)/\alpha,
$
we readily obtain \eqref{w 0 leq semilinear}.
Finally, if $w(p)=0$, we readily obtain \eqref{w 0 leq semilinear II} from \eqref{w 0 leq semilinear} and by letting $s$ go to zero.
\end{proof}
The following result is crucial to treat the case (S2).
\begin{lm}
\label{lemma 2 semilinear}
Set $B_R^+=\{ x\in B_R: x_1 >0\}$ and $T=\{ x\in\partial B_R^+: x_1=0\}$. For
any point $q\in\partial B_R\cap T$ and $s\in [0,R)$, define
$q^s = q + s \nu(q)=(1-s/R)\,q.$
\par
Let $c\in L^\infty(B_R)$ and suppose $w \in C^2(B_R^+) \cap C^1(B_R^+ \cup T)$ satisfies the conditions:
\begin{equation*}
\Delta w + c(x) w = 0
\, \mbox{ and } \, w \geq 0 \ \textmd { in } \ B_R^+,
\quad w = 0 \ \textmd{ on } \ T.
\end{equation*}\par
Then, there is a constant $A^*=A^*(N, R, \|c\|_{\infty})$ such that
\begin{equation}
\label{bound-wx1}
w_{x_1}(0) \leq A^* s^{-1}\, w_{x_1} (q^s)
\ \mbox{ for any } \ s \in (0,R/2].
\end{equation}
\end{lm}
\begin{proof}
As in Lemma \ref{lemma 1 semilinear}, we can assume that $w>0$ in $B_R^+$.
\par
Inequality \eqref{bound-wx1} will be the result of a chain of estimates: with this goal, we introduce the half-annulus $\mathcal{A}^+= B_R^+ \setminus \overline{B^+_\rho}$ and the cube
$$
Q_\sigma=\{(x_1,\dots, x_N)\in{\mathbb R^N}: \, 0<x_1<2\sigma, \,|x_i|<\sigma,\, i=2,\dots, N\}.
$$
For the moment, we choose $0<\rho<R$ and $0<\sigma\le R/\sqrt{N+3}$, that is in such a way that $Q_{\sigma}\subset B_R^+$; the precise value of $\rho$ will be specified later.
\par
The first estimate of our chain is \eqref{up-bound-sup} below; in order to prove it, we introduce the auxiliary function
$$
v(x)=[|x|^{-\alpha}- R^{-\alpha}]\,x_1 \ \mbox{ for } \ x \in \overline{A^+}.
$$
It is clear that $v>0$ in $\mathcal{A}^+$ and $v=0$ on $\partial B^+_R$; also, we can choose $\alpha>0$ so that $\Delta v-c^-(x)\, v\ge 0$ in $\mathcal{A}^+$ ($\alpha=(N+\sqrt{N^2+4 R^2 \|c^-\|_\infty})/2$ will do).
\par
We then consider the function $w/v$ on $\partial B^+_\rho$: it is surely well-defined, positive and continuous in $\partial B^+_\rho\setminus T$; also, it can be extended to be a continuous function up to $T\cap\partial B^+_\rho$ by defining it equal to its limiting values
$
w_{x_1}(x)/(\rho^{-\alpha}-R^{-\alpha}|)
$
for $x\in T\cap\partial B^+_\rho$. With this settings, $w/v$ also turns out to be positive on the whole $\partial B^+_\rho$ since, on $T\cap\partial B^+_\rho$, it is positive by a standard application of Hopf lemma.
\par
These remarks tell us that the minimum of $w/v$ on $\partial B^+_\rho $
is well-defined and positive, and hence that the function
$$
z=w-\min_{\partial B^+_\rho}(w/v)\,v
$$
satisfies the inequalities
\begin{equation*}
\Delta z - c^-(x)\, z \leq 0 \ \textmd{ in } \ \mathcal{A}^+ \quad \mbox{ and } \quad
z \geq 0 \ \textmd{ on } \ \partial \mathcal{A}^+.
\end{equation*}
Thus, by the maximum principle, $z\ge 0$ on $\overline{\mathcal{A}^+}$ and hence
$$
\min_{\partial B^+_\rho}(w/v)\le w(x)/v(x) \ \mbox{for every} \ x\in \overline{\mathcal{A}^+}.
$$
\par
For $s<R-\rho$, we then can take $x=q^s+\varepsilon e_1$, with $\varepsilon>0$ so small that $x\in \mathcal{A}^+$,
take the limit as $\varepsilon\to 0$ and, since $w(q^s)=v(q^s)=0$, obtain the inequality
$$
\min_{\partial B^+_\rho}(w/v)\le \frac{w_{x_1}(q^s)}{v_{x_1}(q^s)}=
\frac{w_{x_1}(q^s)}{(R-s)^{-\alpha}-R^{-\alpha}}.
$$
Again, by the convexity of $t\mapsto t^{-\alpha}$, we find that
\begin{equation*}
\min_{\partial B^+_\rho}(w/v)\le \frac{R^{\alpha+1}}{\alpha\,s}\,w_{x_1}(q^s),
\end{equation*}
and hence it holds that
\begin{equation}
\label{up-bound-sup}
\min_{x\in\partial B^+_\rho}\frac{w(x)}{x_1}\le \frac{R^{\alpha+1}(\rho^{-\alpha}-R^{-\alpha})}{\alpha\,s}\,w_{x_1}(q^s).
\end{equation}
\par
The second estimate \eqref{w 2rho e1 leq} below shows that, up to a constant, the minimum in \eqref{up-bound-sup} can be
bounded from below by the value of $w$ at the center of the cube $Q_{\sigma}$. To do this, we let $y\in\partial B^+_\rho$ be a point at which the minimum in \eqref{up-bound-sup}
is attained and set
\begin{equation*}
\hat y =(0,y_2,\ldots,y_N)\quad\textmd{and}\quad \bar y=(\sigma,y_2,\ldots,y_N);
\end{equation*}
notice that $\hat y$ and $y$ coincide when $y_1=0$.
\par
The ball $B_{\sigma}(\bar y)$ is tangent to $\partial B_R^+\cap T$ at $\hat y$ and we can choose $\rho$ such that $B_{\sigma}(\bar y)\subset B_R^+$; thus, by applying Lemma \ref{lemma 1 semilinear} to $B_{\sigma}(\bar y)$
with $p=\hat y$, $p^s=y$ and $\nu=e_1$,
we obtain that
\begin{equation*}
w(\bar y) \leq A\,w(y)/y_1 \ \mbox{ if } \ y_1>0,
\end{equation*}
and
\begin{equation*}
w(\bar y) \leq A\,w_{x_1} (y) \ \mbox{ if } \ y_1=0, \ \mbox{ being $w(\hat y)=0$. }
\end{equation*}
Thus, we have proved that
\begin{equation*}\label{w bar y leq}
w(\bar y) \leq A \min_{x\in\partial B^+_\rho}\frac{w(x)}{x_1}.
\end{equation*}
Moreover, if we also choose $\rho$ such that $B_\rho(\bar y)\subset B_{2\rho}(\bar y) \subset B_R^+$, since the point $\sigma e_1\in B_\rho(\bar y)$, Harnack's inequality shows that
\begin{equation*}
w (\sigma e_1) \leq \frak{H}_{1/2}\, w(\bar y),
\end{equation*}
and hence we obtain that
\begin{equation}\label{w 2rho e1 leq}
w(\sigma e_1) \leq A\,\frak{H}_{1/2}\, \min_{x\in\partial B^+_\rho}\frac{w(x)}{x_1}.
\end{equation}
\par
To conclude the proof, we use two estimates contained in \cite{BCN} (see also \cite{Ba}).
First, after some rescaling, we can apply \cite[Lemma 2.1]{BCN} to the square $Q_{\sigma/2}$ and obtain that
\begin{equation}
\label{harnack-up-to-boundary}
w(t\sigma/2\, e_1)\le t \,C_1\,\max_{\overline{Q}_{\sigma/2}} w,
\end{equation}
for every $t\in(0,1)$, where $C_1$ is the constant in \cite[Lemma 2.1]{BCN} that, in our case, only depends
on $N$, $\| c\|_\infty$ and $R$ (by means of $\sigma$). Thus, since $w(0)=0$, taking the limit as $t\to 0^+$ gives that
\begin{equation}
\label{pa w pa x1 leq C1 M}
w_{x_1}(0)\le\frac{2 C_1}{\sigma}\,\max_{\overline{Q}_{\sigma/2}} w.
\end{equation}
\par
Secondly, we consider the cube $Q_{\sigma}$ and again after some rescaling, we use the Carleson-type estimate \cite[Theorem 1.3]{BCN} to obtain that
\begin{equation}
\label{carleson}
\max_{\overline{Q}_{\sigma/2}} w \leq 2^q\,B\,w (\sigma\, e_1),
\end{equation}
where, in our case, the constants $B$ and $q$ in \cite[Eq.(1.6)]{BCN} again only depend on $N$, $R$ and $\| c\|_\infty$. Thus, by \eqref{pa w pa x1 leq C1 M} we have that
\begin{equation}\label{pa w pa x1 leq C3 w}
w_{x_1}(0) \leq \frac{2^{q+1} B\,C_1}{\sigma}\, w (\sigma e_1).
\end{equation}
\par
Therefore, by applying \eqref{pa w pa x1 leq C3 w}, \eqref{w 2rho e1 leq} and
\eqref{up-bound-sup}, inequality \eqref{bound-wx1} holds with
$$
A^*=2^{q+1} A\,\frak{H}_{1/2}\, B\,C_1\,\frac{(R/\rho)^\alpha-1}{\alpha}\,\frac{R}{\sigma} ,
$$
where the constants $\rho$ and $\sigma$ can be chosen as specified along the proof.
\end{proof}
For the treatment of case (S1), we must pay attention to the fact that the point of tangency $p$ may be very close to $\pi_\lambda$ and the interior touching ball at $p$ may not be contained in the cap. For this reason, we need the following lemma which gives a uniform treatment of all cases occurring when (S1) takes place.
\begin{lm} \label{lemma 3 semilinear}
Let $\xi =\xi_1 e_1$ with $\xi_1>0$ and set
\begin{equation*}
B_R^+(\xi)=\{x \in {\mathbb R^N}:|x-\xi|<R, \, x_1>0\}, \ T=\{x \in {\mathbb R^N}:|x-\xi|<R, \, x_1=0\}. \footnote{Notice that $T$ may be the empty set.}
\end{equation*}
For $p\in \partial B_R^+(\xi) \setminus T$, define $p^s$
as in Lemma \ref{lemma 1 semilinear}.
\par
Let $c$ be essentially bounded on $B_R(\xi)$ and suppose that $w\in C^2(B_R^+(\xi)) \cap C^0(B_R^+(\xi) \cup T)$ satisfies
\begin{equation*}
\Delta w + c(x)\,w = 0 \
\mbox{ and } \ w \geq 0 \textmd{ in } B_R^+(\xi),
\quad w = 0 \ \textmd{ on } \ T.
\end{equation*}
\par
Then, there is a constant $A^\#=A^\#(N, R, \|c\|_{\infty})$ such that
\begin{equation}
\label{bound-wx1-s1}
w(\xi)/\xi_1 \leq A^\#\,s^{-1}\, w (p^s)/p_1^s
\ \mbox{ for any } \ s \in (0,R/2].
\end{equation}
\end{lm}
\begin{proof}
We proceed similarly to the proof of Lemma \ref{lemma 2 semilinear}, with some modifications.
We shall still use the cube $Q_\sigma$, but we will instead consider the half annulus $\mathcal{A}^+=B_R^+(\xi) \setminus \overline{B}_\rho^+(\xi)$.
\par
Next, we change the auxiliary function $v$:
$$
v(x)=[|x-\xi|^{-\alpha}- R^{-\alpha}]\,x_1;
$$
of course $v=0$ on $\partial B_R^+(\xi)$ and we can still choose $\alpha$ so large that $v$ satisfies the inequality
$\Delta v-c^-(x)\, v\ge 0$ in $\overline{\mathcal{A}^+}$. Thus, the function\footnote{As in the proof of Lemma \ref{lemma 2 semilinear}, we observe that the function $w/v$ can be extended continuously on $T$ and $w/v>0$ on $\partial B_\rho^+(\xi)$.}
$$
z=w-\min_{\partial B^+_\rho}(w/v)\,v
$$
is such that
$
\Delta z - c^-(x)\, z \leq 0
$
in $\mathcal{A}^+$ and $z \geq 0$ on $\partial \mathcal{A}^+$.
By the maximum principle, we obtain that $z\ge 0$ on $\overline{\mathcal{A}^+}$ and hence
\begin{equation*}
\min_{\partial B^+_\rho(\xi)}(w/v)\le w(x)/v(x) \ \mbox{for every} \ x\in \overline{\mathcal{A}^+}.
\end{equation*}
Again, by arguing as in the proof of Lemma \ref{lemma 2 semilinear}, we find that
\begin{equation}\label{up-bound-sup II}
\min_{x\in\partial B^+_\rho(\xi)} \frac{w(x)}{x_1}
\le \frac{R^{\alpha+1}(\rho^{-\alpha}-R^{-\alpha})}{\alpha\,s}\,\frac{w(p^s)}{p_1^s}.
\end{equation}
\par
Now, to conclude the proof we will treat the cases $\xi_1\geq 2\rho$ and $\xi_1\leq2\rho$, separately.
\par
If $\xi_1\geq 2\rho$, the ball $B_{2\rho}(\xi)$ is contained in $B_R^+(\xi)$ and hence, by Harnack's inequality, we have:
$$
\frac{w(\xi)}{\xi_1}\le \frac{w(\xi)}{2\rho}\le \frac{\frak{H}_{1/2}}{2\rho} \min_{\partial B^+_\rho(\xi)} w\le
\frac{\eta+\rho}{2\rho}\,\frak{H}_{1/2} \min_{x\in\partial B^+_\rho(\xi)}\frac{w(x)}{x_1}.
$$
Thus, \eqref{up-bound-sup II} gives that
\begin{equation*}
\frac{w(\xi)}{\xi_1} \le (\eta+\rho)\,\frak{H}_{1/2} \,\frac{R}{\rho}\,\frac{(R/\rho)^{\alpha}-1}{4\alpha\,s}\,\frac{w(p^s)}{p_1^s}.
\end{equation*}
\par
If $\xi_1\leq 2\rho$, we repeat the arguments of the last part of the proof of Lemma \ref{lemma 2 semilinear}, with some slight modification. We take a point $y \in \partial B_\rho^+(\xi)$ at which
the minimum in \eqref{up-bound-sup II} is attained
and set $\bar y=(\sigma,y_2,\ldots,y_N)$, $\hat y=(0, y_2,\dots,y_N)$. We apply Lemma \ref{lemma 1 semilinear} to $B_\sigma(\bar y)$, with $p=\hat y$, $p^s=y$ and $\nu=e_1$ and, by inspecting the two cases $y_1>0$ and $y_1=0$, we obtain that
\begin{equation*}
w(\bar y) \leq A \min_{x\in\partial B^+_\rho(\xi)}\frac{w(x)}{x_1}.
\end{equation*}
As before, we choose $\rho$ such that $B_\rho(\bar y)\subset B_{2\rho}(\bar y) \subset B_R^+$ and, since $\sigma e_1\in B_{\rho}(\bar y)$, by Harnack's inequality we find that
$
w(\sigma e_1) \leq \frak{H}_{1/2}\, w(\bar y),
$
and hence
\begin{equation}\label{w 4rho e1 leq}
w(\sigma e_1) \leq \frak{H}_{1/2} \,A \min_{x\in\partial B^+_\rho(\xi)}\frac{w(x)}{x_1}.
\end{equation}
\par
Now, we apply \cite[Theorem 1.3]{BCN} to the cube $Q_\sigma$ and obtain \eqref{carleson} as before.
Moreover, again we use
\cite[Lemma 2.1]{BCN} in the cube $Q_{\sigma/2}$; if $\rho$ is sufficiently small, we have that $\xi_1\le 2\rho\le \sigma/2$ and hence, applying \eqref{harnack-up-to-boundary} with $t=2\xi_1/\sigma$ gives that
\begin{equation*}
\frac{ w(\xi)}{\xi_1} \leq C_1 \max_{\overline{Q}_{\sigma/2}} w;
\end{equation*}
therefore, \eqref{carleson} yields:
\begin{equation*}
\frac{ w(\xi)}{\xi_1} \leq 2^q B\,C_1 w (\sigma e_1).
\end{equation*}
From \eqref{w 4rho e1 leq} and \eqref{up-bound-sup II}, we conclude in this case, as well.
The constant $A^\#$ can be computed by suitably choosing $\rho$ and $\sigma$ according to the
instructions specified in the proof.
\end{proof}
\setcounter{equation}{0}
\setcounter{theorem}{0}
\section{Approximate symmetry}
In this section, we assist the reader to adapt the theorems obtained in \cite{CMS2} in order to prove our new result of approximate symmetry for $\Omega$. First, we prove the analogue of \cite[Theorem 3.3]{CMS2}, that gives an estimate on the symmetry of $\Omega$ in a fixed direction.
\begin{theorem} \label{th:w bounded usn}
Let $\Omega\subset{\mathbb R^N}$ be a bounded domain with boundary $\Gamma$ of class $C^{2,\tau}$, $0<\tau<1$,
and set $G=\Omega^{\frak r_\Omega/4}$. For a unit vector $\omega\in{\mathbb R^N}$, let $G_\lambda$ and $\Omega_\lambda$ be the maximal caps in the direction
$\omega$ for $G$ and $\Omega$, respectively.
\par
Let $u\in C^{2,\tau}(\overline{\Omega})$ be a solution of \eqref{serrin1} and let $w$ be defined by \eqref{w def}.
\par
Then, for every $\delta\in (0,\frak r_\Omega/8)$, we have that
\begin{equation}\label{smallness}
w^{\lambda}\leq C \delta^{-1} [u]_{\Gamma^\delta} \ \mbox{ on (a connected component of)} \ G_\lambda.
\end{equation}
Here, $C$ is a constant depending on $N$, $\frak L$, $\frak d_\Omega$
and the $C^{2,\tau}$-regularity of $\Gamma$.
\end{theorem}
\begin{proof}
We point out that $G$ is connected.
Also, as already done before, we can assume that $\omega=e_1$ and $\lambda=0$.
\par
Let $p$ and $q$ be the points defined in (S1) and (S2), respectively; $P$ and $Q$ are the points in $\partial G$ already defined.
\par
In what follows, we chose to still denote by $\Omega_\lambda$ and $G_\lambda$ the connected components of the maximal caps $\Omega_\lambda$ and $G_\lambda$ that intersect $B_{\frak r_\Omega/4}(P)$, if case (S1) occurs,
and the connected components of $\Omega_\lambda$ and $G_\lambda$ that intersect $B_{\frak r_\Omega/4}(Q)$, if case (S2) occurs.
\par
Lemma \ref{lemma ball ABR} ensures that the interior ball of radius $\frak r_\Omega$ touching $\partial \Omega$ at $p$ or $q$ has its center in $\overline{\Omega}_\lambda$; hence, $P\in \Omega_\lambda$ and $Q\in \partial \Omega_\lambda \cap \pi_\lambda$. We then apply \cite[Lemma 4.2]{CMS2} with the following settings: $D_1=G_\lambda$, $D_2=\Omega_\lambda$, $R=\frak r_\Omega/4$, and $z=P$, if case (S1) occurs, and $z=Q$, if case (S2) occurs. Thus, we find that
\begin{equation} \label{w leq in Gm I}
w(x) \leq C\,w(P)/P_1 \ \mbox{ for } \ x \in\overline{G}_\lambda,
\end{equation}
and
\begin{equation} \label{w leq in Gm II}
w(x) \leq C\,w_{ x_1}(Q) \ \mbox{ for } \ x \in \overline{G}_\lambda,
\end{equation}
respectively.
Here, the constant $C$ depends only on $N, \frak r_\Omega, \frak L$ and $\frak d_\Omega$.
If (S1) occurs, we apply Lemma \ref{lemma 3 semilinear} by letting $R=\frak r_\Omega/4$ and $\xi=P$ (this is always possible after a translation in a direction orthogonal to $e_1$), and from \eqref{w leq in Gm I} we obtain that
\begin{equation} \label{w leq in Gm I II}
w(x) \leq C\,A^\# \delta^{-1} w(p^\delta)/p_1^\delta \ \mbox{ for } \ x \in \overline{G}_\lambda,
\end{equation}
for any $\delta\in (0,\frak r_\Omega/8)$.
\par
If (S2) occurs, we apply instead Lemma \ref{lemma 2 semilinear} (with $\xi = 0$ and $R=\frak r_\Omega/4$) and \eqref{w leq in Gm II}: we find that
\begin{equation} \label{w leq in Gm II II}
w(x) \leq C\,A^* \delta^{-1} w_{x_1}(q^\delta) \ \mbox{ for } \ x \in \overline{G}_\lambda,
\end{equation}
for any $\delta \in (0,\frak r_\Omega/4)$.
The rest of the proof runs similarly to that of \cite[Theorem 3.3]{CMS2}, where the estimates of \cite[Lemma 3.2]{CMS2} should be replaced by \eqref{w leq in Gm I II} and \eqref{w leq in Gm II II}. For the reader's convenience, we give a sketch of the proof with the usual settings ($\omega=e_1$ and $\lambda=0$). In particular, we show how to relate $w(p^\delta)/p_1^\delta$ and $w_{x_1} (q^\delta)$ to $[u]_{\Gamma_\delta}$, which is the main argument of the proof.
Let us assume that case (S1) occurs. If $p_1^\delta \geq \frak r_\Omega/2$, since $p^\delta$ and its reflection $(p^\delta)'$ about $\pi_\lambda$ lie on $\Gamma^\delta$, then
\begin{equation*}
w(p^\delta) = u((p^\delta)')-u(p^\delta) \leq \frak d_\Omega\,[u]_{\Gamma^\delta},
\end{equation*}
and hence we easily obtain that
\begin{equation} \label{estim I}
w(p^\delta)/p_1^\delta\leq 2 \frak d_\Omega \frak r_\Omega^{-1}\, [u]_{\Gamma^\delta}.
\end{equation}
If $p_1^\delta < \frak r_\Omega/2$, then $|p^\delta-(p^\delta)'|<\frak r_\Omega$, then every point of the segment joining $(p^\delta)'$ to $p^\delta$ is at a distance not greater than $\frak r_\Omega$ from some connected component of $\Gamma^\delta$. The curve $\gamma$ obtained by projecting that segment on that component has length bounded by $\widehat C\,|p^\delta-(p^\delta)'|$, where $\widehat C$ is a constant depending on $\frak r_\Omega$ and the regularity of $\Gamma^\delta$ (and hence on the regularity of $\Gamma$ since $\delta< \frak r_\Omega/8$).
An application of the mean value theorem to the restriction of $u$ to $\gamma$ gives that $u((p^\delta)')-u(p^\delta)$ can be estimated by the length of $\gamma$ times the maximum of the tangential gradient of $u$ on $\Gamma^\delta$. Thus,
\begin{equation}\label{estim II}
w(p^\delta) \leq 2\,\widehat C\, p_1^\delta \,[u]_{\Gamma^\delta}.
\end{equation}
Therefore, \eqref{estim I} and \eqref{estim II} yield the conclusion, if case (S1) is in force.
\par
Case (S2) is simpler. Since $e_1$ belongs to the tangent hyperplane to $\Gamma_\delta$ at $q^\delta$, we readily obtain \eqref{smallness}.
\end{proof}
As outlined in Section \ref{sec:preliminaries}, Theorem \ref{th:w bounded usn} completes steps (i)-(iii) and leads to stability bounds for the symmetry in one direction. Now, we complete steps (iv)-(viii).
\par
As described in steps (iv) and (vi), we define a symmetric open set $X$ and show that $G$ is almost equal to $X$. In order to do that, we need a priori bounds on $u$ from below in terms of the distance function from $\partial G$, as specified in (v). As observed in \cite{ABR} and \cite{CMS2}, such a bound requires a positive lower bound for $u_\nu$ on $\Gamma$,
$$
u_\nu\ge\frak c_{\frak 0} \ \mbox{ on } \ \Gamma.
$$
If $f(0)>0$, this is guaranteed by Hopf lemma.
If $f(0)\le 0$ instead, such a bound must be introduced as an assumption, as
it can be realized by considering any (positive) multiple of the first Dirichlet eigenfunction $\phi_1$ for $-\Delta$. In fact, for any $n\in\mathbb N$ the function $\phi_1/n$ satisfies \eqref{serrin1} with $f(u)=\lambda_1\,u$, being $\lambda_1$ the first Dirichlet eigenvalue, and it is clear that,
although $(\phi_1/n)_\nu\to 0$ on $\Gamma$ as $n\to\infty$, one cannot expect to derive any
information on the shape of $\Omega$.
The final stability result, step (viii), is obtained by defining an approximate center of symmetry $\frak 0$ as the intersection of $N$ orthogonal hyperplanes as described in step (vii) (see also \cite[Proof of Theorem 1.1]{CMS2}).
\par
We can now conclude this section with our improved stability estimate on the symmetry of $\Omega$.
\begin{theorem}
\label{thm stability dependence on s}
Let $\Omega\subset{\mathbb R^N}$ be a bounded domain with boundary $\Gamma$ of class $C^{2,\tau}$, $0<\tau<1$.
Let $u\in C^{2,\tau}(\overline{\Omega})$ be a solution of \eqref{serrin1}.
\par
There exist constants $\varepsilon, C>0$ and $\delta_0 \in (0,\frak r_\Omega/4)$ such that, if
\begin{equation}
\label{seminorm-ep}
[u]_{\Gamma^{\delta_0}} \leq \varepsilon,
\end{equation}
then there are two concentric balls $B_{r_i}$ and $B_{r_e}$ such that
\begin{equation} \label{Bri-Omega-Bre}
B_{r_i}\subset \Omega \subset B_{r_e}
\end{equation}
and
\begin{equation} \label{stability t}
r_e-r_i\le C\,\delta^{-1}\, [u]_{\Gamma^{\delta}},
\end{equation}
for any $\delta \in (0, \delta_0]$.
The constants $\varepsilon$ and $C$ only depend on $N$, $\frak r_\Omega$, $\frak d_\Omega$, $\frak L$, $\frak c_{\frak 0}$, $\max_{\overline\Omega} u$ and the $C^{2,\tau}$-regularity
of $\Gamma$.
\end{theorem}
\begin{proof}
Thanks to Theorem \ref{th:w bounded usn}, we can repeat the argument of the proof of \cite[Theorem 4.2]{CMS2} in which we replace formula \cite[(3.15)]{CMS2} by \eqref{smallness}. Hence, there exists two concentric balls $B_{r^*_i}$ and $B_{r^*_e}$ and two constants $\varepsilon$ (independent of $\delta$) such that
$$
B_{r^*_i}\subset G \subset B_{r^*_e} \ \mbox{ and } \ r^*_e-r^*_i\le C\,\delta^{-1}\, [u]_{\Gamma^{\delta}},
$$
if $[u]_{\Gamma^\delta}\leq \varepsilon$.
Moreover, since $\varepsilon$ does not depend on $\delta$ and
\begin{equation*}
\lim_{\delta\to 0^+} [u]_{\Gamma_\delta} = 0,
\end{equation*}
we can find $\delta_0 \in (0,\frak r_\Omega)$ such that $[u]_{\Gamma^\delta}\leq \varepsilon$ for any $\delta\in (0, \delta_0)$.
To complete the proof, we observe that \eqref{Bri-Omega-Bre} and \eqref{stability t}
hold with $r_i=r^*_i+\frak r_\Omega/4$ and $r_e=r^*_e+\frak r_\Omega/4$.
\end{proof}
\setcounter{equation}{0}
\setcounter{theorem}{0}
\section{Serrin's problem} \label{section Serrin via stability}
In this section, we give a new proof of Serrin's symmetry result and a corresponding stability estimate for spherical symmetry by using the improved stability inequality for the parallel surface problem \eqref{serrin1}, \eqref{parallelcondition}, as just proved in Theorem \ref{thm stability dependence on s}. We need the following lemma.
\begin{lm} \label{lemma seminorm asympt}
Let $\Omega$ be a bounded domain with boundary $\Gamma$ of class $C^2$ and set $r=\min(\frak r_\Omega, \frak R_\Omega)$. Let $u$ be
of class $C^2$ in a neighborhood of $\Gamma$ and such that $u=0$ on $\Gamma$.
\par
Then
\begin{equation}
\label{seminorms}
[u]_{\Gamma^\delta}\le \frac{\delta}{1-\delta/r}\,[u_\nu]_{\Gamma}
+\int_0^\delta \frac{(\delta-t)(r-t)}{r-\delta}\,[u_{\nu\nu}]_{\Gamma^t}\,dt,
\end{equation}
for every $\delta\in [0,r)$.
\par
In particular, if $\delta\le r/2$, we have that
\begin{equation}
\label{seminorms2}
[u]_{\Gamma^\delta}\le 2\,\delta\,\left\{[u_\nu]_{\Gamma}
+\int_0^\delta [u_{\nu\nu}]_{\Gamma^t}\,dt\right\}.
\end{equation}
\end{lm}
\begin{proof}
Let $p_1$ and $p_2$ be two points on $\Gamma$, so that $p_i^\delta=p_i+\delta\,\nu(p_i)$, $i=1, 2$,
are points on $\Gamma^\delta$. It is clear that $|p_1^\delta-p_2^\delta|\ge |p_1-p_2|-\delta\,|\nu(p_1)-\nu(p_2)|$
and hence:
\begin{equation}
\label{p1p2}
|p_1^\delta-p_2^\delta|\ge (1-\delta/r)\,|p_1-p_2|.
\end{equation}
\par
By applying Taylor's formula to the values of $u$ at $p_1^\delta$ and $p_2^\delta$ and taking the difference, we have that
\begin{multline*} \label{u k 2}
u(p_1^\delta)-u(p_2^\delta) = \delta\,[u_\nu(p_1)-u_\nu(p_2)]
+ \int_0^\delta (\delta-t)\, [u_{\nu\nu}(p_1^t) - u_{\nu\nu}(p_2^t)]\,dt,
\end{multline*}
since $u=0$ at $p_1$ and $p_2$ and being $\nu(p_i^t)=\nu(p_i)$, $i=1, 2$.
Dividing both sides by $|p_1^\delta-p_2^\delta|$ and using \eqref{p1p2}, gives that
\begin{multline*}
\frac{|u(p_1^\delta)-u(p_2^\delta)|}{|p_1^\delta-p_2^\delta|}\le \frac{\delta}{1-\delta/r}\,\frac{|u_\nu(p_1)-u_\nu(p_2)|}{|p_1-p_2|}
+\\
\int_0^\delta \frac{(\delta-t)(r-t)}{r-\delta}\,
\frac{|u_{\nu\nu}(p_1^t) - u_{\nu\nu}(p_2^t)|}{|p_1^t-p_2^t |}\,dt\le \\
\frac{\delta}{1-\delta/r}\,[u_\nu]_{\Gamma}
+\int_0^\delta \frac{(\delta-t)(r-t)}{r-\delta}\,[u_{\nu\nu}]_{\Gamma^t}\,dt.
\end{multline*}
\par
Therefore, \eqref{seminorms} and hence \eqref{seminorms2} follow at once.
\end{proof}
We are now in position to prove both symmetry and stability for Serrin's problem.
Of course, stability implies symmetry, when the normal derivative of $u$ is exactly constant on $\Gamma$.
However, we prefer to present the two results separately.
\begin{theorem}[Symmetry] \label{thm Serrin by stability}
Let $\Omega\subset{\mathbb R^N}$ be a bounded domain with boundary $\Gamma$ of class $C^{2,\tau}$, $0<\tau<1$. Let $u\in C^{2,\tau}(\overline\Omega)$, satisfy \eqref{serrin1} and suppose that \eqref{serrin2} holds with $\frak a>0$.
\par
Then $\Omega$ is a ball.
\end{theorem}
\begin{proof}
The assumed regularity of $u$ implies that $[u_{\nu\nu}]_{\Gamma^t}=O(t^{\tau-1})$ as $t\to 0$
and hence,
since \eqref{serrin2} is in force, \eqref{seminorms2} tells us that \eqref{seminorm-ep} holds for some $\delta_0>0$. Thus, Theorem \ref{thm stability dependence on s} can be applied and, by \eqref{seminorms2}, we have that
$$
B_{r_i}\subset \Omega \subset B_{r_e} \ \mbox{ and } \ r_e-r_i\le 2 C \int_0^{\delta} [u_{\nu\nu}]_{\Gamma^t}\,dt
$$
for any $\delta\in(0,\delta_0)$. The behavior of $[u_{\nu\nu}]_{\Gamma^t}$ as $t\to 0$ then implies that the integral at the right-hand side can be made
arbitrarily small. Therefore, $r_e=r_i$, that implies that $\Omega$ is a ball.
\end{proof}
\begin{theorem}[Stability] \label{thm stability by stability}
Let $\Omega\subset{\mathbb R^N}$ be a bounded domain with boundary $\Gamma$ of class $C^{2,\tau}$, $0<\tau<1$, and let $u\in C^{2,\tau}(\overline\Omega)$ be solution of \eqref{serrin1}.
Let $C$ be the constant in \eqref{stability t}.
\par
There are two concentric balls $B_{r_i}$ and $B_{r_e}$ such that \eqref{Bri-Omega-Bre}
holds with
\begin{equation}
\label{stability Serrin}
r_e-r_i\le 2C\, [u_\nu]_{\Gamma}.
\end{equation}
\par
In particular, if \eqref{serrin2} is in force with $\frak a>0$, then $\Omega$ is a ball.
\end{theorem}
\begin{proof}
The regularity of $u$ and \eqref{seminorms2} imply that \eqref{seminorm-ep} holds for some $\delta_0>0$. Thus, Theorem \ref{thm stability dependence on s} can be applied and, by \eqref{seminorms2}, we have that \eqref{Bri-Omega-Bre} holds with
$$
r_e-r_i\le 2 C \left\{ [u_\nu]_\Gamma+\int_0^{\delta} [u_{\nu\nu}]_{\Gamma^t}\,dt\right\},
$$
for every $\delta\in (0,\delta_0)$; \eqref{stability Serrin} then follows by letting $\delta$ tend to $0$.
\end{proof}
\begin{remark}
We notice that in Theorem \ref{thm stability by stability} we are not assuming the smallness of $[u_\nu]_{\Gamma}$ to prove \eqref{stability Serrin}.
\end{remark}
\section*{Acknowledgements}
The authors have been supported by the Gruppo Nazionale per l'Analisi Matematica, la Probabilit\`a e le loro
Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM).
The paper was completed while the first author was visiting \lq\lq The Institute for Computational Engineering and Sciences\rq\rq (ICES) of The University of Texas at Austin, and he wishes to thank the institute for the hospitality and support. The first author has been also supported by the NSF-DMS Grant 1361122 and the project FIRB 2013 ``Geometrical and Qualitative aspects of PDE''.
|
\section{Introduction}
Quantum walks have been a particularly fruitful field of research in
quantum information going back to ideas from Feynman \cite{Feynman},
Meyer \cite{Meyer_automata}, and Aharonov \cite{Aharonov}. One focus
of recent work
on quantum walks is their application to spatial search
algorithms. It is well-known that Grover's algorithm \cite{Grover1,
Grover2} for searching on unstructured databases offers a quadratic
improvement in search time over classical models. Grover's algorithm
is not designed to search physical systems where only local operations
are possible and quantum walk algorithms have been employed for this
purpose. In developing such algorithms, the major consideration is
the search time attempting to match the quadratic improvement over the
classical case offered by Grover's algorithm. The first spatial search
algorithm to do this, using the standard model of quantum walks, was
developed by Shenvi, Kempe and Whaley \cite{Shenvi} for searches on
a hypercube.
While there exist discrete-time searches on $d$-dimensional cubic
lattices which are faster than classical searches for $d\geq 2$
\cite{AKR} , effective continuous-time quantum searches only exist for
$d\geq 4$ \cite{CG04a} or else they require additional memory (in the
form of spin degrees of freedom) in order to improve their search time
in lower dimensions \cite{CG04b}. In \cite{FGT}, we demonstrated that
effective searches over two-dimensional lattices may be achieved in an
arguably simpler way which does not require extra degrees of freedom,
and could, therefore, be viewed as more efficient. This is achieved
through the choice of a different lattice, specifically, a honeycomb
lattice which is the underlying lattice structure of carbon atoms in
the material graphene.
The association with graphene is important as, although we first study
a purely theoretical problem in quantum information, the use of a
graphene lattice also offers a potential physical
realisation.
We thus envisage using the quantum walk and quantum search algorithm
framework to investigate the effect of perturbations on the dynamics
on graphene and other carbon structures. This offers not only the
possibility for demonstrating two-dimensional continuous-time quantum
searching, but also paves the way for looking for novel effects in the
material graphene.
In this paper, we offer a detailed account of the theory and numerical
results thus expanding on the findings as presented in \cite{FGT}.
Generalisations of the results in \cite{FGT} have been given in
\cite{Childscrystal}, where a approach to solving the problem
has been taken through encoding extra degrees of freedom into crystal
lattices. Recent experiments on artificial microwave
graphene \cite{Nice} have shown that additional site perturbations, as
discussed in Sec.\ \ref{sec:alternativemarking}, can
be used to create a search protocol. Searching on honeycomb
lattices in discrete-time setting has been considered in \cite{ADMP10};
however, no improvement over discrete time searches on cubic lattices
was found.
The
paper is structured as follows: In
Section~\ref{sec:quantumwalksandsearching} we give an introduction to
the formalism of continuous-time quantum walks and the construction of
quantum walk search algorithms. We also explain why previous search
algorithms struggled on lower-dimensional lattices and why graphene
offers a solution. Section~\ref{sec:graphene} will detail the relevant
properties of graphene and set-up the notation we shall use
throughout.
Sections~\ref{sec:threebondperturbation}~\&~\ref{sec:threebondperturbation_analysis}
contain our main analytical results, where we detail the specifics of
our search algorithm and offer an analysis of the search running time
and success probability. We shall then show in
Section~\ref{sec:communication} how this search protocol can be
adapted to demonstrate novel communication
setups. Sections~\ref{sec:alternativemarking}~\&~\ref{sec:alternativenanostructures}
contain numerical work demonstrating the possibility of using
alternative methods of marking to create search behaviour and other
carbon nanostructures. We conclude with a review and discussion of our
results in Section~\ref{sec:discussion}.
\section{Quantum walks and searching}
\label{sec:quantumwalksandsearching}
Continuous-time quantum walks (CTQW), first defined in \cite{Farhi},
are the quantum analogue of continuous-time Markov chains. They are
defined purely on the state space, that is, the Hilbert space
$\mathcal{H}_{p}$, spanned by the states $\ket{j}$ which represent the
$j^{th}$ site of the lattice. Thus, the time-evolution of such systems
is defined by the Sch\"{o}dinger equation
\begin{equation}\label{ctqw}
\frac{d}{dt}\alpha_{j}\left(t\right) = -i\sum_{l=1}^{N} {\bf H}_{jl}\alpha_{l}\left(t\right)\, ,
\end{equation}
where $\alpha_{j} = \braket{j}{\psi\left(t\right)}$ is the probability
amplitude at the $j^{\text{th}}$ vertex of a system described by the
state vector $\ket{\psi\left(t\right)}$, and $\bf H$ is the
Hamiltonian describing the connectivity of the lattice.
Note that we are using a dimensionless description setting e.g. $\hbar=1$.
The dynamics of a quantum walk over a network is defined by the nature
of the interaction between connected sites. Therefore, the Hamiltonian
is generally constructed from the adjacency matrix of the underlying
lattice. The adjacency matrix $\mathbf{A}$ is defined as
\begin{equation}
\mathbf{A}_{jl} = \left\{ \begin{array}{l l}
1 & \quad \text{if $j$ and $l$ are connected}\\
0 & \quad \text{if $j$ and $l$ are not connected.}
\end{array} \right.
\end{equation}
Typically, the Hamiltonian is chosen as $\mathbf{H} =
\epsilon_{D}\mathbf{I} + v\mathbf{A}$. The parameter $v$ determines
the coupling strength between connected sites and the parameter
$\epsilon_{D}$ is an on-site energy that only enters the dynamics in a
trivial way and thus can be set to a desired value. If $v=-1$ and $\epsilon_{D}$ is equal to the valency of the lattice $\mathbf{H}$
is the discrete Laplacian.
This form of Hamiltonian is closely related to the
tight-binding model for condensed matter systems \cite{Kittel}.
As first explained in \cite{HeinTanner09, HeinTannerCom}, a quantum
walk is transformed into a search protocol by introducing a localised
perturber state, forming an avoided crossing in the spectrum of the
search Hamiltonian between an unperturbed eigenstate and the localised
perturber. Thus, initialising the system in the unperturbed eigenstate
involved in the crossing and allowing the system to evolve in time,
one finds that the system rotates into the localised perturber state.
The first CTQW search over $d$-dimensional cubic lattices \cite{CG04a} introduced the localised perturber
state as a projector onto a single site $w$, resulting in the search
Hamiltonian
\begin{equation}\label{Hgamma}
\mathbf{H}_{\gamma} = -\gamma \mathbf{A} + \ket{w}\bra{w}\, .
\end{equation}
Here, $\gamma$ is a parameter governing the strength of
interactions between sites in the lattice. This parameter is chosen carefully \cite{CG04a} such that
the perturber state $\ket{w}$ is brought into resonance with the
ground state of the unperturbed Hamiltonian, the uniform superposition
$\ket{s}$.
Let us assume for the moment that
all other unperturbed eigenstates are energetically
sufficiently
separated from the ground state -- we will come back to this
assumption later. Then there will be an avoided
crossing of two eigenvalues corresponding to the perturber state
and the ground state.
Perturbation theory estimates that the energy splitting of these
resonant states is proportional to the overlap of the localised
perturber state $\ket{w}$ and the uniform ground state $\ket{s}$, which scales as
\begin{equation}
\Delta E \sim |\langle w | s\rangle | \sim N^{-1/2}\, ,
\end{equation}
and that the corresponding eigenstates are of the form $(\ket{s} \pm \ket{w})/\sqrt{2}$.
By preparing the
system initially in $\ket{s}$ and allowing it to evolve for a set
period of time $T=\pi / \Delta E \propto \sqrt{N}$ , a measurement of the system will result in the state
$\ket{w}$ being measured with high probability. This explains the
speed-up of the quantum walk search. The Grover algorithm and its
speed-up can be
understood in analogous terms and one can show that all other states
are indeed energetically very well separated.
In the present case this
separation only holds above a critical dimension $d_c=4$.
A simple argument for the critical dimension can be obtained by
comparing the scaling behaviour of the energy splitting $\Delta E \sim
N^{-1/2}$ with the energy separation of the first excited state from
the ground state in the
unperturbed lattice. For a cubic lattice we have a quadratic
dispersion relation which allows us to estimate the energy separation $$E(\underline{k}) - E_0 \sim
|\underline{k}|^2 \sim N^{-2/d}$$ (as $|\underline{k}|\sim N^{-1/d}$
for the first excited state in a $d$-dimensional lattice). For $d>4$
one then has $\Delta E/ (E(\underline{k}) - E_0) \to 0$ as $N \to
\infty$ and a detailed analysis indeed proves that the quantum walk
search works with optimal speed-up \cite{CG04a}. At the critical
dimension $d=4$ the two energy scales scale in the same way -- the
detailed analysis shows that the search still works but the dynamics
is more complicated due to the interference of excited states.
While there is still a speed-up it is only almost optimal; the
optimal search time $T \propto N^{1/2}$ gets multiplied with a
logarithmically
increasing factor.
For the
experimentally relevant regimes of 2- and 3-dimensional cubic
lattices, all states participate in the dynamics as any avoided
crossing gets dissolved completely as the number of sites grows.
As a result
no speed-up over classical searches is found.
The above estimate offers
a simple way how to reduce the critical dimension.
If one can construct a search around a uniformly distributed state at
an energy $E_0$ where the dispersion relation is conic (linear in
$|\underline{k}|$), then in $d$ dimensions $$E-E_0 \sim
|\underline{k}| \sim N^{1/d}$$
which results in a critical dimension $d_c=2$.
In \cite{CG04b} this was implemented using a
modified Dirac Hamiltonian. However, this requires the addition of a
spin degree of freedom, essentially a doubling of memory, with the
added complication that it is not immediately clear how such a system
would be physically realised.
Instead, our solution here is to change the lattice
from a square (cubic) to a graphene lattice. This change of
topology automatically implies the first step in our solution, as one
of the important electronic features of graphene is the conic
dispersion relation around the Dirac energy. This arises naturally
from the tight-binding descriptions of graphene. Note that
graphene has the critical dimension $d=2$. Indeed our construction as
presented in \cite{FGT} has an almost optimal speed-up with logarithmic corrections
that need to be evaluated in a detailed analysis that goes beyond the
simple perturbative description given above.
\section{Relevant properties of graphene}
\label{sec:graphene}
Graphene is a single layer of carbon atoms arranged in a honeycomb
lattice. The lattice is bipartite with two sublattices, labelled $A$
and $B$, and a unit cell containing two carbon atoms. The spatial and
reciprocal lattices are shown in Figure~\ref{fig:both_lattices}. The
primitive vectors describing the lattice are
$\underline{a}_{1\left(2\right)}$, such that the position of a unit
cell in the lattice is given by
$\underline{R}\left(\alpha,\beta\right) = \alpha a_{1} + \beta
a_{2}$. We use dimensionless units in space where the
distance between nearest neighbor sites is $a=1$.
The reciprocal lattice shows two important points, the Dirac
points $\underline{K}$ and $\underline{K}'$ at the two inequivalent
corners of the Brillouin zone.
\begin{figure}[t]
\centering
\includegraphics[width=0.45\linewidth]{graphene_lattice_vectors}
\hspace{0.2cm}
\includegraphics[width=0.45\linewidth]{reciprocal_lattice}
\label{fig:reciprocal_lattice}
\caption{Left: Graphene with lattice vectors $\underline{a}_{1/2}$,
translation vectors $\underline{\delta}_{i}$ and unit cell (dashed
lines).
Right: Reciprocal lattice with basis vectors
$\underline{b}_{1/2}$, symmetry points $\underline{\Gamma}$,
$\underline{K}$, $\underline{K}'$, $\underline{M}$ and first
Brillouin zone (hexagon).}
\label{fig:both_lattices}
\end{figure}
The energy spectrum of electrons in graphene was first derived by
Wallace \cite{Wallace_graphene} when considering the band structure of
graphite using a tight-binding Hamiltonian
\begin{equation}
\mathbf{H}= \epsilon_D\mathbf{1} + v \mathbf{A}
\end{equation}
where $\epsilon_D$ is the on-site energy (which we will identify as the energy
of the Dirac points) and
$v$ the hopping strength (both dimensionless in our setting).
The tight-binding model
for graphene and the derivation of the solution are well-known
\cite{Wong_nanotube, Neto_graphene} and give rise to the dispersion
relation
\begin{align}
\epsilon\left(\underline{k}\right) &= \epsilon_{D} \pm \label{eq:graphene_unperturbed_eigenenergies} \\
&v\sqrt{1 + 4\cos^{2}\left(\frac{k_{x}}{2}\right) +
4\cos\left(\frac{k_{x}}{2}\right)\cos\left(\frac{\sqrt{3}k_{y}}{2}\right)}\,
, \nonumber
\end{align}
shown in Figure~\ref{fig:infinite_graphene_lattice} for an infinite
graphene lattice.
\begin{figure}
\centering
\includegraphics[trim = 0mm 63mm 0mm 92mm,clip = true,
width=0.9\linewidth]{infinite_graphene_spectrum.pdf}
\caption{(Color online) Dispersion relation for infinite graphene
sheet ($\epsilon_{D}=0$ and $v=-1$).}
\label{fig:infinite_graphene_lattice}
\end{figure}
As there are two atoms per unit cell the spectrum has two branches,
the upper branch being the conduction band and the lower the valence
band, which meet at the corners of the Brillouin zone, the
$\underline{K}$-points. The energy at the $\underline{K}$-points is
$\epsilon_{D}$ which we name the Dirac energy. It is around these
points, that the behaviour of the spectrum is conical, that is,
\begin{equation}\epsilon\left(\underline{k}\right) \approx
\epsilon_{D} \pm v\frac{\sqrt{3}}{2}\sqrt{\delta k_{x}^{2} + \delta
k_{y}^{2}} = \epsilon_{D} \pm v\frac{\sqrt{3}}{2}\left|\delta
k\right|\, ,
\label{eq:linear_dispersion}
\end{equation}
with a reduced density of states, a necessary feature for the creation
of the search dynamics.
As the lattice possesses a translational symmetry the Hamiltonian can
be solved using linear superpositions of Bloch functions over both
sublattices. As a basis we use the orthonormal states
$\{\ket{\alpha,\beta}^{A},\ket{\alpha,\beta}^{B}\}$ to denote states
on either the $A$ or $B$ sublattice in the cell at position
$\underline{R}\left(\alpha,\beta\right)$. For the majority of what
follows (except in Section~\ref{sec:alternativenanostructures}), we
will focus on finite-sized lattices with assumed periodic boundary
conditions along the axes of both primitive vectors so that the
topology of our lattice is a torus; that is, our wavefunction is of
the general form $|\psi\rangle= \sum_{\alpha=1}^m\sum_{\beta=1}^n
\left(\psi_{\alpha,\beta}^A |\alpha,\beta\rangle^A + \allowbreak
\psi_{\alpha,\beta}^B |\alpha,\beta\rangle^B\right)$. Our boundary
conditions imply that the state vector must satisfy
$\psi^{A(B)}_{\alpha,\beta}=\psi^{A(B)}_{\alpha+m,\beta}=\psi^{A(B)}_{\alpha,\beta+n}$
where $m,n$ denote the period of the lattice. Thus, the wavefunctions
on the torus take the Bloch function form
\begin{align}
\ket{\psi} = &\sum_{\left(\alpha,\beta\right)} \left[ \frac{1}{\sqrt{N}}e^{i\underline{k}\cdot\underline{R}\left(\alpha,\beta\right)}\ket{\alpha,\beta}^{A} \right. \nonumber \\
&\left. + \frac{C\left(\underline{k}\right)}{\sqrt{N}}
e^{i\underline{k}\cdot\left(\underline{R}\left(\alpha,\beta\right)+\underline{\delta}_{1}\right)}\ket{\alpha,\beta}^{B}\right]
\, , \label{eq:wavefn}
\end{align}
where $N$ is the number of sites in the lattice, $\underline{k}$ is
the momentum, and $C\left(\underline{k}\right)$ is a relative phase
contribution dependent upon whether the state belongs to the
conduction or valence bands; it can be calculated by explicitly
working through the tight-binding model.
Application of the periodic boundary conditions results in the
following quantised momenta
\begin{equation}
k_{x} = \frac{2\pi p}{m}\, , \hspace{10mm} k_{y} = \frac{1}{\sqrt{3}}\left(\frac{4\pi q}{n} - k_{x}\right)\, ,
\label{eq:quantised_momenta}
\end{equation}
where $p \in \{0,1,\ldots m-1\}$ and $q \in \{0,1,\ldots n-1\}$. In
the following and whenever we consider quantum walk dynamics on a
torus, the number of cells in each direction is generally chosen to be
the same, that is, $m = n = \sqrt{\frac{N}{2}}$. This choice is
purely for simplifying the notation; alternative torus dimensions are
possible as are other choices of boundary conditions
corresponding to alternative carbon structures (e.g.\ nanotubes or a
graphene sheet) as will be demonstrated in
Section~\ref{sec:alternativenanostructures}. For our choice of torus
dimensions, we find there are momenta equal to the
$\underline{K}$-points and, consequently, eigenstates with energies
equal to the Dirac energy when both $m$ and $n$ are some multiple of
3.
In fact, using the quantised momenta in
Eq.~\eqref{eq:quantised_momenta} obtained for periodic boundary
conditions, we find that there are four degenerate eigenstates with an
energy that coincides exactly with the Dirac energy when $m$ and $n$
are both multiples of 3. These four states, known as the Dirac states,
can be constructed to live only on one of the sublattices, and are given by
\begin{align}\label{eq:K_state}
\ket{\underline{K}}^{A(B)} =& \sqrt{\frac{2}{N}} \sum_{\left(\alpha,
\beta\right)} e^{i\frac{2\pi}{3}\left(\alpha + 2\beta + 2\sigma
\right)}
\ket{\alpha,\beta}^{A(B)}\nonumber \\
\ket{\underline{K}'}^{A(B)} =&\sqrt{\frac{2}{N}} \sum_{\left(\alpha,
\beta\right)} e^{i\frac{2\pi}{3}\left(2\alpha +
\beta\right)}\ket{\alpha,\beta}^{A(B)}\, ,
\end{align}
where $\sigma = 0$ for states on the $A$-sublattice or $\sigma = 1$
for states on the $B$-sublattice.
\section{Quantum search on graphene - Reduced model}
\label{sec:threebondperturbation}
In this section we will describe how a site is marked and will derive
our optimal search starting state. Our approach will be to analyse the
system's spectrum and its dynamics in a reduced Hamiltonian model
involving only the relevant states from the avoided crossing. In the
next section we will then validate this approach with a more detailed
analysis, using the results obtained here as an initial guide.
As already established, we introduce the localised perturber state in
a region of the spectrum with a conic dispersion relation and thus a
low density of states. Simply altering the on-site energy in Eq.\
(\ref{Hgamma}) as done in \cite{CG04a} does, however, not work here.
As the on-site energy $\epsilon_{D}$ and the Dirac energy are equal,
one finds that the perturbation only interacts with the Dirac states
in the limit of zero perturbation strength, returning the unperturbed
lattice. Therefore, we choose an alternative perturbation method:
namely, we modify the coupling strength between the site we wish to
mark and its neighboring vertices. We focus in this section on
changing the coupling to all three neighboring vertices of a
particular site equally. Our choice of perturbation matrix $\bf{W}$ to
mark the $A$-type vertex $\left(\alpha_{o},\beta_{o}\right)^{A}$ is
then
\begin{equation}\label{eq:3bond_perturbation_matrix} {\bf{W}} =
\sqrt{3}\ket{\ell}\bra{\alpha_{o},\beta_{o}}^{A} +
\sqrt{3}\ket{\alpha_{o},\beta_{o}}^{A}\bra{\ell}\, ,
\end{equation}
where the state $\ket{\ell}$ is the symmetric superposition over the
three neighbors of the perturbed site
\begin{equation}
\ket{\ell} = \frac{1}{\sqrt{3}}\left(\ket{\alpha_{o}-1,\beta_{o}}^{B} + \ket{\alpha_{o}-1,\beta_{o}+1}^{B} + \ket{\alpha_{o},\beta_{o}}^{B} \right)\, .
\end{equation}
This leaves us with the search Hamiltonian
\begin{equation}
\bf{H}_{\gamma} = -\gamma \bf{A} + \bf{W}\, ,
\label{eq:full_search_Hamiltonian}
\end{equation}
where $\gamma$ is a free parameter. In what follows, we always set the
on-site energy $\epsilon_{D}=0$.
Considering our search Hamiltonian, we can see that setting $\gamma =
1$ corresponds to a coupling strength of $v=0$ from the perturbed
vertex and its nearest-neighbors; our perturbation essentially
removes the marked vertex $\left(\alpha_{o},\beta_{o}\right)^{A}$ from
the lattice. Note that vacancies are a common, naturally occurring,
defects in graphene lattices \cite{Meyer}.
In order to establish the critical value of $\gamma$, we numerically
calculate the spectrum of $\bf{H}_{\gamma}$ as a function of $\gamma$,
plotted in Figure~\ref{fig:3bond_spectrum} for a torus of dimensions
$m = n = 12$. As $\bf{W}$ is a rank-2 perturbation, we see in
Figure~\ref{fig:3bond_spectrum} two perturber states working their way
through the spectrum to an avoided crossing around $\epsilon_{D} = 0$
when $\gamma = 1$, that is, when the perturbed vertex is removed from
the underlying lattice.
\begin{figure}
\begin{overpic}[width =
1.05\linewidth]{12x12_3p_spectrum_colour}
\put(54.25,26.75){\fbox{\includegraphics[width=0.37\linewidth]{3bond_gap_scale}}}
\end{overpic}
\caption{(Color online) Spectrum of $ {\bf H}_\gamma$ in
Equation~\eqref{eq:full_search_Hamiltonian} as a function of
$\gamma$ for a $12\times 12$ cell torus ($N=288$). The spectrum is
symmetric around $\epsilon_{D} = 0$. Inset: Scaling of the gap
$\Delta = \tilde{E}_{+} - \tilde{E}_{-}$ (dots) and curves
$c_{1}/\sqrt{N}$ (solid blue), $c_{2}/\sqrt{N\log N}$ (dashed red)
for comparison. }
\label{fig:3bond_spectrum}
\end{figure}
As well as establishing a critical value for $\gamma$, we also find
from this figure the states involved in the avoided crossing: the four
Dirac states at the Dirac energy and the two perturber states which
form our perturbation matrix $\bf{W}$. However, with further
consideration, we may reduce the number of states involved further. As
we have removed the site $\left(\alpha_{0},\beta_{0}\right)^{A}$ from
the lattice it can no longer interact with the rest of the lattice and
the corresponding state drops out. Also, by direct calculation one
finds that the $B$-type Dirac states do not interact with the
perturbation, that is, ${\bf{W}}\ket{\underline{K}}^{B} =
{\bf{W}}\ket{\underline{K}'}^{B} = 0$. Thus, they remain an eigenstate
of the search Hamiltonian and do not interact with the perturbation.
We are left with three states taking part in the avoided crossing:
$\{\ket{\underline{K}}^{A},\ket{\underline{K}'}^{A},\ket{\ell}\}$.
We reduce our search Hamiltonian in this three state basis at the
critical point $\gamma = 1$ to obtain the following reduced
Hamiltonian describing the local dynamics at the avoided crossing
\begin{equation}\label{redH}
\tilde{\bf H} = \sqrt{\frac{6}{N}} \begin{bmatrix}
0 & 0 & e^{-i\mu_o} \\
0 & 0 & e^{-i\nu_o} \\
e^{i\mu_o} & e^{i\nu_o} & 0
\end{bmatrix}\,
\end{equation}
with $\mu_o = \frac{2\pi}{3}\left(\alpha_{o} + 2\beta_{o}\right)$ and
$\nu_o = \frac{2\pi}{3}\left(2\alpha_{o} + \beta_{o}\right)$. This
reduced Hamiltonian has eigenvalues $ \tilde{E}_{\pm} = \pm
2\sqrt{\frac{3}{N}}, \tilde{E}_{0} = 0$, and eigenvectors
\begin{align} \label{eq:reduced_matrix_eigenvectors}
\ket{\tilde{\psi}_{\pm}} &= \frac{1}{2}\left(e^{-i\mu_o}\ket{K}^A + e^{-i\nu_o}\ket{K'}^A \pm \sqrt{2}\ket{\ell}\right) \\
\ket{\tilde{\psi}_{0}} &=
\frac{1}{\sqrt{2}}\left(e^{-i\mu_o}\ket{K}^A - e^{-i\nu_o}\ket{K'}^A
\right)\, .
\end{align}
Using the eigenvectors of the reduced Hamiltonian, we can construct a
search starting state which is a superposition of Dirac states
\begin{align}
\ket{s} &= \frac{1}{\sqrt{2}}\left(\ket{\tilde{\psi}_{+}} + \ket{\tilde{\psi}_{-}}\right) \label{eq:starting_state} \\
&= \frac{e^{-i\mu_o}}{\sqrt{2}}\left(\ket{\underline{K}}^{A} +
e^{-i\frac{2\pi}{3}\left(\alpha_{o} -
\beta_{o}\right)}\ket{\underline{K}'}^{A}\right)\, . \nonumber
\end{align}
Allowing our search starting state $\ket{s}$ to evolve under the
reduced Hamiltonian we find
\begin{align}
\ket{\psi\left(t\right)} &= e^{-i{\bf{\tilde{H}}}t}\ket{s} \nonumber \\
&= \frac{1}{\sqrt{2}}\left(e^{-i \tilde{E}_{+}t}\ket{\tilde{\psi}_{+}} + e^{-i \tilde{E}_{-}t}\ket{\tilde{\psi}_{-}}\right) \label{eq:search_time}\\
&= \cos\left(\tilde{E}_{+}t\right)\ket{s} -
i\sin\left(\tilde{E}_{+}t\right)\ket{\ell} \nonumber\, ,
\end{align}
so that our system rotates from our Dirac superposition $\ket{s}$ to a
state which is localised on the neighbors of the perturbed vertex
$\ket{\ell}$ in a time $t=\frac{\pi}{4}\sqrt{\frac{N}{3}}$. Thus, we
find a $\mathcal{O}\left(\sqrt{N}\right)$ search time, a polynomial
improvement over the classical search time.
\begin{figure}[t]
\centering
\includegraphics[width= 1.0\linewidth]{12x12cell_walk_3p}
\caption{(Color online) Search on $12\times 12$ cell graphene lattice with a
triple-bond perturbation, using starting state $\ket{s}$ from
Equation~\ref{eq:starting_state}. For tori with $m = n$ the
dynamics at each neighboring site is the same so only one is
shown.}
\label{fig:search_prob}
\end{figure}
We plot in Figure~\ref{fig:search_prob} a numerically calculated
search for a lattice with $N = 288$ sites. The system has been
prepared in $\ket{s}$ and allowed to evolve under the full search
Hamiltonian from Eq.~(\ref{eq:full_search_Hamiltonian}) at the
critical value $\gamma = 1$. One finds that the system localises on
the neighbors of the marked vertex and, for this particular system
size, the probability for a measurement of the system to return one of
the neighbors is around $45\%$ at its peak. This is orders of
magnitude higher than the average probability to measure other
vertices in the lattice, $100/N$, which for this system size is around
$0.35\%$. The localised state interacting with the spectrum under the
full Hamiltonian has a tail into the rest of the lattice; this leads
to a loss of probability to be found at the neighboring vertices
below $100\%$.
It is worth emphasising that the marked vertex plays no role in the
search - it is removed from the lattice. Rather, we force the system
to localise on the nearest-neighbors, essentially making the marked
vertex conspicuous by its absence. This feature differentiates our
algorithm from previous searches, which have generally concentrated on
localising directly on the marked vertex. An approach considering the
neighboring vertices was detailed in \cite{Ambainisneighbors},
however, the focus was here still on localising on a single site but
with classical post-processing steps considering the tail of the
localised state extending into the neighborhood around the marked
vertex.
At this point, there are a number of issues which need to be
mentioned. The first is that our starting state in
Eq.~(\ref{eq:starting_state}) contains information about the marked
vertex in the form of the relative phase between the Dirac states.
However, this phase can only take three different values. Therefore, we have
three possible optimal starting states for an $A$-type perturbation and the
same number for $B$-type perturbations; there are thus in total six
possible optimal states. As not all of these states are orthogonal,
we only find an increase of necessary runs for a successful search by
a factor of 4; this additional overhead is independent of $N$ and,
therefore, does not affect the overall time complexity of the search.
The particular representation of the Dirac states in
Eq.~(\ref{eq:K_state}) was chosen in such a way as to make the
calculations and conceptualisation of the system dynamics easier. In
reality, constructing starting states which exist on
one sublattice only, experimentally is extremely difficulty; rather, in an experiment,
one is likely to excite a superposition of Dirac states, reducing the
success probability, on average, by a factor of $1/4$.
Our search is based on the conic dispersion relation in the spectrum
and the $\mathcal{O}\left(1/\sqrt{N}\right)$ scaling between
successive energy levels in the linear regime, giving rise to the
$\mathcal{O}\left(\sqrt{N}\right)$ search time found in our reduced
model. However, previous searches in continuous-time using a modified
Dirac Hamiltonian \cite{CG04b} or in discrete-time \cite{AKR} have
found logarithmic corrections to the search time. As we have seen in
Eq.~(\ref{eq:search_time}) the search time is related to the energy
gap at the avoided crossing. In the inset of
Figure~\ref{fig:3bond_spectrum} we have numerically calculated the
scaling of the gap in the spectrum of the full search Hamiltonian and
we indeed find evidence for a logarithmic correction. The lack of a
logarithmic term in the search time for our reduced model is due to
the neglecting contributions from the rest of the spectrum.
In what follows we establish a more accurate estimate of the running
time and success probability by working with the full Hamiltonian. We
also justify the findings of our reduced model: while this model is
insufficient to estimate the finer details, such as logarithmic correction terms,
the accurate calculations show that our search
algorithm indeed takes place mainly in a two-dimensional subspace
spanned by the Dirac states and a state localised on the neighbors of
the marked vertex.
\section{Quantum search on graphene - Detailed analysis}
\label{sec:threebondperturbation_analysis}
We follow here closely our derivation as presented in \cite{FGT} which builds on ideas given in \cite{CG04a} for a similar
a similar calculation for regular rectangular lattices. We focus first on the search success amplitude, that is,
\begin{equation}
\bra{\ell}e^{-i{\bf{H}}T}\ket{start} = \sum_{\ket{\psi_{a}}
} \braket{\ell}{\psi_{a}}\braket{\psi_a}{start}e^{-iE_{a}T},
\label{eq:amplitude}
\end{equation}
where $T$ is the search time, that is, the time our search probability
reaches a maximum, and our search Hamiltonian ${\bf H}$ is given by
Eq.\ (\ref {eq:full_search_Hamiltonian}) with $\gamma = 1$, that is,%
\begin{equation} {\bf{H}} = -{\bf{A}} +
\sqrt{3}\ket{\ell}\bra{\alpha_{o},\beta_{o}} +
\sqrt{3}\ket{\alpha_{o},\beta_{o}}\bra{\ell}\, ;
\label{eq:specific_search_Hamiltonian}
\end{equation}
here, $\ket{\psi_a}$, $E_{a}$ are the eigenstates and eigenenergies of
${\bf H}$. We assume, without loss of generality, a specific
starting state $\ket{start}$ where the marked vertex is chosen such
that $e^{i\frac{2\pi}{3}\left(\alpha_{o}+2\beta_{o}\right)} = 1$
leading to
\begin{equation}
\ket{start} = \frac{1}{\sqrt{2}}\left(\ket{\underline{K}} + \ket{\underline{K}'}\right)\, .
\end{equation}
In the following, we suppress the sublattice superscript as the
analysis is the same regardless on which sublattice the perturbation
lives. We also denote $\epsilon\left(\underline{k}\right)$ the positive
eigenenergies of $-\bf{A}$ from Eq.~(\ref{eq:graphene_unperturbed_eigenenergies})
and set $\epsilon_{D} = 0$.
For eigenstates $\ket{\psi_{a}}$ with eigenenergies $E_a$
that are not in the unperturbed graphene spectrum ($E_a \neq
\epsilon(\underline{k})$ for all points $\underline{k}$ in the dual lattice),
we may rewrite
Eq.~(\ref{eq:specific_search_Hamiltonian}) in the form
\begin{equation}
\ket{\psi_{a}} = \sqrt{3 R_{a}}(E_{a} + {\bf{A}})^{-1} \ket{\alpha_{o},\beta_{o}}\, ,
\label{eq:eigenstate_representation}
\end{equation}
where $\sqrt{R_{a}} = \braket{\ell}{\psi_{a}}$ and the phase of
$\ket{\psi_{a}}$ is chosen such that $\braket{\ell}{\psi_{a}} \geq 0$.
At this point, we can remove several states from the summation in
Eq.~(\ref{eq:amplitude}). Note that the basis state associated with
the marked vertex, $\ket{\alpha_{o},\beta_{o}}$, is itself an
eigenstate of the search Hamiltonian with ${\bf
H}\ket{\alpha_{o},\beta_{o}} = 0$, and also,
$\braket{\ell}{\alpha_{o},\beta_{o}} = 0$ so that
$\ket{\alpha_{o},\beta_{o}}$ does not contribute to the
time-evolution. We may also remove eigenstates of $\bf{H}$ which are
at the same time eigenstates of $-\bf{A}$ with the same energy.
This can be seen in the following way. We first consider an
unperturbed eigenstate $\ket{\psi^{o}_{a}}$ such that
$-\mathbf{A}\ket{\psi^{o}_{a}} = E_{a}\ket{\psi^{o}_{a}}$. Let us
assume that there is an eigenvector $\ket{\psi_{a}}$ of the search
Hamiltonian with the same eigenenergy $E_{a}$, that is,
$\mathbf{H}\ket{\psi_{a}} = E_{a}\ket{\psi_{a}}$. Considering the
matrix element $\bra{\psi^{o}_{a}}\mathbf{H}\ket{\psi_{a}}$, we find
$\braket{\psi^{o}_{a} }{ \ell}\braket{\alpha_{o},\beta_{o} }{ \psi_{a}
} + \braket{\psi^{o}_{a} }{ \alpha_{o},\beta_{o}}\braket{\ell }{
\psi_{a} } = 0$. As $\ket{\alpha_{o},\beta_{o}}$ is an eigenvector
of the search Hamiltonian we know that $\braket{\alpha_{o},\beta_{o}
}{ \psi_{a} } = 0$. This leaves us with $\braket{\psi^{o}_{a} }{
\alpha_{o},\beta_{o}}\braket{\ell }{ \psi_{a} } = 0$. As the
unperturbed eigenstate $\ket{\psi^{o}_{a}}$ is simply a Bloch state we
know $\braket{\psi^{o}_{a}}{\alpha_{o},\beta_{o}} \neq 0$. Thus, we
obtain $\braket{\ell }{ \psi_{a}} = \sqrt{R_{a}} = 0$. It is then
clear that eigenstates of the search Hamiltonian whose eigenenergies
remain in the spectrum of $-\mathbf{A}$ do not play a role in the
time-evolution of the search.
We now derive an eigenvalue condition for those perturbed eigenvalues
which are in the spectrum of $\bf{H}$. Using the orthogonality of
eigenstates $\braket{\alpha_{o},\beta_{o}}{\psi_{a}}=0$,
Eq.~(\ref{eq:eigenstate_representation}) leads to
\begin{equation}
\sqrt{3R_{a}}\bra{\alpha_{o},\beta_{o}}\left(E_{a} + {\bf{A}} \right)^{-1}\ket{\alpha_{o},\beta_{o}} = 0\,.
\end{equation}
By expressing $\bra{\alpha_o,\beta_o}$ in terms of the eigenstates of
the {\em unperturbed } walk Hamiltonian $-{\bf{A}}$, we may write this
as a quantization condition
\begin{equation} \label{FE} F\left(E_{a}\right) = 0
\end{equation}
with
\begin{equation} \label{FE}
F\left(E\right) = \frac{\sqrt{3}}{N}
\sum_{\underline{k}}
\left[
\frac{1}{E - \epsilon\left(\underline{k}\right)}
+
\frac{1}{E + \epsilon\left(\underline{k}\right)}
\right]\, ,
\end{equation}
were $N$ is the total number of sites in the lattice.
Eq.\ (\ref{FE}) is written as two summations to
incorporate the fact that the spectrum of $-\bf{A}$ as well as
$\bf{H}$ is symmetric around $E=0$.
Choosing $\ket{\psi_{a}}$ to be normalised
$\braket{\psi_{a}}{\psi_{a}}=1$,
Eq.~(\ref{eq:eigenstate_representation}) also implies
\begin{equation}
3 R_{a}\bra{\alpha_{o},\beta_{o}}\left(E_{a} + {\bf A}\right)^{-2}
\ket{\alpha_{o},\beta_{o}} = 1,
\end{equation}
which allows $R_{a}$ to be rewritten as
\begin{equation}
R_{a} = \frac{1}{\sqrt{3}|{F'\left(E_{a}\right)}|}.
\label{eq:Ra}
\end{equation}
We may now rewrite the amplitude in Eq.~\eqref{eq:amplitude} in the
form
\begin{equation}
\bra{\ell}e^{-i{\bf H}T}\ket{start} = \braket{\alpha_{o},\beta_{o}}{start}\sum_{a}\frac{e^{-iE_{a}T}}{E_{a}|{F'\left(E_{a}\right)}|}\, ,
\label{eq:time_evolution}
\end{equation}
where we have used the adjoint of
Eq.~(\ref{eq:eigenstate_representation}). Note again that eigenstates
of $\bf{H}$, which are also in the spectrum of $-\bf{A}$, do not
contribute to the time-evolution of the search; it may be seen from
the definition of $F\left(E\right)$ in Eq.~(\ref{FE}) that $\left|
F'\left(E_{a}\right) \right| \rightarrow \infty$, where $E_{a}$ is
in the spectrum of $-\bf{A}$.
As we saw in the previous section, the avoided crossing is formed by
two perturbers approaching symmetrically either side of the Dirac
states with energy $\epsilon_{D} = 0$. Thus, we concentrate on
evaluating the contribution to the time-evolution from these perturbed
eigenstates of $\bf{H}$ either side of the Dirac point and we label
these states $\ket{\psi_{\pm}}$. In order to evaluate these
contributions we calculate the perturbed eigenenergy $E_{+}$ and also
derive a leading-order expression for $F'\left(E_{+}\right)$ (since
the spectrum is symmetric, we have $E_{+} = -E_{-} > 0$).
Using the definition of $F\left(E\right)$ in Eq.~(\ref{FE}), we
estimate $F\left(E_{+}\right)$ by separating out the Dirac points,
where
$\epsilon\left(\underline{K}\right)=\epsilon\left(\underline{K}'\right)
= 0$, from the summations and then Taylor expanding the remaining
terms at $E=0$, to find
\begin{equation}\label{eq:exp}
F\left(E_{+}\right) = \frac{4\sqrt{3}}{NE_{+}} - \sum_{n=1}^{\infty} I_{2n}E_{+}^{2n-1}\, ,
\end{equation}
where the sums $I_{n}$ are given by
\begin{equation}
I_{n} = \frac{\sqrt{3}}{N}\sum_{\underline{k} \neq \underline{K},\underline{K}'}
\left[
\frac{1}{\left[\epsilon\left(\underline{k}\right)\right]^{n}}
+
\frac{1}{\left[-\epsilon\left(\underline{k}\right)\right]^{n}}
\right]
\,.
\end{equation}
As the unperturbed spectrum is symmetric only those $I_{n}$ with even
$n$ are non-zero; thus from now on we focus only on $I_{2k}$ where $k\geq
1$.
We stated earlier that the spectrum of graphene is well-approximated
by a conic dispersion relation around the Dirac points. Thus, it is
from around these points that the major contributions to the $I_{2k}$
sums arise. By applying the linear approximation from
Eq.~(\ref{eq:linear_dispersion}) and the momenta quantum numbers from
Eq.~(\ref{eq:quantised_momenta}) we may approximate the $I_{2k}$
summations as
\begin{eqnarray}
I_{2k}&= &4\sqrt{3} N^{k-1} \left[
\sum_{(p,q)\in L}
Z_{2}\left(S_{\underline{K}},2\right)
+ \sum_{(p,q)\in L'}
Z_{2}\left(S_{\underline{K}'},2\right)
\right] \nonumber \\
&+& O(1) \, .
\label{eq:I_epstein}
\end{eqnarray}
Here $Z_2(S,x)$ is the Epstein zeta-function \cite{Siegel}
\begin{equation}
Z_2(S,x)
= \frac{1}{2}
\sum_{ (p,q) \in \mathbb{Z}^2
\backslash (0,0)}
\left( S_{11}p^2
+2 S_{12} pq + S_{22} q^2
\right)^{-x}\, ,
\end{equation}
for a real positive definite real symmetric $2 \times 2$ matrix
$S$. For our purposes we use
\begin{equation}
S_{\underline{K}}= S_{\underline{K}'}=
4\pi^2 \begin{pmatrix}
2 & -1 \\
-1 & 2
\end{pmatrix} \, ,
\end{equation}
which describes the spectrum close to the Dirac points. The linear
approximation around both Dirac points $K$ and $K'$ is the same and,
thus, the matrices $S_{\underline{K}}$ and $S_{\underline{K}'}$ are
equal.
Our summation in Eq.~(\ref{eq:I_epstein}) is over the rectangular
regions $L$ and $L'$ of the lattice $\mathbb{Z}^2$. Both are centered
on $(0,0)$ and have side lengths proportional to $\sqrt{N}$, however
the center, $(0,0)$, corresponding to the relevant Dirac point, is
omitted from the summation.
Convergence of the Epstein zeta function is well-known for $k\geq 2$
\cite{Siegel}, and leads to the bounds
\begin{equation}\label{eq:Ikest}
\lim_{N \to \infty } \frac{I_{2k}}{N^{k-1}}=
4 \sqrt{3} \left(Z_2(S_{\underline{K}},k)+ Z_2(S_{\underline{K}'},k)
\right)
\end{equation}
for $k\ge 2$.
A sharp estimate for
\begin{equation}\label{eq:I2est}
I_2= O \left(\ln N\right)\, .
\end{equation}
is given in Appendix~\ref{appendix_a}. In order to calculate an
estimate for $E_{+}$, we truncate the Taylor expansion of
$F\left(E\right)$ in Eq.~(\ref{eq:exp}) at the $I_{2}$ term (the first
term in the summation), and apply the eigenvalue condition
$F\left(E_{+}\right) = 0$. That is, we solve $\frac{4 \sqrt{3}}{N
E_+} - I_2 E_+=0$ and obtain the approximation $E_+^2\approx \frac{4
\sqrt{3}}{N I_2}$. This solution also leads to the estimate
$F'\left(E_{+}\right) \approx -2 I_{2}$.
We consider next whether our solution for $E_{+}$ lies inside the
radius of convergence of the Taylor expansion. We note that each term
in $Z_{2}\left(S_{\underline{K}},k\right)$ is smaller than the
corresponding term in $Z_{2}\left(S_{\underline{K}},2\right)$ for
$k>2$, and so it follows that $Z_2(S_{\underline{K}},k)<
Z_2(S_{\underline{K}},2)$ for $k>2$. This property of the Epstein
zeta function and the estimate from Eq.~(\ref{eq:Ikest}) imply
\[\sum_{n=2}^\infty I_{2n} E_+^{2n-1}< \frac{C}{N
E_+}\sum_{n=2}^\infty (N E_+^2)^n\] i.e.\ the infinite sum in
Eq.~(\ref{eq:exp}) converges for $E_+< 1/\sqrt{N}$. Thus, for large
$N$, our solution lies within the radius of convergence of our Taylor
expansion in
Eq.~(\ref{eq:exp}).
In order to show that the dominant contributions to the search come
from $\ket{\psi_{\pm}}$, we need to establish the leading order error
term. All the $I_{2k}$ sums are positive so that, given the sign of
all the terms in the Taylor expansion in Eq.~(\ref{eq:exp}), the true
value of $E_{+} > 0$ has to be smaller than the estimate we have
obtained. Thus, we write the true value of $E_{+}$ as
\begin{equation}
E_+^2=\frac{4 \sqrt{3}}{N I_2}- \Delta >0\, ,
\end{equation}
with $\Delta> 0$. One may rewrite $F(E_+)=0$, using the true value of
$E_{+}$, to give
\begin{equation}
I_2 \Delta = \sum_{n=2}^\infty I_{2n} E_+^{2n}\, .
\end{equation}
We follow the same arguments as used for the calculation of the radius
of convergence to obtain an upper bound for the summation. This
together with the already established fact that $E_+$ is inside this
radius for sufficiently large $N$, we get the following inequality:
\begin{eqnarray*}
0 < N I_2 \Delta &<& C \sum_{n=2}^\infty
(N E_+^2)^n\\
&=& \frac{C N^2 E_+^4}{1- N E_+^2}= O(I_2^{-2}).
\end{eqnarray*}
So $\Delta = O(I_2^{-3}N^{-1})= O\left((\ln N)^{-3}
N^{-1}\right)$. Thus, we obtain
\begin{equation}
E_+^2 = \frac{4 \sqrt{3}}{N I_2}\left( 1 + O({(\ln N)^{-2}})\right)\, .
\end{equation}
It also follows that
\begin{equation}
F'\left(E_{\pm}\right) = -2I_{2} + O\left(\frac{1}{\ln N}\right).
\end{equation}
This shows that our perturbed eigenstates $\ket{\psi_{\pm}}$ have an
$O\left(1\right)$-overlap with the starting state and are, therefore,
the relevant states to be considered in the time-evolution of the
algorithm. Using the definitions of $\ket{\psi_{a}}$ and $R_{a}$ in
Eqs.~(\ref{eq:eigenstate_representation}), (\ref{eq:Ra}), the inner
product of the starting state and the perturbed eigenvectors can be
expressed as
\begin{equation}
\braket{start}{\psi_{a}} = \frac{1}{E_a}\sqrt{ \frac{\sqrt{3}}{|F'\left(E_{a}\right)|}}\braket{start}{\alpha_{o},\beta_{o}}\, ,
\end{equation}
where $\braket{start}{\alpha_{o},\beta_{o}}$ is the overlap of the
starting state with the marked vertex state.
Applying our previous approximations for $E_{+}$ and
$F'\left(E_{+}\right)$, that is, for our perturbed eigenstates closest
to the Dirac point, we find
\begin{equation}
|\braket{start}{\psi_{\pm}}| = \frac{1}{\sqrt{2}} + O\left(\frac{1}{\ln^{2} N}\right).
\end{equation}
Thus, our starting state is indeed a superposition of the perturbed
eigenstates, $\ket{\psi_{\pm}}$ as assumed in the previous section.
This makes it possible to investigate the running time and success
amplitude of the algorithm in more detail, that is,
\begin{align}
&\left|\bra{\ell}e^{-iHt}\ket{start}\right| \\
&\approx \left|\frac{1}{\sqrt{2}}\left(e^{-i E_{+}t}\braket{\ell}{\psi_{+}} - e^{i E_{+}t}\braket{\ell}{\psi_{-}}\right)\right| \\
&=
\frac{1}{3^{\frac{1}{4}}I_{2}^{\frac{1}{2}}}\left|\sin\left(E_{+}t\right)\right|\,
.
\end{align}
It is clear from our earlier results for $E_{\pm}$ and $I_{2}$ that
our algorithm localises on the neighbor state $\ket{\ell}$ in time $T
= \frac{\pi}{2 E_{+}} = O\left(\sqrt{N\ln N}\right)$ with probability
amplitude $O\left(1/\sqrt{\ln N}\right)$. This confirms the
logarithmic correction observed numerically and displayed in the inset
of Fig.\ \ref{fig:3bond_spectrum}.
\section{Communication}
\label{sec:communication}
It has been demonstrated in \cite{HeinTannerCom} that discrete-time
search algorithms can be modified to create a communication protocol.
We show here that a communication setup can be established also in the
continuous-time search algorithm with minor changes due to the
subtleties of our search.
We use the same unperturbed walk Hamiltonian as before, that is,
${\bf{H_{0}}} = -\bf{A}$ and the same type of perturbation matrix as
in Eq.~(\ref{eq:3bond_perturbation_matrix}), but now at two different
sites in the
lattice.
Explicitly, our communication Hamiltonian is
\begin{equation} {\bf{H}} = -{\bf{A}} + {\bf{W_{s}}} + {\bf{W_{t}}}\,
,
\label{eq:communication_hamiltonian}
\end{equation}
where the perturbation matrices $\bf{W_{s/t}}$ mark the source and
target sites, respectively.
As prescribed in \cite{HeinTannerCom}, the communication protocol
operates by preparing a state localised on the source perturbation, in
our case on the neighbors of the source site, and allowing the system
to evolve under our communication Hamiltonian. The system then
localises on the neighboring vertices of the target site. The effect
of the two perturbation matrices $\bf{W_{s/t}}$ amounts to
disconnecting the two vertices from the underlying lattice.
As noted before, our search algorithm has several different optimal
starting states depending on the position of the marked
vertices. Therefore, we divide our communication analysis into three
different cases: the two sites i) are on the same sublattice and have
the same optimal search starting state; ii) are on the same sublattice
but \emph{do not} share the same optimal search starting state; iii)
are on different sublattices. Both of the first two cases can be
treated by applying the reduced Hamiltonian method demonstrated
earlier, but communication between different sublattices is not
tractable by this method and so we focus on numerics in this case.
For signal transfer between two sites on the same sublattice, we will
assume, without loss of generality, that our communication takes place
between two sites on the $A$-sublattice,
$\left(\alpha_{s/t},\beta_{s/t}\right)^{A}$; here, the subscript $s$
or $t$ denotes the source or target vertices respectively. Using
arguments as employed in Sec.~\ref{sec:threebondperturbation}, we can
reduce the number of relevant states. Ultimately, the search dynamics
takes place in the subspace spanned by the basis
$\ket{\underline{K}}^{A},\allowbreak\ket{\underline{K}'}^{A}$, the
$A$-type Dirac states, and $\ket{\ell_{s}},\allowbreak\ket{\ell_{t}}$,
the uniform superpositions over the neighbors of the source and
target vertices. This basis leads to the reduced Hamiltonian
\begin{equation}\label{eq:reduced_communication_hamiltonian} {\bf
\tilde{H}} = \sqrt{\frac{6}{N}}
\begin{pmatrix}
0 & 0 & e^{i\mu_{s}} & e^{i\mu_{t}} \\
0 & 0 & e^{i\nu_{s}} & e^{i\nu_{t}} \\
e^{-i\mu_{s}} & e^{-i\nu_{s}} & 0 & 0 \\
e^{-i\mu_{t}} & e^{-i\nu_{t}} & 0 & 0
\end{pmatrix}\, ,
\end{equation}
where $\mu_{s/t} \equiv \frac{2\pi}{3}\left(\alpha_{s/t} +
2\beta_{s/t}\right)$ and $\nu_{s/t} \equiv
\frac{2\pi}{3}\left(2\alpha_{s/t} + \beta_{s/t}\right)$.
In the first case considered we assume that there are two
perturbations on the graphene lattice located at the points
$\left(\alpha_{s},\beta_{s}\right)^{A}$ and
$\left(\alpha_{t},\beta_{t}\right)^{A}$, chosen such that
$e^{i\frac{2\pi}{3}\left(\alpha_{s} + 2\beta_{s}\right)} =
e^{i\frac{2\pi}{3}\left(\alpha_{t} + 2\beta_{t}\right)}$. This
implies that a search for either vertex, using our search algorithm
from
Secs.~\ref{sec:threebondperturbation}~\&~\ref{sec:threebondperturbation_analysis},
would use the same optimal starting state. As the phases are equal, we
drop the subscript on the perturbation coordinates in what follows.
Fixing our phases and diagonalising the reduced Hamiltonian, we find
it has eigenvalues $\lambda^{\pm}_{2} = \pm 2\sqrt{\frac{6}{N}}$ and
$\lambda^{1,2}_{0} = 0$ with eigenvectors
\begin{align}
\ket{\tilde{\psi}_{\pm 2}} &= \frac{1}{2}\left(e^{i\mu}\ket{\underline{K}}^{A} + e^{i\nu}\ket{\underline{K}'}^{A} \pm \ket{\ell_{s}} \pm \ket{\ell_{t}}\right) \\
\ket{\tilde{\psi}_{0}^{1}} &= \frac{1}{\sqrt{2}}\left(e^{i\mu}\ket{\underline{K}}^{A} - e^{i\nu}\ket{\underline{K}'}^{A} \right) \\
\ket{\tilde{\psi}_{0}^{2}} &= \frac{1}{\sqrt{2}}\left(\ket{\ell_{s}}
- \ket{\ell_{t}}\right)\, .
\end{align}
Using these eigenstates, we may rewrite the source neighbor state as
\begin{equation}
\ket{\ell_{s}} = \frac{1}{2}\left(\ket{\tilde{\psi}_{2}} - \ket{\tilde{\psi}_{-2}}\right) + \frac{1}{\sqrt{2}}\ket{\tilde{\psi}_{0}^{2}}\, .
\end{equation}
Placing the system in the source state $\ket{\ell_{s}}$ and allowing
the system to evolve under the reduced Hamiltonian, one finds
\begin{align}
\ket{\psi\left(t\right)} &= e^{-i{\bf \tilde{H}}t}\ket{\ell_{s}} \\
&= \frac{1}{2}\left(e^{-i\lambda^{+}_{2}t}\ket{\tilde{\psi}_{2}} - e^{-i\lambda^{-}_{2}t}\ket{\tilde{\psi}_{-2}}\right) + \frac{1}{\sqrt{2}}\ket{\tilde{\psi}_{0}^{2}} \\
&= \frac{-i}{2}e^{i\mu}\sin\left(\lambda^{+}_{2} t\right)\left(\ket{\underline{K}}^{A} + e^{i\frac{2\pi}{3}\left(\alpha -\beta\right)}\ket{\underline{K}'}^{A}\right) \nonumber \\
& \hspace{5mm} + \frac{1}{2}\left(\cos\left(\lambda^{+}_{2} t\right)
+ 1\right)\ket{\ell_{s}} +
\frac{1}{2}\left(\cos\left(\lambda^{+}_{2} t\right) -
1\right)\ket{\ell_{t}}\, .
\end{align}
We note that, in the last line above, the term in the brackets
involving only the Dirac states is actually the optimal search
starting state for both perturbed vertices, defined in
Eq.~(\ref{eq:starting_state}). The system thus oscillates between the
states localised on the neighbors of the perturbed vertices,
$\ket{\ell_{s}}$ and $\ket{\ell_{t}}$, in a time $T =
\frac{\pi}{2}\sqrt{\frac{N}{6}}$, via their optimal search starting
state.
\begin{figure}[t]
\centering
\includegraphics[width =
1.0\linewidth]{equivalent_communication}
\caption{(Color online) Numerically calculated signal transfer on a $12\times 12$
cell graphene lattice between equivalent vertices, using the
communication Hamiltonian in
Eq.~(\ref{eq:communication_hamiltonian}). The system is
initialised in $\ket{\ell_{s}}$ and localises on
$\ket{\ell_{t}}$. Only the sum of probabilities to be found on the
neighbor vertices is shown.}
\label{fig:equivalent_communication_torus}
\end{figure}
Figure~\ref{fig:equivalent_communication_torus} shows the system
evolving under the full communication Hamiltonian,
Eq.~(\ref{eq:communication_hamiltonian}). The initial state used for
the time evolution shown in
Figure~\ref{fig:equivalent_communication_torus} is the true localised
state, that is, we run the quantum search with a single perturbation
located at vertex $\left(\alpha_{s},\beta_{s}\right)^{A}$ until it
reaches maximum success probability, and then apply the second
perturbation to the vertex
$\left(\alpha_{t},\beta_{t}\right)^{A}$. The figure confirms the
behaviour expected from the reduced model calculation.
The communication mechanism essentially works in the same way as the
quantum search algorithm, as it can be viewed as one marked vertex
`finding' another. The initial localised source state decays back
towards the search starting state, and the system then searches for
the target state.
In our second case of signal transfer we analyse the behaviour of a
communication system where we have two perturbed vertices on the same
sublattice, but with the restriction that
$e^{i\frac{2\pi}{3}\left(\alpha_{s} + 2\beta_{s}\right)} \neq
e^{i\frac{2\pi}{3}\left(\alpha_{t} + 2\beta_{t}\right)}$. As such,
the two marked sites cannot interact via the same search starting
state. However, the optimal search starting states are not orthogonal
so that signal transfer is still possible, but the resulting
interference effects make the analysis slightly more complicated.
We rewrite the coordinates of the target in terms of the source,
$\alpha_{t} = \alpha_{s} + x$ and $\beta_{t} = \beta_{s} + y$. Again,
fixing our phases and diagonalising the reduced Hamiltonian in
Eq.~(\ref{eq:reduced_communication_hamiltonian}), we find it has
eigenvalues $\lambda^{\pm}_{\sqrt{3}} = \pm\sqrt{3}\sqrt{\frac{6}{N}}$
and $\lambda^{\pm}_{1} = \pm \sqrt{\frac{6}{N}}$ with eigenvectors
\begin{align}
&\ket{\tilde{\psi}_{\pm\sqrt{3}}} = \frac{e^{i\mu_s}}{2\sqrt{3}}\left(e^{i\frac{2\pi}{3}\left(x + 2y\right)} - 1\right)\ket{\underline{K}}^{A} \\
&+ \frac{e^{i\nu_s}}{2\sqrt{3}}\left(e^{-i\frac{2\pi}{3}\left(x + 2y\right)} - 1\right)\ket{\underline{K}'}^{A} \nonumber \mp \frac{1}{2}\ket{\ell_{s}} \pm \frac{1}{2}\ket{\ell_{t}} \nonumber \\
&\ket{\tilde{\psi}_{\pm 1}} = \mp\frac{e^{i\mu_s}}{2}\left(e^{i\frac{2\pi}{3}\left(x + 2y\right)} + 1\right)\ket{\underline{K}}^{A} \\
&\pm \frac{e^{i\nu_s}e^{i\frac{2\pi}{3}\left(x +
2y\right)}}{2}\ket{\underline{K}'}^{A} -
\frac{1}{2}\ket{\ell_{s}} - \frac{1}{2}\ket{\ell_{t}}\, . \nonumber
\end{align}
Using these eigenstates, one may write the source perturbation as
\begin{equation}
\ket{\ell_{s}} = \frac{1}{2}\left(-\ket{\tilde{\psi}_{\sqrt{3}}} + \ket{\tilde{\psi}_{-\sqrt{3}}} - \ket{\tilde{\psi}_{1}} - \ket{\tilde{\psi}_{-1}}\right)\, .
\end{equation}
The full expression for the time-evolution is rather cumbersome;
therefore, we only show the terms and prefactors we are interested in,
namely $\ket{\ell_s}$ and $\ket{\ell_t}$
\begin{align}
\ket{\psi\left(t\right)} = &\frac{1}{2}\left[\cos\left(\lambda^{+}_{\sqrt{3}}t\right)+\cos\left(\lambda^{+}_{1}t\right)\right]\ket{\ell_{s}} \label{eq:simulated_nonequivalent_communication} \\
&-
\frac{1}{2}\left[\cos\left(\lambda^{+}_{\sqrt{3}}t\right)-\cos\left(\lambda^{+}_{1}t\right)\right]\ket{\ell_{t}}
\nonumber \, .
\end{align}
We can see here that the prefactors do not depend upon the coordinates
of either the source or the target sites; the transport signal between
sites on the same sublattice but with different optimal search
starting state is thus independent of the position of the source or
target site.
In Figure~\ref{fig:both_nonequivalent_communication} we show the
system evolved under the full communication Hamiltonian. Again, the
initial state is the true localised state on the nearest-neighbors of
the source vertex, obtained by running the search algorithm using one
marked vertex until it reaches its peak success probability. The
time-evolution is radically different to the previous communication
case; the behaviour here is erratic with uneven peaks of probability
at the two perturbations involved in the protocol. However, there are
still significant probability revivals.
\begin{figure}[t]
\centering
\includegraphics[width =
1.0\linewidth]{nonequivalent_communication}
\newline \centering
\includegraphics[width =
1.0\linewidth]{nonequivalent_communication_simulation}
\caption{(Color online) Upper: Numerically calculated signal transfer on a
$12\times 12$ cell graphene lattice between non-equivalent
vertices, using the communication Hamiltonian in
Eq.~(\ref{eq:communication_hamiltonian}). The system is
initialised in $\ket{\ell_{s}}$ and localises on
$\ket{\ell_{t}}$. Only the sum of probabilities to be found on the
neighbor vertices is shown. Lower: Analytically calculated
behaviour for the same system, using the reduced Hamiltonian
method and Eq.~(\ref{eq:simulated_nonequivalent_communication}).}
\label{fig:both_nonequivalent_communication}
\end{figure}
The transport behaviour from
Eq.~(\ref{eq:simulated_nonequivalent_communication}), calculated using
the reduced Hamiltonian, is also shown in
Figure~\ref{fig:both_nonequivalent_communication}. The probability at
time $t = 0$ has been scaled to match that shown in
Figure~\ref{fig:both_nonequivalent_communication}. Our calculated
behaviour has the same signal pattern as the numerically calculated
behaviour from the full Hamiltonian, although over a shorter period of
time. As our reduced model only makes use of the Dirac states and the
perturber states, we lose the contribution to the time-evolution from
the rest of the spectrum giving rise to logarithmic corrections as
discussed in Sec.~\ref{sec:threebondperturbation_analysis}. This leads
to differences in the overall time scales for the reduced model and
the full Hamiltonian. This also supports our findings that signal transfer
between all non-equivalent vertices on the same sublattice is the
same. The communication protocol is again set up by a
search mechanism in reverse, where one vertex finds another. The
slightly erratic behaviour that emerges here is due to interference
between the two separate search mechanisms which interact due to the
non-zero inner product of the three possible optimal search starting
states for vertices on the $A$-sublattice.
In our final case of signal transfer, we consider communication
between sites on different sublattices which we can not treat in the reduced
Hamiltonian model. It has been demonstrated previously that
perturbations to one sublattice do not interact with the Dirac states
which live on the other sublattice. Therefore, when attempting to
reduce the communication Hamiltonian as done before, we merely find
that it decouples into two non-interacting reduced Hamiltonians, each
describing a search protocol on one sublattice. The interaction
between the two sublattices is here facilitated due to interactions
via the bulk of the spectrum. In what follows, we focus on numerics
and inspect the behaviour for systems involving the same source
perturber but different targets. The numerics show a finite number of
different signal patterns, two examples are shown in
Figure~\ref{fig:both_differentsublattice_communication}. We can see
from these examples that the communication mechanism takes place over
a much longer timescale with a superimposed oscillatory dynamics.
\begin{figure}[t]
\centering
\includegraphics[width =
1.0\linewidth]{differentsublattice_communication_1}
\newline \centering
\includegraphics[width =
1.0\linewidth]{differentsublattice_communication_2}
\caption{(Color online) Numerically calculated signal transfer on a $12\times 12$
cell graphene lattice between vertices on different sublattices,
using the communication Hamiltonian in
Equation~\ref{eq:communication_hamiltonian}. The system is
initialised in $\ket{\ell_{s}}$ and localises on
$\ket{\ell_{t}}$. Only the sum of probabilities to be found on the
neighbor vertices is shown. The two figures have the same source
position but different positions of the target site.}
\label{fig:both_differentsublattice_communication}
\end{figure}
The increase in timescale can be attributed to the weak nature of the
interaction between the two perturbations due the fact that the
localised states live mainly on one sublattice. The signal pattern in
Figure~\ref{fig:both_differentsublattice_communication} thus decouples
into a fast oscillation between one site, say the source site, and the
delocalised lattice state and a slow time scale on which a small
amount of probability amplitude escapes into the other sublattice due
to the weak interaction. This process continues until the recurrence
probability at the target perturbation reaches the same peak as the
initial localised source state, and then the behaviour reverses.
\section{Alternative methods of marking}
\label{sec:alternativemarking}
In the previous sections, we have discussed a method of marking a
vertex through modifying the hopping potential from the perturbed site
to all three of its nearest-neighbors. Here, we will discuss
alternative methods of marking a vertex which still keep the perturber
interaction in the conic dispersion region of the spectrum. This can
be done by altering the hopping potential in different ways as
discussed above; this approach is necessarily a rank-2 perturbation
and these two perturber states must meet at the Dirac energy. We will
discuss several ways of applying such a perturbation in the next
paragraph. Another approach is based on coupling extra sites to the
lattice; this set-up will be discussed at the end of this section.
Focussing on hopping potential perturbations, we may consider
perturbing the bonds to any number of nearest-neighbors to any
strength. We have seen previously that a symmetric three-bond
perturbation successfully creates a search. Here, we demonstrate a
search based on perturbing the hopping potential from a given site to
only one of its nearest-neighbors. That is, we use the same search
Hamiltonian as in Eq.~(\ref{eq:full_search_Hamiltonian}) with the
perturbation matrix
\begin{equation} {\bf{W}} =
\ket{\alpha_{o},\beta_{o}}^{A}\bra{\alpha_{o},\beta_{o}}^{B} +
\ket{\alpha_{o},\beta_{o}}^{B}\bra{\alpha_{o},\beta_{o}}^{A}\, ,
\label{eq:singlebond_perturbation}
\end{equation}
and eigenstates
\begin{align}
\ket{W_{g}} &= \frac{1}{\sqrt{2}}\left(\ket{\alpha_{o},\beta_{o}}^{A} -\ket{\alpha_{o},\beta_{o}}^{B}\right) \\
\ket{W_{e}} &=
\frac{1}{\sqrt{2}}\left(\ket{\alpha_{o},\beta_{o}}^{A}
+\ket{\alpha_{o},\beta_{o}}^{B}\right).
\end{align}
Note that it no longer makes sense to speak of a single marked
vertex. Rather, our perturbation marks both vertices
$\left(\alpha_{o},\beta_{o}\right)^{A/B}$ simultaneously; thus, it may
be more accurate to view our perturbation matrix as marking a single
cell of the lattice.
For this type of perturbation, the avoided crossing used to generate
search behaviour is not necessarily at $\gamma = 1$, see
Figure~\ref{fig:singlebond_spectrum}, where the spectrum of our search
Hamiltonian as a function of $\gamma$ for a $12\times12$ cell torus is
shown. As stated before, our single bond perturbation is also a rank-2
matrix, and so again there are two perturber states approaching the
spectrum from the negative and positive regions of the
spectrum. Inspecting the region around the Dirac energy, see inset in
Figure~\ref{fig:singlebond_spectrum}, we see that avoided crossings
are formed by four states in total, the two blue (solid) curves and the two
red (dashed) curves. At $\gamma = \frac{1}{3}$ there is an exact crossing
between the red and blue curves indicating that these states are
orthogonal.
\begin{figure}
\begin{overpic}[width = 1.05\linewidth]{singlebond_spectrum}
\put(54.26,26.75){\fbox{\colorbox{white}{\includegraphics[width=0.45\linewidth]{singlebond_spectrum_zoom}}}}
\end{overpic}
\caption{(Color online) Spectrum of single-bond search Hamiltonian in
Eq.~(\ref{eq:full_search_Hamiltonian}) using perturbation from
Eq.~\ref{eq:singlebond_perturbation} as a function of $\gamma$ for
a $12\times 12$ cell torus ($N=288$). The spectrum is symmetric
around $\epsilon_{D} = 0$. Inset: States nearest the Dirac energy
have been coloured depending on their parity with respect to the
$C_{2}$ operator: even (solid blue), odd (dashed red), undefined (dot-dashed black).}
\label{fig:singlebond_spectrum}
\end{figure}
We can make this avoided crossing picture clearer by breaking down the
eigenstates of the search Hamiltonian in terms of the symmetries of
the lattice. In particular, we use the rotation operator $C_2$, which
is a rotation of $\pi$ about the mid-point of the perturbed
bond. Considering the action of $C_{2}$ on the marked
vertices and the eigenstates of the perturbation matrix, one finds
$C_{2}\ket{\alpha_{o},\beta_{o}}^{A/B} =
\ket{\alpha_{o},\beta_{o}}^{B/A}$, $C_{2}\ket{W_{g}} = -\ket{W_{g}}$
and $C_{2}\ket{W_{e}} = \ket{W_{e}}$. The inset of
Figure~\ref{fig:singlebond_spectrum} shows the states nearest to the
Dirac energy, where the same colour indicates the same $C_{2}$
parity. It now becomes clear that the states near the Dirac point
actually form two avoided crossings, one for each parity with a
minimum energy gap at $\gamma = \frac{1}{3}$.
\begin{figure}[t]
\centering
\includegraphics[width = 1.0\linewidth]{singlebond_search}
\caption{(Color online) Search on $12\times 12$ cell graphene lattice with
single-bond perturbation, using an optimal starting state. The
behaviour at both marked vertices is the same, as is the behaviour
at each of their neighboring sites.}
\label{fig:singlebond_search}
\end{figure}
Without going into details, we can again find optimal starting states
using a reduced Hamiltonian approach (one for each of the avoided crossings); the evolution of the system
using these optimal states is shown in
Figure~\ref{fig:singlebond_search} for a $12\times12$ cell lattice.
We find there is a significant localisation on the marked vertices and
their nearest-neighbors, with the probability of being found on
either of the marked vertices peaking at around $16-18\%$. We can
also see that the probability of being found on each of the
nearest-neighbors peaks at around $8\%$, resulting in a total
probability of being found on the marked vertices and their
nearest-neighbors of approximately $48\%$.
The success probability fluctuates, as the probability amplitude
oscillates between the marked vertices and their
nearest-neighbors. This is due to the probability amplitude being
constrained in the local area by the increased hopping potential
between the two marked vertices.
Although our demonstration is for a $12\times12$ cell lattice,
numerical investigations show that the search behaviour remains the
same as the lattice size increases with critical value fixed at
$\gamma = \frac{1}{3}$. We also find that the gap at the two avoided
crossings scales as $\mathcal{O}\left(1/\sqrt{N\ln N}\right)$, in the
same way as for the three-bond perturbation shown in
Figure~\ref{fig:3bond_spectrum}. As the search time is inversely
proportional to the energy splitting at the avoided crossing, it also
gives an estimate of the running time of the search $T =
\mathcal{O}\left(\sqrt{N\ln N}\right)$.
We now turn to coupling additional sites to the lattice as a way of
introducing a perturbation as also considered in \cite{Nice}; such a treatment
is in many ways closer to
experimental realisations as defects due to additional add-on atoms are
quite common in graphene \cite{Banhart}. We focus here on the
idealised case where an additional site is coupled to a single lattice
vertex. We use the perturbation matrix
\begin{equation} {\bf{W\left(\gamma\right)}} =
-\ket{\alpha_{o},\beta_{o}}^{A}\bra{site}\, -\,
\ket{site}\bra{\alpha_{o},\beta_{o}}^{A} +
\gamma\ket{site}\bra{site}\, ,
\end{equation}
for coupling the additional site $\ket{site}$ to the $A$-type vertex
$\left(\alpha_{o},\beta_{o}\right)^{A}$ and $\gamma$ is a free
parameter related to the on-site energy of the additional vertex. The
choice of the coupling terms is such that the binding energy between
the additional site and the lattice vertex is the same as the internal
couplings in the lattice. Thus, our search Hamiltonian is of the form
\begin{equation} {\bf{H}} = -{\bf{A}} + {\bf{W}\left(\gamma\right)}\,
.
\label{eq:additional_site}
\end{equation}
In Figure~\ref{fig:singlesite_spectrum}, the spectrum of the search
Hamiltonian is given as function of $\gamma$. As our free parameter
$\gamma$ only changes a single term, our spectrum only has a single
perturber state. One finds a clear avoided crossing around the Dirac
energy ($E=0$) when $\gamma =0$, that is, when the on-site energy of
the additional site matches the on-site energy of the lattice
vertices.
\begin{figure}
\begin{overpic}[width = 1.05\linewidth]{singlesite_spectrum}
\put(11.75,28.25){\fbox{\colorbox{white}{\includegraphics[width=0.45\linewidth]{singlesite_spearch}}}}
\end{overpic}
\caption{(Color online) Spectrum of the Hamiltonian in
Eq.~(\ref{eq:additional_site}) as a function of $\gamma$ for a
$12\times 12$ cell torus ($N=288$). The symmetry of the graphene
spectrum is broken by the choice of the perturbation in terms of
an additional site. Inset: Search on $12\times 12$ cell graphene
lattice with a single additional site perturbation, using starting
state $\ket{s}$ from Eq.~(\ref{eq:starting_state})}
\label{fig:singlesite_spectrum}
\end{figure}
In the reduced Hamiltonian picture, we write
Eq.~\ref{eq:additional_site} in terms of a basis consisting of the
Dirac states and our perturber state, $\ket{site}$. Only three states
are involved in the search, that is,
$\{\ket{\underline{K}}^{A},\ket{\underline{K}'}^{A},\ket{site}\}$, and
$\bf H$ reduces, up to a different prefactor, to the same $3\times3$
matrix found in Eq.~(\ref{redH}) for the three-bound
perturbation.
Thus, our previous reduced Hamiltonian analysis holds for this case
and the optimal search starting states are the same; the
time-evolution for the additional site search is shown in the inset in
Figure~\ref{fig:singlesite_spectrum}. The only differences between
this case and the three-bond marking are that the system localises on
the additional site, and also the change in prefactor leads to a
search time of $T = \frac{\pi}{4}\sqrt{N}$.
We note that other, more realistic types of perturbations involving
additional sites coupling to a lattice site and its neighbors, can be
shown to result in effective search
protocols.
Also, the single additional site perturbation described here can, like
the three-bond perturbation, be used to setup a communication
protocol, with the same signal patterns and behaviours as found
previously. The initial state is in this case completely localised on
the additional site.
\section{Alternative nanostructures}
\label{sec:alternativenanostructures}
So far we have described the dynamics of searches on graphene
lattices with periodic boundary conditions, that is, graphene on a
torus. Here we consider more realistic boundary conditions such
as nanotubes and graphene sheets.
For the dynamics on nanotubes, we move to periodic boundary condition
along one axis only and impose Dirichlet boundary conditions along the
other directions. The properties of nanotubes are well-known
\cite{Wong}, and it has been shown that the band structure of armchair
nanotubes, that is, nanotubes with armchair boundaries, always allows
for an energy at the Dirac energy regardless
of the nanotube diameter; we will focus on these types of nanotubes.
We are interested in searching on finite length nanotubes, where the
band structure becomes discrete; the length of the nanotube are in
addition chosen such that there exists an eigenenergy at the Dirac
energy. An example of the cell we use to construct finite armchair
nanotubes is shown in Figure~\ref{fig:armchairnanotube_example}. We
choose the finite length of the nanotube to be along the horizontal
axis and we close the underlying graphene lattice into a nanotube
along the vertical axis.
\begin{figure}[t]
\centering
\includegraphics[width =
0.8\linewidth]{armchair_nanotubes_example}
\caption{Example of a armchair nanotube cell. The nanotube is
periodic along the vertical axis and finite in the horizontal
direction with a width of $N_{x}$ sites. }
\label{fig:armchairnanotube_example}
\end{figure}
We denote the basis states of sites in our nanotube as
$\ket{m,A/B,l}$, where $m$ indicates the $m^{th}$ $A/B$-type vertex in
the horizontal direction in the $l^{th}$ cell. It is simple to see
that, due to the periodic boundary conditions along the circumference
of the tube, the vertical component of the eigenstates must be Bloch
states and the horizontal component of the amplitudes must be
sinusoidal in nature. That is, the eigenstates of the finite nanotube
are standing waves along its length.
By working through the tight-binding model for this system, one can
show that $\left(k_{x},k_{y}\right) = \left(\frac{2\pi}{3},0\right)$
is the only point where there exists an eigenenergy equal to the Dirac
energy. It is also possible to show that the spectrum includes Dirac
points when the number of sites along the length of the nanotube
$N_{x} = 3r -1$, where $r$ is an integer.
As there is only one potential Dirac point for finite armchair
nanotubes, it follows that there are only two Dirac states (one from
the bonding and the anti-bonding regions of the spectrum). As we have
$k_{y} = 0$ at the Dirac point, the Bloch wave around the
circumference of the nanotube is simply a uniform
superposition. Another important feature of the Dirac states on the
nanotube, which we have not encountered previously, is the existence
of nodal points where the amplitude of the eigenstate is 0. We find
these nodal points occur at every third site along the horizontal
axis.
We focus on applying our three-bond perturbation from
Sections~\ref{sec:threebondperturbation}~\&~\ref{sec:threebondperturbation_analysis}
to the armchair nanotube. In our new labelling our perturbation takes
the form
\begin{align}
W = &\ket{m_{o},A,l_{o}}\left(\bra{m_{o}+1,B,l_{o}} +
\bra{m_{o}-1,B,l_{o}} \right. \nonumber \\ &\left. +
\bra{m_{o},B,l_{o}}\right) + h.c.\, .
\end{align}
We have assumed that we are perturbing an $A$-type vertex on an even
horizontal coordinate, so that the perturbed site and its
nearest-neighbors remain within one nanotube cell. While it is easy
to re-write this perturbation matrix for other sites we restrict
ourselves to this form for simplicity. Note even in this form it can
still be expressed in terms of a marked site, $\ket{m_{o},A,l_{o}}$,
and a state which lives on the neighbors, $\ket{\ell} =
\frac{1}{\sqrt{3}}\left(\ket{m_{o}+1,B,l_{o}} + \ket{m_{o}-1,B,l_{o}}
+ \ket{m_{o},B,l_{o}}\right)$.
Our search Hamiltonian for the nanotube is the same as in
Eq.~(\ref{eq:full_search_Hamiltonian}). Inspecting the spectrum of the
search Hamiltonian as a function of the free parameter $\gamma$ where
the perturbation is not located on a nodal point of a Dirac state, one
finds an avoided crossing around the Dirac energy when $\gamma = 1$.
This spectrum is very similar to the one shown in
Figure~\ref{fig:3bond_spectrum} for the search on the torus and will
be
omitted.
If the perturbation is located on a nodal site the picture is
different, however; one finds that there is an exact crossing at the
Dirac energy when $\gamma = 1$ and searching is not
possible.
Recall that the effect of the perturbation at the critical point is to
completely remove the site from the lattice, if the site already has
zero amplitude then the perturbation will not interact with the Dirac
state.
Similar to the previous section, we numerically reduce the full search
Hamiltonian in a basis consisting of the two Dirac states and the
state living purely on the neighbors, $\ket{\ell}$. The perturbed
site, $\ket{m_{o},A,l_{o}}$, is decoupled from the lattice. Through
this process, we find two possible starting states, one for each
sublattice. The starting states are weighted superpositions of the
Dirac states over the sublattice containing the marked vertex. The
nodal points are excluded from this treatment.
\begin{figure}[t]
\centering
\includegraphics[width =
1.0\linewidth]{armchairnanotube_3p_bulk_search}
\newline \centering
\includegraphics[width =
1.0\linewidth]{armchairnanotube_3p_edge1_search}
\caption{(Color online) Searches on an armchair nanotube ($N=320$) with triple-bond
perturbation located in the bulk (upper figure) and placed near
the edge of the nanotube(lower figure), using numerically
calculated optimal starting state: $\ket{m_{o}-1,B,l_{o}}$ (dotted blue),
$\ket{m_{o}+1,B,l_{o}}$ (dashed red), $\ket{m_{o},B,l_{o}}$ (dot-dashed green), sum
of neighbor probabilities (solid black).}
\label{fig:armchair_search}
\end{figure}
Figure~\ref{fig:armchair_search} shows the system evolution for two
searches using the numerically found optimal starting states. One
search has the marked site located in the centre of the nanotube, the
other has the marked site positioned near the edge of the nanotube.
The dimensions of the nanotube have been chosen so that there are 320
sites, comparable to the searches shown in earlier sections. Comparing
these searches to the three-bond perturbation search on the torus in
Figure~\ref{fig:search_prob}, we see that there is a marked difference
in behaviour and success probability induced by relaxing the periodic
boundary conditions along one axis.
Using a perturbation located near the edge, the lower image in
Figure~\ref{fig:armchair_search}, we see a reduction in success
probability by a factor of 2-3 and strong fluctuations in peak height
when compared to the search on the torus. Searching in the bulk,
shown in the lower image of Figure~\ref{fig:armchair_search}, displays
behaviour closer to searches on a graphene torus, but again with
significant fluctuations at each peak. We propose that this effect is
due to the reflection of probability amplitude from the edges of the
nanotube. This is supported by the changes in the interference pattern
in the signal as the perturbation is moved across the lattice.
Numerics demonstrate that additional site perturbations and
communication protocols can also be used on nanotube
lattices.
We now move to working on graphene sheets, that is, removing the periodic boundary conditions along both axis. A detailed account of a specific version of this set-up together with experimental results has been presented in \cite{Nice} for graphene sheets consisting of armchair boundaries only. Note that such a configuration can not be achieved on rectangular sheets; the simplest configuration has the form of a parallelogram, see \cite{Nice} for further details.
The advantage of such a configuration is that the form of the boundary does not admit so-called `edge-states'. In the following, we will look at the influence of these edge-states in more detail by considering rectangular graphene sheets, such as in Figure \ref{fig:finite_lattices_example}. In addition to armchair edges, here along the vertical axis, these sheets have boundaries formed by bearded edges (left) and zigzag edges (right) along the horizontal axis. These boundaries support edge states, i.e.\ states localised along these edges, with an eigenenergy close to the Dirac energy \cite{Wakabayashi, Yao}. In the following, we will investigate, how the existence of these edge states influences the search.
\begin{figure}[t]
\centering
\includegraphics[width =
1.0\linewidth]{beardedandzigzag_lattice}
\caption{Examples of two finite graphene sheets, with dimensions in
terms of primitive cells $\left(N_{x},N_{y}\right) =
\left(4,4\right)$. Along the vertical axis of both sheets are
armchair edges. The horizontal boundaries are formed by bearded
edges (left) and zigzag edges (right). }
\label{fig:finite_lattices_example}
\end{figure}
As in the case of the finite armchair nanotube, the
imposition of Dirichlet boundary conditions at an edge generates
sinusoidal eigenstates. As a result, there are no extended (bulk) eigenstates at the Dirac energy due to the inability to
equate the quantised momenta with the necessary points in
$\underline{k}$-space. We thus need to find other extended eigenstates in the sea of edge states
near the Dirac point to undertake a search in this set-up.
We mark sites using the triple-bond perturbation such as in
Eq.~(\ref{eq:full_search_Hamiltonian}). Throughout this section we
choose the dimensions in terms of primitive cells of the graphene
sheets, $\left(N_{x} , N_{y}\right) = \left(10, 10\right)$; a bearded
edge sheet thus consists of $N = 200$ sites and a zigzag sheet is
formed of $N = 218$ sites. The spectra of the search Hamiltonians for
the bearded lattice is shown in Figure~\ref{fig:spectrum_bearded_3p}
constrained to the energy region of interest. The spectrum for zigzag
edges is very similar and is not show here. One can clearly see an
avoided crossing around the Dirac energy at $\gamma = 1$, the critical
value for this type of perturbation. There are several states very
close to the Dirac energy; these are the edge states which are all
non-degenerate. Therefore, it is not possible for us to construct a
superposition of degenerate eigenstates which is optimal for
searching. Rather, our initial starting state must be a single
eigenstates of the unperturbed Hamiltonian $\mathbf{H}_{o} =
-\mathbf{A}$.
\begin{figure}[t]
\centering
\includegraphics[width = 1.0\linewidth]{spectrum_bearded_3p}
\caption{(Color online) Spectrum of triple-bond perturbation search Hamiltonian in
Equation~\ref{eq:full_search_Hamiltonian} as a function of $\gamma$ for a
$10\times 10$ cell bearded graphene sheet (N = 200). We focus only
on the relevant section of the spectrum.}
\label{fig:spectrum_bearded_3p}
\end{figure}
Using the search Hamiltonian from
Eq.~(\ref{eq:full_search_Hamiltonian}) and fixing $\gamma = 1$, we
proceed by allowing the system to evolve after being prepared in one
of the unperturbed eigenstates. One finds for both types of lattice,
that the edge states near the Dirac energy fail to produce
localisation behaviour; the probability at the neighboring vertices
of the marked site do not rise above noise levels. Rather, search
behaviour only begins to emerge when we use the first delocalised,
non-edge state as our initial state. We also note that we
\emph{cannot} search for sites near the zigzag or bearded edges, where
the edge states exist. Only as we move further into the bulk of the
lattice or along an armchair-type edge, the localisation behaviour
returns.
\begin{figure}[t]
\centering
\includegraphics[width =
1.0\linewidth]{search_bearded_0xhalfy}
\includegraphics[width =
1.0\linewidth]{search_bearded_central}
\includegraphics[width =
1.0\linewidth]{search_bearded_midxthirdy}
\caption{(Color online) Searching on a bearded edge graphene sheet with dimensions
$\left(N_{x},N_{y}\right) = \left(10, 10\right)$ using the
Hamiltonian, Eq.~(\ref{eq:full_search_Hamiltonian}). All the
searches are initialised in the first unperturbed, non-edge
eigenstate above the Dirac energy. The locations of the marked
vertices are: upper - halfway along the left armchair edge, middle
- centre of the sheet, lower - at the mid-point of the lower
bearded edge and a third of the way along the vertical axis. (Colors and linestyles as in Figure
\ref{fig:armchair_search}) }
\label{fig:bearded_searches}
\end{figure}
Note that the edge states can be viewed as a kind-of one-dimensional
system and, as we saw from the scaling argument towards the end of
Sec.~\ref{sec:quantumwalksandsearching}, one-dimensional systems imply
an energy spacing between successive energy levels of $E_{n+1}- E_{n}
= \mathcal{O}\left(N^{-1}\right)$. Thus, the perturber state interacts
with many states in a dense part of the spectrum and the search fails.
In Figure~\ref{fig:bearded_searches} we show the search behaviour
arising from marked sites placed at different positions on a bearded
graphene lattice (zigzag lattice types display similar behaviour).
The results are similar to those found for the nanotube searches.
There is a slight increase in variation of signal pattern with
position, and an increase in success probability maxima as we move
towards the centre of the lattice from either of the bearded edges.
\section{Discussion}
\label{sec:discussion}
We have shown that continuous-time quantum search can be done
effectively on a two-dimensional graphene lattice without the use of
internal degrees of freedom. This is achieved by making use of the
conical (linear) dispersion relation in the graphene spectrum. The
search succeeds in time $ T = \mathcal{O}\left(\sqrt{N\ln N}\right)$
with probability $\mathcal{O}\left(1/\ln N\right)$. This is the same
time complexity found in \cite{Aharonov} for discrete-time searches
and in \cite{CG04a} for continuous-time searches. To boost the
probability to $\mathcal{O}\left(1\right)$, $\mathcal{O}\left(\ln
N\right)$ repetitions are required giving a total time $T =
\mathcal{O}\left(\sqrt{N}\ln^{\frac{3}{2}}N\right)$. Amplification
methods \cite{Gro97a,Tul08,PGKJ09} may be used to reduce the total
search time further.
Our main result focusses on perturbations which involve altering the
hopping potential from a marked site to all three of its
nearest-neighbors equally. We have also demonstrated other types
searches based on perturbing the hopping potential in a cell and the
adding extra sites. We have shown that search mechanisms can be
utilised for the purposes of signal transfer.
Our findings point towards applications in directed signal transfer,
state reconstruction, or sensitive switching. This opens up the
possibility of a completely new type of electronic engineering using
single atoms as building blocks of electronic devices. Our results
demonstrate that a range of nanostructures constructed from graphene
could be used to this end.
\acknowledgements
We thank Julian B\"{o}hm, Ulrich Kuhl and Fabrice Mortessagne for discussions on how the ideas of \cite{FGT} could
be adapted to more physical systems, which lead to some of the work on additional perturbation types and graphene
flakes.
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.